PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINE return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE ma campus-p.14 ENCAPSULATION OF METAL SULFIDE AND OXIDE CLUSTERS IN Mos2 AND ws2 AND CHARACTERIZATION OF RESTACKED Mos2 AND ws2 By Joy Marie Heising 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 ENCAPSULATION OF METAL SULFIDE AND OXIDE CLUSTERS IN Mos2 AND ws2 AND CHARACTERIZATION OF RESTACKED Mos2 AND WS2 By Joy Marie Heising The synthesis Of [A1,3O4(OH)24(HZO),2],‘MS2 (M = MO,W; x = 0.02 - 0.05; d spacing = 16.0A) was accomplished by a precipitative encapsulation technique in which a suspension Of single MS2 layers in H20 was added to an aqueous solution of the [A1,3O4(OH)24(HZO),2]7* cluster. One- dimensional electron density mapping and Rietveld refinement revealed that the cluster is oriented with its ”C3 axis perpendicular to the layers. Thermal analysis of the samples indicate that the layers remain expanded up to 100°C and partially expanded to 300°C, but surface area measurements reveal that void space (if any) in the material is inaccessible. The synthesis of [Fe688(PEt3)y],‘MS2 (x = 0.0502; y = 2-3; d-spacing = 10.5 -1 1.4 A ) and [Ni989(PEt3)6]xMSZ (x = 0.04-0.077; d-spacing = 14.0 - 14.2 °) is similar but a solution of the cluster in a non-aqueous, water miscible solvent is used. [C0688(PPh3)6]0mMoS2 (d-spacing = 15.3 A) was synthesized by the addition of a solution of the CoéSs(PPh3)6 cluster in CHzCl2 to an aqueous suspension Of single MOS2 layers. Thermal analysis reveals that the volatile product in all samples is phosphine sulfide. The hydrodesulfurization (HDS) activity of [C0688(PPh3),.«,]o.02MOS2 is comparable tO that Of commercially available catalysts; the activity Of [N iS,SS,(PEt3)6]xMS2 is inferior. The structure of restacked MS2 (M = MO,W) was solved by electron crystallographic studies. Using two-dimensional hk0 data indexed to a 3a x a orthorhombic unit cell, Patterson projections were calculated which revealed a severe distortion in the Mo/W plane, forming infinite zig-zag chains. The projection of the structure suggests M-M distances of 2.92 A and 2.74 A for MOS2 and W82, respectively. Least squares refinement from the single crystal data gives R,=13.3% for WS2 and R1=15.3% for M052, and reveals a WTe2 type structure. The relationship between charge and structure in restacked MS2 has been probed by encapsulation of alkali cations, forming A,,(HZO)YMS2 (x = 0.15-0.25; y = 0.3045) and chemical oxidation with I2 and Brz. Chemical oxidation results in a change in the structure of restacked M082, giving rise to a V51 x 1/51 superlattice. Differential Scanning Calorimetry studies show an irreversible exothermic transition to 2H-MS2 which shifts in temperature with oxidation. Thermopower measurements indicate that restacked MOS2 and WS2 are p-type metallic conductors. ACKNOWLEDGMENTS Nothing presented here would have been possible without the support, enthusiasm, patience, and guidance Of my advisor, Professor Mercouri Kanatzidis, for which I am deeply grateful. I would also like to thank Professor Pinnavaia for his helpful advice, for allowing me to use his equipment, and his comments as a second reader. Much thanks to Professor McCracken and Professor Baker for serving as members Of my graduate committee as well. I thank Dr. Chris Marshall and Dr. Jim Brenner for HRTEM and HDS studies, Dr. Eric Prouzet for EXAFS measurements, and Dr. Duck-Young Chung for thermopower measurements. I owe a great deal to Dr. John Heckman, who taught me to use a TEM and Offered many useful suggestions throughout my studies. I would also like tO thank the members Of the Kanatzidis group for their kindness and all that I have learned from them, especially Rabin Bissessur, Wakgari Hirpo, Hui-Lien Tsai, Lei Wang, Jason Hanko, and Rhonda Patschke. Much thanks also to the chemistry department staff. Thanks to my friends, especially Dean Lantero and Carl Iverson for help with the NMR and Gary Lavine for just listening. Finally, I thank my family for their support. iv TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES ABBREVIATIONS CHAPTER 1 Introduction 1.1 Insertion Reactions 1.2 Open Framework Materials 1.3 MOS, and WS, as hosts 1.4 Pillaring agents References CHAPTER 2 viii ix xiv Encapsulation and Rietveld Structural Characterization of the [AI,,O4(OH),.,(H,O),,]7+ cluster into MOS, and WS,. 2.0 Abstract 2.1 Introduction 2.2 Experimental 2.2.1 Synthesis 2.2.2 Characterization 2.3 Results and Discussion 2.3.1 Evidence Of Cluster Encapsulation 2.3.2 One-Dimensional Electron Density Mapping 2.3.3 One-Dimensional Rietveld Refinement 2.3.4 Thermal Stability and Absorption Measurements 2.4 Conclusions References CHAPTER 3 Encapsulation of Neutral and Cationic Metal Chalcogenide Clusters in MOS, and WS, 3.0 Abstract 3.1 Introduction 3.2 Experimental 3.2.1 Synthesis 3.2.2 Characterization 3 5 36 37 38 38 40 41 41 49 57 60 63 65 68 69 71 72 72 7S 3.3 Results and Discussion 77 3.3.1 Evidence Of Cluster Encapsulation 77 3.3.2 Thermal Analysis 90 3.3.3 HDS Activity 101 3.4 Conclusions 109 References 1 13 CHAPTER 4 Structure of Restacked MOS, and WS, Elucidated by Electron Crystallography 1 1 7 4.0 Abstract 118 4.1 Introduction 119 4.2 Experimental 124 4.3 Results and Discussion 125 4.3.1 Data Collection 125 4.3.2 Structure solution from the “triplet” crystal of WS, 128 4.3.3 Dynamic Scattering: Two-Beam Approximation 136 4.3.4 Structure solution from “single” crystals of WS,. 146 4.3.5 Dynamic Range Correction 153 4.3.6 Structure Solution from “single” crystals Of MOS,. 155 4.3.7 Dynamic scattering: n-beam approximation. 160 4.3.8 Electron Crystallographic Studies of an alternate superlattice in MOS, 168 4.4 Conclusions 174 References 175 CHAPTER 5 Exfoliated and Restacked MOS, and WS,: Ionic or Neutral Species? Encapsulation and Ordering Of Hard Electropositive Cations l 7 9 5.0 Abstract 180 5.1 Introduction 181 5.2 Experimental 184 5.2.1 Synthesis 184 5.2.2. Characterization. 191 5.3 Results and Discussion 192 5.3.1 Cation encapsulated MS,. 192 5.3.2 Electron diffraction studies of (cation),MS,. 194 5.3.3 Modeling the cation ordering. 202 5.3.4. Oxidation state of M (M = MO,W) and the charge of the layers. 209 5.3.5 Oxidation reactions of LiMS,. 212 5.3.6 X-ray diffraction studies of oxidized MS,. 218 vi 5.3.7 DSC studies Of oxidized M8,. 225 5.3.8 Acid restacked MS,. 231 5.3.9 Thermopower measurements. 237 5.4 Conclusions 241 References 246 vii LIST OF TABLES Table 2.1. IF(l)obsl and F(1)calc for [Al,,O4(OH),4(H,O)1,]0_05MOS,. 52 Table 2-2 IF(l)obsl and F(1)calc for [N1304(OH)24(H20)1210.055W82- 53 Table 2.3. Rietveld refinement summary Of [M13O4(OH)24(H20)12]o.ossws2 59 Table 3.1. HDS catalytic activity and surface area (SA) of selected samples and references. 103 Table 4.1. Crystal data and structure refinement for restacked WS,. 134 Table 4.2. Preliminary atomic coordinates ( x 104) and equivalent isotropic displacement parameters (A2 x 103) for restacked WS,. 135 Table 4.3. Refinement for WS, as a function of data correction methods. 141 Table 4.4. Crystal data and structure refinement for crystals of WS,. 150 Table 4.5. Atomic coordinates ( x 104) and equivalent isotropic displacement parameters (A2 x 103) Of restacked “single” and “triple” WS,. 15 1 Table 4.6. Crystal data and structure refinement Of restacked MOS,. 158 Table 4.7. Atomic coordinates ( x 10‘) and equivalent isotropic displacement parameters (A2 x 103) in restacked MOS,. 159 Table 4.8. A comparison Of simulated dynamic electron diffraction data to the experimental data for MOS, and WS, as a function Of layer thickness. 167 viii LIST OF FIGURES Figure 1.1. The chemical insertion of Li atoms into TaS,. 2 Figure 1.2. Projection of 2H-MOS2 (A) in the ab plane (one layer) and (B) along the c axis. 2H-WS, is isostructural. 11 Figure 1.3. Structure Of lT-TiS, viewed (A) in the ab plane and (B) along the c axis. 13 Figure 1.4. Illustration Of the exfoliation and flocculation process used to encapsulate neutral and cationic species. 15 Figure 1.5. (A) the [A111,O,,(OH),,,(H,O),,]7+ cluster. (B) A tn'mer Of edge- sharing octahedral Al” ions with a common vertex (Open circle). 19 Figure 1.6. (A) MGQ, core of Co,S,(PPh,)6. (B) M6Q8 core with phosphine ligands. 21 Figure 1.7. (A) Ni3S3 units which are stacked in a staggered fashion to form (B), the [Ni,S,(PEt3)6] 2* cluster. 24 Figure 2.1. X-ray diffraction patterns of (A) [A1,3O,,(OH),4(H,O)1,],05MOS2 and (B) [A11304(OH),4(H,O),,]0.055WS, used for electron density maps and Rietveld refinements. 42 Figure 2.2. Three possible orientations of the [Al,,O4(OH),4(H,O),,]7* cluster between the layers. 43 Figure 2.3. Solid State MAS-NMR 27Al spectra of (A) [N1304(0H)24(H20)1210.03M052 and (B) Na[A11304(0H)24(H20)121600246 Figure 2.4. HRTEM Micrographs Of (A) [A1,,O,(OH),,(H,O),,]0_03MOS, and (B) [A]1304(OH)24(H2O)12]O.04WS2- 47 Figure 2.5. One-dimensional electron density maps of (A) [Al1304(OH),4(H,O)1,]005MOS, and (B) [A]1304(OH)24(H20)1210.055W52- 54 Figure 2.6. (A) Electron density maps calculated from [A1,,O,,(OH),,,(H,O)1,]0055WS2 F(1)obs data using 12-19 001 reflections (a-h). (B) Maps calculated from theoretical data for a) a W atom and b) a W atom and a S atom, illustrating the Fourier truncation error and the dynamic range problem. 56 Figure 2.7. (A) TGA of [Al,,O,(OH),4(H,O)1,],03MOS, under nitrogen, showing gradual weight loss with an inflection point near 300°C. (B) XRD patterns of the material (a) at room temperature (b) after heating to 120°C and (c) after heating tO 300°C. 62 Figure 3.1. X-ray diffraction patterns of (A) [C068, PPh3)6]O,4MOS, and restacked MOS,, showing an expansion of 15.4 upon encapsulation of the cluster (B) [C0688(PPh,)6]0'96MoS,, showing only 4 001 reflections. 78 Figure 3.2. The C068,,(PPh3)6 cluster oriented with its (A) C3 axis perpendicular to the layers and its (B) C4 axis perpendicular to the layers. 79 Figure 3.3. HRTEM micrographs of [C0688(PPh3)x]0.02MOS,. (A) Region with larnellar features showing expanded and unexpanded layers. Expansion suggests that PPh, ligands have been removed. Small (B) Region showing in-plane structure. 80 Figure 3.4. X-ray diffraction patterns of (A) [Fe6S,(PEt3),]O_096MOS, (A = 10.5 A) and (B) [FeéS,(PEt3),]0,0WS, (A = 11.4 A). 84 Figure 3.5. HRTEM micrograph Of [Fe6Ss(PEt3)3]o,0WS, showing highly ordered in-plane structure. 85 Figure 3.6. X-ray diffraction patterns Of (A) [N1939(PEI3)6]0.07MOS, (A = 14.2 A) and (B) [Ni,s,(PEt,),]o,,,ws, (A = 14.0 A). 88 Figure 3.7. A view of the [Ni,S,(PEt3)6]2* cluster with its Ni,S3 planes (A) parallel to and (B) perpendicular to the MS, layers 89 Figure 3.8. TGA Of (A) CO6S,(PPh3)6. (B) [Co,,S,(PPh,)6]0,,MoS,. 91 Figure 3.9. An illustration of the original plan to create a microporous sulfide: encapsulation Of the cluster with phosphine ligands acting as “spacers”, followed by thermal treatment to remove the phosphine ligands. 92 Figure 3.10. An illustration of the observed data: encapsulation Of the cluster with phosphine ligands acting as “spacers”, followed by thermal treatment. The phosphine ligands attack the layers as they depart, forming SPR3 species and possible 82' anion vacancies. 93 Figure 3.11. X-ray diffraction patterns upon thermal treatment of [C0638(PPh3)6]0,4MOS,. 94 Figure 3.12. TGA of (A) [Fe6S,(PEt3)6](BPh4),, (B) [F6633(PEI3)2]0.096MOS, (solid line) and [Fe6S,(PEt,)3]0_,OWS, (dashed line). 97 Figure 3.13. X-ray diffraction patterns upon thermal treatment of [F66$8(PEt3),]o'096MOS,. 98 Figure 3.14. TGA of (A) [Ni,S,(PEt,)6](ClO4),. (B) TGA of [Ni,S,(PEt,)6]OMWS,. 100 Figure 3.15. HRTEM micrograph of [C068,(PPh3)6]0'0,MOS, after HDS. 105 Figure 3.16. Illustration of the “rim-edge” model for (A) unpillared and (B) pillared MOS,. 107 Figure 4.1. Schematic illustrating the proposed superstructures Of restacked/ 1T-MOS,. 121 Figure 4.2. BFI and SAED Of “triple” and “single” WS,. 126 Figure 4.3. (A) Relationship between the 2a x a and 1/371 x a lattice. (B) Illustration Of how three flu x a patterns can be overlapped to form a 2a x 2a pattern. 127 Figure 4.4. Illustration of the “Gel Plotting Macro” in NIH Image 1.60. 130 Figure 4.5. Two-dimensional Pattersons along the c-axis: (A) lT-MS, (theoretical), (B) restacked WS, with background subtraction and Lorentz correction, (C) restacked WS, with only Lorentz correction, (D) restacked WS, with no corrections. 131 Figure 4.6. Schematic illustrating (A) kinematic scattering and (B) dynamic scattering. 137 Figure 4.7. Graph Of the function K(A,Q). 139 Figure 4.8. The intersection Of the Ewald sphere with the reciprocal lattice at an accelerating voltage Of (A) ~50kV and (B) ~100kV. 143 Figure 4.9. (A) The limit Of detectable reflections (in a perfectly oriented crystal) as a function Of the thickness (t) Of the sample (q = 1/t), the radius Of the Ewald sphere (r), and angle 7. (B) An illustration Of secondary scattering. 144 Figure 4.10. Selected Area Electron Diffraction of “single” WS, patterns used to generate Patterson contour plots. 148 Figure 4.11. Two-dimensional Patterson projections along c-axis: (A) ideal MS, and (B-D) three “single” WS, data sets. 149 Figure 4.12. (A) Two-dimensional projection Of a restacked MS, layer and one WTe, layer. (B) View Of WTe, parallel to c-axis. 152 Figure 4.13. Surface plots Of the (3-10) and the (-2-20) reflections from the patterns in Figures 4.2d and 4.11a. 154 Figure 4.14. Bright Field images and Selected Area Diffraction Patterns of “single” crystal restacked MOS,. 156 Figure 4.15. Two-dimensional Patterson projections Of restacked MOS,. 157 Figure 4.16. Simulated dynamic Electron Diffraction patterns of restacked WS,. 162 Figure 4.17. Patterson projections calculated from the data in 4.16. 165 Figure 4.18. Simulated Patterson projections for WS, calculated as a function Of thickness. 166 Figure 4.19. Selected Area Electron Diffraction pattern Of alternate MOS, superlattice. 171 Figure 4.20. Two-dimensional Patterson projections calculated for (A) an ideal MOS, lattice and (B) from experimental data 172 Figure 4.21. Simulated Patterson projections of alternate MOS, superlattice. 173 Figure 5.1. (A) X-ray diffraction pattern Of Nao‘,8(H,O)0.45WS,. (B) TGA Of Nao.18(H20)0.45W82 ’ Ko.21(H20)o.4oW52, and Rbo.24(H20)o.. 193 Figure 5.2. Selected Area Electron Diffraction patterns of (A) CSO_,3MOS, and (B) CstS, 196 Figure 5.3. Selected Area Electron Diffraction patterns Of (A) K0_,,MOS,; (B) K0,,MOS,; (C) Rb0.,5MOS,; and (D) K0,,MOS,. 199 Figure 5.4. Selected Area Electron Diffraction patterns from (A) a thick crystal Of LiXWS, and (B) a thin crystal Of waMOS, 200 Figure 5.5. (A) Structural model and (B) simulated electron diffraction pattern to explain the Observed patterns in Figures 5.2-5.3. 203 Figure 5.6. Selected Area Electron Diffraction patterns Of (A) BamMOS, and (B) NaOMMOS, 207 Figure 5.7. (A) Structural model and (B) simulated electron diffraction pattern to explain the Observed Selected Area Electron Diffraction patterns in Figure 5.5. 208 Figure 5.8. Selected Area Electron Diffraction pattern Of LiMoS, oxidized with Br, in MeCN. 214 Figure 5.9. (A) Patterns which show only the alternate superlattice. (B) Selected Area Electron Diffraction pattern of LiMoS, oxidized with Br, in MeCN over a longer period Of time with sonication. 215 Figure 5.10. Selected Area Electron Diffraction pattern Of a minority phase formed upon oxidation of LiWS, with Br, in MeCN. 216 Figure 5.11. Transmission X-ray diffraction (XRD) pattern Of LiMoS, oxidized in MeCN with increasing time. 219 Figure 5.12. Transmission XRD pattern of LiWS, oxidized with Br, in MeCN. The black circle marks a peak at 2.55 A which does not appear to be an hk0 reflection Of any Of the three ex ected lattices. The asterisk marks a contribution from the Saran wrap substrate. 220 Figure 5.13. Transmission XRD pattern Of LiWS, oxidized with Br, (A) before and (B) after heating. 224 Figure 5.14. Differential Scanning Calorimetry (DSC) plots of restacked MOS, and LiMoS, oxidized with Br,. 226 Figure 5.15. DSC plots showing the irreversible exothermic phase transition to 2H-WS,. 227 Figure 5.16. Transmission XRD pattern Of (A) LiWS, and (B) HXMOS, Obtained from LiMoS, exfoliated in concentrated HCl. 232 Figure 5.17. DSC plots Of H,WS, Obtained from the reaction Of HCl with (A) LiWS, (155°C) and Li,,n(H,O),WS, (191°C). 233 Figure 5.18. Schematic illustrating (A) an ideal lT-MS, lattice, (B) a flu x a lattice with infinite zig-zag metal chains, (C) a 1/3a x I/Ba lattice with trimers, and (D) a 21/361 x 21/3a lattice with possible tetramers formed due to M-M distortions. 236 Figure 5.19. Schematic band diagram illustrating the distortion of an ideal octahedral system to form zig-zag chains. 239 Figure 5.20. Thermopower measurements Of restacked MOS, and WS,. 240 Figure 5.21. Reaction scheme illustrating the different phases Obtained with various treatment Of (A) MOS, and (B) WS,. 243 ABBREVIATIONS BET:Brunauer, Emmett, and Teller (Surface Area) BFI: Bright Field Image DSCzDifferential Scanning Calorimetry EDS:Energy Dispersive Spectroscopy EXAFS:Extended X-ray Absorption Fine Structure HDS:HydrOdesulfurization HRTEMzHigh Resolution Transmission Electron Microscopy MAS-NMR: Magic Angle Spinning Nuclear Magnetic Resonance SAED:Selected Area Electron Diffraction SEM:Scanning Electron Microscope TEM:Transmission Electron Microscope TGA:Thermogravimetic Analysis XRD: X-ray Diffraction PEt3: triethyl phosphine PPh,: triphenyl phosphine A1130“: [M1304(0H) 24(H20) 12]” xiv CHAPTER 1 Introduction 1.1 Insertion Reactions The demand for materials with new and unique combinations Of properties has required scientists to adopt many clever synthetic strategies. One approach is tO alter the properties Of two existing materials by combining them. For example, the combination of nylon—6 with clays results in nanocomposites with enhanced thermal stability and mechanical properties relative tO those Of the individual components.’ Materials which are mixed on a molecular level are of particular interest to chemists, and can be synthesized through processes such as encapsulation and insertion. In insertion or encapsulation reactions, a guest species is incorporated into a host species, resulting in a net gain in the energy Of the system. The “host” is usually a one-dimensional, two-dimensional, or three-dimensional materialz'3 and the “guest” includes molecular species which range from a proton4 to large clusters92 or polymers.‘5'16 An example is the chemical or electrochemical insertion of Li atoms into the layered compound TaS,:5 n-butyl Lithium Li Li Li Li Li Li ti Li Li Li Li ti l I l' l' l' !' I' l' Figure 1.1. The chemical insertion Of Li atoms into TaS,. An insertion is said to be “topotactic” or “topochernical” when the structure Of the host framework is not significantly altered by the insertion process. Layered materials have been explored perhaps most extensively as hosts because their structure is sufficiently robust to allow topotactic insertion, but flexible enough to allow species of many different sizes tO be inserted. Intercalation, a special kind of insertion process, is a very popular example of host/guest chemistry. In the most general chemical sense, “intercalation” refers to a reversible insertion. The word “intercalation” is sometimes loosely employed in the literature to describe irreversible insertion reactions as well. The driving force for the insertion process is the net gain in energy Of the system, which can be realized through a variety Of mechanisms. One Of the most common ways is through electron transfer from the guest species to the host species. For example, in the insertion Of the Li atoms into TaS, the layers are reduced and the lithium atoms are oxidized.5 Another frequently encountered insertion mechanism is ion-exchange, which has been used to replace gallery N a” ions in smectite clays with other cations such as tetramethyl or tetraethylammonium.6 This is accomplished by swelling the clay with water, then exposing it to a solution with an excess Of the cation which is to be intercalated. A third manner is through coordination of the guest species to the host, which is Observed when alcohols replace coordinated water in vanadyl phosphonates.’ Other methods exist, and some layered hosts, such as V,Os, can undergo insertion through more than one method.8 The variety in mechanisms through which guest species can be inserted, coupled with the large number of known layered structures, has resulted in an enormous body of literature on the subject Of the insertion Of guest species in lamellar materials.2'3"9'10 The earliest reports of intercalation involve graphite as a host material.‘l Layered hosts which can undergo insertion reactions also include smectite clays”; double hydroxides”; metal oxidesg'”, oxyhalides‘s, and halides9; transition metal phosphates, phosphonates, and sulfophosphonates”; and transition metal dichalcogenides.3'5'lo Even the high temperature superconducting cuprates can undergo insertion reactions.‘4 The diversity of available layered hosts and guest species has led tO the study of these systems for many applications. One of the most important applications which has been explored for those systems which undergo insertion via redox reactions is their use as solid state battery cathodes. In particular, the insertion of alkali metals in the transition metal dichalcogenides both chemically and electrochemically has been studied extensively due to the interesting electronic properties of the chalcogenides, which can be affected by the intercalation reaction.”lo Related to this is the study of polymers as guest species, desirable because they lead to an increase in ion mobility in the lattice. Studies have not been limited to the layered dichalcogenides, but include hosts such as clays, V,O,, FeOCl, and M003. Strategies to incorporate polymers have included insertion Of the monomer followed by polymerizationls or direct incorporation of the polymer.16 Another important application for layers which undergo insertion processes is as ion-exchange materials17 or open framework catalysts.18 The layers in smectite clays can be separated by solvation, called swelling, which allows ion exchange.19 When larger cations are used, void spaces may be created.6 1.2 Open Framework Materials The demand for open framework structures, which originates from their application as size selective molecular sieves and catalysts, has inspired the discovery and characterization of a vast array of materials in the last fifty years. Those with accessible void spaces in the range 3 - 20 A are called microporous; compounds with larger void spaces (up tO 500 A) are mesoporous. The most widely used Open framework materials are the aluminosilicate zeolites, which possess three dimensional structures with pore sizes in the range 3-9 A”, and zeolite-like aluminophosphates, which contain micropores up to 14 A.21 Applications for zeolites include ion exchange17 and a variety of size selective catalytic processes.22 Zeolites with larger void spaces are desirable but have not been discovered. A breakthrough in the synthesis of mesoporous materials was the discovery Of MCM-41 in 1992. The synthesis involves the polymerization Of silicate around micelles Of long-chain alkylammonium cations, followed by calcination, and results in a material which has uniform channels with accessible void space.23 The dimensions Of these channels can be controlled synthetically by adjusting the micelles, allowing the pore size tO range from 16 - 100 A. Prior to MCM-41, one strategy for attaining pore sizes larger than 14 A was the “pillaring” of smectite clays with large cations, creating materials with accessible void space. Typically void space is quantified through surface area values measured by N, absorption. The exchange of small gallery cations for larger cations requires an expansion of the layers, and is possible in smectites due tO their ability to swell and even delaminate in water. The first example of pillaring was reported by Barrer and MacLeOd in 1955, and involved the exchange Of Na+ for tetramethyl and tetraethyl ammonium in smectite clays.6 These cations are small pillars, however, and interest in pillared clays only began to develop in the 1970’s, particularly with the discovery that pillaring with polynuclear metal hydroxyl cations such as A113O4(OH),4(H,O),,7‘° and Zr,,(OH),,“+ results in materials with thermal stability above 500°C and moderately high surface area values.77 These pillars convert to metal oxide clusters upon calcination, resulting in a material which retains the layer expansion. Other pillaring agents have included bicyclic amine cations“, metal and silicon chelate complexes”, metal halide clusters such as Nb6C1,,2*,'8" and imogolite, a large tubular structure.26 Unfortunately Open framework silicates are insulators. The chalcogenides, however, usually have semiconducting or metallic properties, which, when combined with porosity, could result in a unique new class Of materials with interesting practical applications which include use as chemical sensors or catalysts. For this reason many researchers are using a variety Of techniques in an effort to synthesize Open framework chalcogenides.27 A logical approach is hydrothermal synthesis, as it is the method used to make zeolites. An important breakthrough came in 1989, when Bedard and coworkers discovered A,Sn3S7 (A = alkyl ammonium cation) and A,MGe,,Sw.2“'34 Perhaps the most exhaustively studied open framework chalcogenide is A,Sn3S,, which has a layered structure with large pores in the basal plane. The solution precursors have been isolated and a mechanism Of formation proposed.29 The effects of progressive selenium substitution and included water upon the band gap have been reported.30 Unfortunately, adjustment of the layer stacking to minimize the void space in the basal plane occurs when organic cations Of various sizes are used in the synthesis.31 The tremendous flexibility Of the framework seems tO confound efforts to maximize void space in the structure. The materials are only stable to 250-350°C (depending on the identity Of the organic cation), and microporous behavior is only Observed when CO, is used in the absorption studies.32 Cesium sulfide and selenide analogs have been isolated, but the selenide analog is fairly air sensitive and the sulfide analog contains an S8 molecule included in the structure which cannot be removed.33 Most other Open framework chalcogenides are not so thoroughly investigated. Members Of the family A,MGe,,Qlo (M = transition metal; Q = S, Se) are perhaps the second most reported and characterized compounds.34 Some interesting antimony sulfides and selenides have also been published.” Molten salt synthesis has resulted in KBi3S5, a three dimensional material with channels, and (Ph4P)[M(Se6),] (M = Ga, In, T1), a two-dimensional framework with void spaces in the basal plane occupied by tetraphenylphosphonium cations.36 Perhaps the most promising recent development in the search for chalcogenide analogs of zeolites are the hydrothermally synthesized indium sulfide frameworks ASU-31 and ASU- 32, but no absorption studies are reported.37 Concurrent with the hydrothermal self-assembly approach to synthesize microporous chalcogenides is an effort to create mesoporous chalcogenides through supramolecular assembly, as in the synthesis Of MCM-41. The polymerization reaction about the micelles is accomplished by the linkage of Ge4Qm" clusters with transition metals. Preliminary results indicate that structures with worrnholes and channels are formed; however, the thermal stability Of the materials is poor and no absorption studies are reported.38 As hydrothermal synthesis and supramolecular assembly are both synthetic techniques developed for the silicates, a logical extension is the application Of pillaring to the chalcogenides as well. Pillaring has been extended to Other non-silicate layered materials with some success. Pillared phosphonates have been reported, but these are not synthesized by an insertion process.39 The layered materials which are most Obvious candidates for pillaring by insertion are ones which can undergo ion exchange, as this is the mechanism for pillaring in clays. Hydrotalcite has been pillared with polyoxometallates in this manner.12 The cations [Bi,5(OH),,]6+ [A1,3O,,(OI-I),4(H,O)1,]7+ have been inserted in M003, and the latter cation into V,Os as well, by ion exchange.”82 Insertion of large molecules via a redox reaction has been Observed, however, in the insertion Of [Fe4(n-C5H4Me)4(u3-S)4] and [MO4(n-C5H4Pr‘)4(|.t3-Se)4] into M003 and FeOCl.40 As mentioned previously, the layered dichalcogenides undergo intercalation chemistry and have been explored extensively as battery cathode materials due to their ability to accept alkali metal cations through redox intercalation.2'5'10 Redox intercalation has also been used to insert larger molecules such as pyridine" and phthalocyanine dye derivatives42 in TaS,. Also by redox chemistry, metallocenes can be inserted into group IV and V layered dichalcogenides43 and into SnS, and SnSe,.44 Redox insertion is limited, however, in that the pillaring agent must be capable Of reducing the host. An important development in the pillaring of the layered dichalcogenides was the insertion of [Al,,O,,(OH),,,(H,O),,]7° and [Fe,,S,(PEt_,)6]2+ into Nax(H,O)yTaS, by ion exchange.“92 The layers were dispersed (similar to delamination) in a mixture Of n-methylformarnide and water and the clusters introduced, resulting in layer flocculation. The use i of ion exchange to pillar TaS, expands the identity of guest species which can be incorporated to include species which cannot be inserted by redox chemistry, which encompasses most Of the pillaring agents used in clays. 1.3 MOS, and WS, as hosts Unlike the group IV and V chalcogenides, MOS, was notorious for its resistance to intercalation until 1983, when it was discovered that redox insertion is possible but induces a transformation in the coordination environment Of the metal.47 MOS, is found in nature as the mineral molybdenite. The structure, first solved in 1923 by Dickinson and Pauling," consists of a layer of MO atoms sandwiched by layers of sulfur atoms such that the coordination environment of the MO atom is trigonal prismatic. The stacking of the MOS, layers is staggered (Figure 1.2). 10 Figure 1.2. Projection of 2H-MOS, (A) in the ab plane (one layer) and (B) along the c axis. 2H-WS, is isostructural. ll Each MOS, layer has only van der Waals interactions with the neighboring MOS, layers. This allows the layers to slide with respect to one another with relative ease; as such, other polytypes of MOS, can be found with trigonal prismatic coordination Of MO, but a different stacking arrangement. Using a simple system Of prefixes for naming the polytypes, molybdenite is called 2H—MOS, because there are two layers per unit cell and the crystal has hexagonal symmetry. Another polytype, 3R-MoS,, has three layers per unit cell and rhombohedral symmetry. The structure and polytypes Of WS, are analogous to MOS,.46 As mentioned previously, it was discovered that upon intercalation of MOS, and WS, with lithium a structural transformation takes place in which the metal coordination environment changes from trigonal prismatic to octahedral." The new structure, denoted lT-MOS, (and WS,), has been called analogous to the structure type observed for TiS2 (Figure 1.3).“ The practical applications for MOS, (and to a lesser extent WS,) are diverse, and include use as a high temperature lubricant.49 Cobalt and nickel supported MOS, and WS, are among the commercially available catalysts for hydrodesulfurization.’0 Exfoliated MOS, has also been explored as a catalyst for coal liquefaction5l and methanation of CO.52 2H- MOS2 is a semiconductor with a band gap of 1.1 eV. Because it is found in nature, molybdenite is a relatively inexpensive material. 12 *tA‘t-‘Tw. var/aw Va: warm iii; Figure 1.3. Structure of 1T-TiS, viewed (A) in the ab plane and (B) along the c axis. Restacked MOS, and WS, are resemble this structure type if distortions due to metal-metal associations are neglected. 13 Electrochemical studies Of Li insertion in MOS, for solid state battery applications have been conducted,53 but a more important development in the behavior Of LiMoS, was the discovery in 1986 that it exfoliates in water, much like the delamination of clays. The following redox reaction with water was proposed: 5" LiMoS, + H,O --> MOS, (single layers) + LiOH + 1/2 H,(g) Eq. (1.1) The suspension of MOS, layers can be recovered tO a lamellar form by centrifugation, evaporation, or filtration. Furthermore, the layers can encapsulate guest molecules in the restacking process. This is accomplished by first washing away the LiOH generated in the exfoliation reaction, then introducing a water-immiscible solvent such as benzene. An emulsion is formed, and eventually the MOS, begins to collect at the water/solvent interface. If a guest species is dissolved in the organic solvent, it can be incorporated in this manner (Figure 1.4).55 Subsequent studies have shown that some species can also be incorporated without the presence of the organic solvent?6 It is remarkable that it is possible to incorporate neutral species in this manner. Neutral species Of all sizes have been encapsulated and include organic molecules,” polymers,55"56 ferrocene,55m fullerenes,58 and metal chalcogenide clusters.873 The behavior Of WS, is similar.”60 14 Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li Li MOS, ‘Hzo, stir W centrifuge, w /\\ “/0” decant H 20 \ \\ \|>_:./\ ‘ cationic/aqueous i .). $44: 1.6% Ml‘éfik Figure 1.4. Illustration of the exfoliation and flocculation process used to encapsulate neutral and cationic species. 15 Due tO the ability Of MOS, to encapsulate neutral species, it was believed by many scientists that the oxidation of MOS, in the exfoliation process was complete, resulting in neutral layers in suspension. An increasing number Of studies have revealed that it is possible to incorporate cationic species as well, without the detectable presence Of co-encapsulated anions. Transition metals61 and small alkyl ammonium cations,62 along with larger cations such as protonated phenanthroline,63 iron porphyrins,64 dihexadecyldimethyl ammonium,“ and poly(allylamine hydrochloride)“ have been incorporated by the exfoliation/encapsulative precipitation method. These results suggest that MOS, (and WS,) may in fact retain some negative charge. The results of the studies reported in Chapters 2 and 3 support this assertion. Experiments designed specifically to address the issue Of charge on the MS, layers are reported in Chapter 5. The octahedral coordination environment, induced by the lithiation reaction, appears to be retained in restacked MOS, and WS,.“ ’9 However, there are several superlattices which have been Observed in lithiated, restacked, and oxidized MOS, (and WS,) due to metal-metal associations which cause a deviation from the ideal 1T structure type depicted in Figure 2. Initially a Za x 2a superlattice was reported based on transmission X-ray diffraction pattems.‘7°59 Other researchers found evidence for a 2a x a lattice by Scanning Tunneling Microscopy (STM).68 Still others have 16 Observed a 3a x [/30 lattice by X-ray diffraction and STM.69 The layered metal dichalcogenides are prone to these kinds of structural distortions due to the formation of charge density waves (CDWs).7O The structural distortion in restacked MOS, and WS, will be discussed in detail in Chapter 4. 1.4 Pillaring agents Due to the unique ability of exfoliated MOS, and WS, to accept both neutral and cationic species, there are an enormous variety of metal chalcogenide, metal oxide, metal halide, and metal pnictide clusters which could be used as pillaring agents?”73 Obviously, metal chalcogenide clusters are requisite for the synthesis Of an open framework chalcogenide. The four clusters selected for this work were chosen for their history as a successful pillaring agent in other layered materials and/or the desirability of their chemical composition for potential catalytic materials. 17 [Al,,O4(OH),,(H,O),,]Cl.,. The aqueous Al'” ion exhibits diverse coordination chemistry at various pH.74 It was known for years that soluble polynuclear species are formed at pH just below 775 before the [Al,,O4(OH),4(H,O),,]7* cluster, sometimes abbreviated as [A1,,O4o]7*, was crystallized from solution as a sulfate salt.76 The structure of the cation is a distorted version Of a Keggin ion structure Observed in polyoxometallates,73W in which twelve octahedrally coordinated A13+ ions surround one tetrahedrally coordinated Al3+ ion (Figure 1.5a). The octahedral Al3+ ions may be broken into four groups of edge-sharing trimers, each with a common vertex (Figure 1.5b). These trimers are linked to one another through edge sharing, forming a cluster with tetrahedral symmetry. The tetrahedral Al3+ ion at the center is coordinated by the common vertex Of each Of the four trimers. The cluster is approximately 10 A in its smallest projection (visible in Figure 1.5a) and nearly 13 A in its largest dimension. Polynuclear hydroxy aluminum species, particularly the [A1,,O4o]7* cluster, have been explored by many as pillaring agents for clays.”78 Solution and solid state 27A1 NMR have proved useful tOOls in the identification of the cluster in solution and in the intercalated samples.79 The thermal stability of clays pillared with this species permits activity as a 18 Figure 1.5. (A) the [A1,,O,,(OH),,,(H,O),,]7+ cluster. The tetrahedral Al3+ polyhedron is shown in white. (B) A trimer of edge-sharing octahedral A13+ ions with a common vertex (open circle). Four of these trimers are linked together though edge sharing (black lines) to form the cluster. molecular sieve and acid catalyst.78 The cluster has also been used as a pillaring agent for M00380, TaS,“, and V,Os.82 The encapsulation of this cluster will be described in Chapter 2. Co,S,(PPh,)6. This cluster is a member Of a large family Of compounds of the general formula CofiQ,(PR,)6n+ (Q = S, Se, Te; R = ethyl, butyl, phenyl; n = 0,1,2).83"84 The structure is comprised Of a cube Of eight chalcogen atoms containing a cobalt atom in the center Of each face (Figure 1.6a). TO each cobalt atom is ligated one phosphine group (Figure 1.6b). This MGQ, core appears frequently in metal chalcogenide cluster chemistry.71 The size of the C0688 core is about 5 A along the faces Of the cube and 7.5 A along the body diagonal. The overall dimensions Of the cluster, including the phosphine ligands, range from 14 - 15.5 A. Research involving molecular clusters such as C06S,(PPh,)6 is motivated (at least in part) by the hope that they might be used as molecular precursors to extended solids.84a One particularly interesting feature Of the CO,,Q,(PR3 6‘” clusters is their structural and compositional similarity to catalytic materials. As mentioned previously, cobalt supported MOS, is used as a catalyst in hydrodesulfurization (HDS), but the mechanism Of the activity is not well understood?0 One theory to explain the activity, called the “contact synergy” model, involves separate C098, 20 Figure 1.6. (A) M6Q, core (M, black circles, Q, gray circles) Of C068,(PPh3)6. (B) MGQ8 core with phosphine ligands (P, dark gray circles). 21 clusters and MOS, domains.85 Another group of HDS catalysts, called the Chevrel phases, are ternary compounds containing a three dimensional array of linked MOGQB (Q = S, Se, Te) units.86 The encapsulation Of neutral members Of the C06Q,(PR3)6 family into MOS, was initiated in the Kanatzidis laboratory by Bissessur.87 Surface area measurements and magnetic properties were reported, but I-IDS activity was not explored. The CO,,S,(PPh,),5 cluster was selected for further studies due to the relative ease of the cluster synthesis, the large expansion afforded by the bulky ligands, and the relatively high thermal stability of the unencapsulated cluster. The encapsulation of this cluster will be discussed in Chapter 3. [Fe6S,(PEt,)6]2‘°. In addition to the possible use for metal chalcogenide clusters as molecular ”precursors to extended solids, the synthesis of many such clusters, particularly iron chalcogenide clusters, is motivated by a need for models Of compositionally and structurally similar redox centers in biological systems.88 Because they are of interest tO materials scientists as well as bioinorganic chemists, the family Of [Fe6Q,(PR3)6]'” (Q = S, Se, Te; R = methyl, ethyl; n = 0 - 4 ) compounds is also fairly large. 8990'” ‘[Fe6S,(PEt,)6]2* is isostructural to C06S8(PPh3)6 with similar core dimensions (4.5 A across faces and 7.5 A down the body diagonal), but smaller overall dimensions due to the smaller phosphine 22 ligands (10 - 12 A). The [F e6Sg(PEt3)6] 2+ cluster was successfully used as a pillaring agent for TaS, but little was reported beyond the structural 81.92 characterization Of the material The encapsulation of this cluster will be discussed in Chapter 3. [Ni,S,(PEt,)6] 2*. Polynuclear nickel chalcogenide species with the M6Q, structure have not been Observed; rather, a prismane structure is found in Ni,,Se5(PPh,)6.72'93 In general, nickel chalcogenide cluster chemistry seems to be dominated by the linkage of nickel trimers to form larger clusters. The [Ni,S,(PEt,)6] 2* cluster is assembled from three Ni,S3 units (Figure 1.7a) stacked in alternating directions.94 The coordination environment of the nickel atoms in the top and bottom trimers is completed by triethyl phosphine ligands (Figure 1.7b). The dimensions Of the core are 6 A perpendicular to and 6.8 A parallel to the N1383 planes. The overall dimensions of [Ni9S9(PEt,)6] 2* range from 11.5 A to 13 A. This cluster has not been used as a pillaring agent for any lamellar material. Ni supported MOS, and WS,, however, are active HDS catalysts; hence, a nickel chalcogenide pillaring agent is desirable. The encapsulation of this cluster will be discussed in Chapter 3. 23 Figure 1.7. (A) Ni3S3 units (Ni = black circles, S = gray circles) which are stacked in a staggered fashion to form (B), the [Ni,S,(PEt3)6] 2* cluster. The top and bottom trimers are ligated by triethyl phosphine (P = dark gray circles, C = Open circles). 24 The insertion Of metal sulfide and metal oxide clusters into the layered dichalcogenides is a step toward open framework sulfides. The hosts MOS, and WS, are particularly well suited for this task due to their remarkable ability tO exfoliate in water, allowing the encapsulation of large guest molecules between the layers. The clusters [Al,;.,O,,(OH),,,(H,O),,]7+ (Chapter 2); CO6S,(PPh,)6, [Fe688(PEt3)6]2*, and [Ni,S,(PEt3)6]2+ (Chapter 3) have been incorporated and characterized. The structure Of the layered hosts MOS, and WS, has been solved by electron crystallography (Chapter 4), and studies regarding the oxidation state of the layers have been conducted (Chapter 5). 25 References l. a) Fukushima, Y.; Okada, A.; Kawasumi, M.; Kurachi, T.; Kamigaito, 0. Clay Minerals, 1988, 23, 27. b) Okada, A.; Kawakado, M. Japanese Patent Application 1985, NO. SHO 60-217396. 2. SchOllhom, R. Comments Inorg. Chem, 1983, 2, 271. 3. Whittingham, M.S. Prog. Solid St. Chem, 1978, 12, 41. 4. Murphy, D.W.; DiSalvo, F.J.; Hull, G.W., Jr.; Waszczak, J.V.; Mayer, S.F.; Steward, G.R.; Early, S.; Acrivos, J.V.; Geballe, T.H. J. Chem. Phys., 1975, 62, 967. 5. a) Dines, M.B. Mater. Res. Bull, 1975, 10, 287. b) Murphy, D.W.; DiSalvo, F.J.; Hull, G.W., Jr., Waszczak, J.V. Inorg. Chem, 1976, 15, 17. c) Hung, P. L.-K.; Sundararajan, G.; San Filippo, J., Jr. Organometallics, 1982, l, 957. 6. Barrer, R.M.; MacLeOd, D.M. Trans. Faraday Soc., 1955, 51, 1290. 7. a) Johnson, J.W.; Jacobson, A.J.; Butler, W.M.; Rosenthal, S.E.; Brody, J.F.; Lewandowski, J.T. J. Am. Chem. Soc., 1989, 111, 381. b) Gendraud, P.; de Roy, M.E.; Besse, J.P. J. Solid State Chem, 1993, 106, 577. 8. a) Bouhaouss, A.; Aldebert, P.; Mater. Res. Bull, 1983, 18, 1247. b) Aldebert, P.; Baffier, N.; Gharbi, N.; Livage, J. Mater. Res. Bull, 1981, 16, 949. 9. O’Hare, D. New J. Chem, 1994, 18, 989. 10. Rouxel, J. Comprehensive Supramolecular Chemistry, Vol. 7., Alberti, G.; Bein, T., Eds. Pergamon Press, New York, 1996, 77-105. 11. a) Fredenhagen, K.; Cadenbach, G. Z. Anorg. Allg. Chem,1926, 158, 249. b) Schaflrautl, C. J. Prakt. Chem, 1840, 3, 129. 12. a) Li, L.; Ma, S.; Liu, X.; Yue, Y.; Hui, J.; Xu, R.; Bao, Y.; Rocha, J. Chem. Mater., 1996, 8, 204. b) Drezdzon, M.A.; Inorg. Chem, 1988, 27, 4628. 26 13. 14. 15. 16. 17. 18. 19. 20. Clearfield, A. Prog. Inorg. Chem, 1998, 47,371. a) Choy, J.—H.; Park, N.-G.; Kin, Y.-I.; Hwang, S.-H.; Lee, J .-S.; YOO, H.-I. J. Phys. Chem, 1995, 99, 7845. b) Choy, J.-H.; Park, N-G.; Hwang, S.-J.; Kim, D.-H. J. Am. Chem. Soc., 1994, 116, 11564. a) Kanatzidis, M.G.; Marcy, H.O.; McCarthy, W.J.; Kannewurf, C.R.; Marks, T.J. Solid State Ionics, 1989, 32/33, 594. b) Bissessur, R.; DeGroot, D.C.; Schindler, J .L.; Kannewurf, C.R.; Kanatidis, M.G. J. Chem. Soc., Chem. Commun., 1993, 687. a) Wang, L.; Schindler, J.; Kannewurf, C.R.; Kanatzidis, M.G. J. Mater. Chem, 1997, 7, 1277. b) Liu, Y.-J.; Schindler, J.L.; DeGroot, D.C.; Kannewurf, C.R.; Hirpo, W.; Kanatzidis, M.G. Chem. Mater., 1996, 8, 525. Sherry, H.S., Ion Exchange, 2, J. A. Marinsky, Ed., Marcel Dekker, New York, 1969. a) Pinnavaia, T.J. Science, 1983, 220, 365. b) Figueras, F. Catal. Rev. -Sci. Eng., 1988, 30, 457. Walker, G.F.; Garret, W.G. Science, 1967, 156, 385. a) Barrer, R.M. Hydrothermal Chemistry of Zeolites, Academic Press, New York, 1972. b) Breck, D.W. Zeolite Molecular Sieves, Wiley, New York, 1984. 21. a) Davis, ME. Acc. Chem. Res., 1993, 26, 111. b) Bu, X.; Feng, P.; Stucky, 6D. Science, 1997, 278, 2080. 22. 23. 24. 25. a) Vaughan, D.E.W., Chem. Eng. Prog., 1988, 25. b) Chen, N.Y.; Garwood, W.E.; Dwyer, F.G. Shape Selective Catalysis in Industrial Applications, Marcel Dekker, New York, 1989. Kresge, C.T.; Leonowicz, M.E.; Roth, W.J.; Vartuli, J.C.; Beck, J.S. Nature, 1992, 359, 710. Mortland, M.M.; Berkheiser, V.E. Clays Clay Miner., 1976, 24, 60. a) Knudson, M.I.; McAtee, J.L. Clays Clay Miner., 1973, 21, 19. b) Pinnavaia, T.J. Clays Clay Miner., 1978, 26, 319. 27 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Johnson, I.D.; Werpy, T.A.; Pinnavaia, T]. J. Am. Chem. Soc., 1988, 110, 8545. Bowes, C.L.; Ozin, G.A. Adv. Mater., 1996, 8, 13. Bedard, B.L.; Wilson, s.T.; Vail, L.D.; Bennet, J.M.; Flanigen, E.M. Zeolites, Facts, Figures, Future. Jacobs, P.A.; van Santen, R.A.; Eds. Elsevier Science Publishers B., Amsterdam, the Netherlands, 1989, 375. Jiang, T.; Lough, A.; Ozin, G.A.; Bedard, R.L. J. Mater. Chem, 1998, 8, 733. a) Ahari, H.; Ozin, G.A.; Bedard, R.L.; Petrov, 8.; Young, D. Adv. Mater., 1995, 7, 370. b) Ahari, H.; Bowes, C.L.; Jiang, T.; Lough, A.; G.A. Ozin, Bedard, R.L.; Petrov, S.; Young, D. Adv. Mater., 1995, 7, 375. Jiang, T.; Lough, A.; Ozin, G.A.; Bedard, R.L.; Broach, R. J. Mater. Chem, 1998, 8, 721. Bowes, C.L.; Petrov, s.; Vovk, G.; Young, D.; Ozin, G.A.; Bedard, R.L. J. Mater. Chem, 1998, 8, 711. a) Sheldrick, W.S.; Braunbeck, H.-G. Z. Naturforsch, 1990, 45b, 1643. b) Marking, G.A.; Kanatzidis, M.G. Chem. Mater., 1995, 7, 1915. a) Yaghi, O.M.; Sun, 2.; Richardson, D.A.; Groy, T.L. J. Am. Chem. Soc., 1994, 116, 807. b) Tan, K.; Darovsky, A.; Parise, J.B. J. Am. Chem. Soc., 1995, 117, 7039. c) Ahari, H.; Garcia, A.; Kirkby, S.; Ozin, G.A.; Young, D.; Lough, A.J. J. Chem. Soc., Dalton Trans, 1998, 2023. a) Parise, J.B. Science, 1991, 251, 293. b) Hanko, J.A.; Kanatzidis, M.G. Angew. Chem. Int. Ed., 1998, 37, 342. c) Stephan, H.-O.; Kanatzidis, M.G. J. Am. Chem. Soc., 1996, 118, 12226. a) McCarthy, T.J.; Tanzer, T.A.; Kanatzidis, M.G. J. Am. Chem. Soc., 1995, 117 , 1294. b) Dhingra, S.; Kanatzidis, M.G. Science, 1992, 258, 1769. 28 37. 38. 39. 40. 41. 42. 43. 45. 46. 47 Li, H.; Laine, A.; O’Keefe, M.; Yaghi, O.M. Science, 1999, 283, 1145. a) Wachhold, M.; Rangan, K.K.; Billinge, S.J.L.; Petkov, V.; Heising, J.; Kanatzidis, M.G. submitted for publication. b) MacLachlan, M.J.; Coombs, N .; Ozin, G.A. Nature, 1999, 397, 681. Dines, M.D.; DiGiacomo, P.M.; Callahan, K.P.; Griffith, P.C.; Lane, R.H.; COOksey, R.E. Chemically Modified Surfaces in Catalysis and Electrocatalysis, Miller, J .S., Ed., Amer. Chem. Soc. Symp. Ser. , 1982, 192, Washington, DC, Chapter 12. Chatakondu, K.; Green, M.L.H.; Qin, J.; Thompson, M.E.; Wiseman, P.J. J. Chem. Soc., Chem. Commun., 1988, 223. Gamble, F.R.; DiSalvo, F.J.; Klemm, R.A.; Geballe, T.H. Science, 1970, 168, 568. Preobazhensky, V.B.; Zhemova, A.Ja.; Babichev, A.P.; KObrin, I.K. Solid State Commun., 1978, 27, 127. a) Dines, M. B. Science, 1975, 188, 1210. b) Clement, R.P.; Davies, W.B.; Ford, K.A.; Green, M.L.H.; Jacobson, A.J. Inorg. Chem, 1978, 17, 2754. . a) O’Hare, D.; Jaegermann, W.; Williamson, D.L.; Ohuchi, F.S.; Parkinson, B.A. Inorg. Chem, 1988, 27, 1537. b) Forrnstone, C.A.; FitzGerald, E.T.; O’Hare, D.; Cox, P.A.; Kurmoo, M.; HOdby, J.W.; Lillicrap, D.; Goss-Custard, M. J. Chem. Soc., Chem. Commun., 1990, 501. a) Dickinson, R.G.; Pauling, L. J. Am. Chem. Soc., 1923, 45, 1466. b) Bronsema, K.D.; de Boer, J .L.; Jellinek, F. Z. Anorg. Allg. Chem, 1986, 540/541, 15. a) Wildervanck, J.C.; Jellinek, F. Z. Anorg. Allg. Chem, 1964, 328, 309. b) Glemser, O.; Sauer, H.; KOnig, P. Z Anorg. Chem, 1948, 257, 241. c) van Arkel, A.E., Rec. Trav., Chim, Pays-Bas, 1925, 45, 437. d) Ehrlich, R; Z. Anorg. Chem, 1948, 257 , 247. Py, M.A.; Haering, R. R. Can. J. Phys, 1983, 61, 76. 29 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. Chianelli, R.R.; Scanlon, J.C.; Thompson, A.H. Mater. Res. Bull, 1975, 10, 1379. Fleischauer, P.D. Thin Solid Films, 1987, 154, 309. a) Chianelli, R.R.; Ruppert, A.P.; Jose-Yacaman, M.; Vézquez-Zavala, A. Catal. Today, 1995, 23, 269. b) Topsoe, H.; Clausen, B.S. Catal. Rev. -Sci. Eng., 1984, 26, 395. c) Weisser, 0.; Landa, S. Sulfided Catalysts, their Properties and Applications. Pergamon Press, New York, 1973. Bockrath, B.C.; Parfitt, D.S. Catal. Lett., 1995, 33, 201. Miremadi, B.K.; Morrison, S.R. J. Catal., 1987, 103, 334. a) Mulhem, P.J. Can. J. Phys, 1989, 67, 1049. b) Julien, C. Microionics: Solid State Integrable Batteries. Balkanski, M., Ed. Elsevier Science, 1991, New York, 309. Joensen, P.; Frindt, R.F..; Morrison, S.R. Mater. Res. Bull, 1986, 21, 547. a) Divigalpitiya, W.M.R.; Frindt, R.F.; Morrison, S.R. Science, 1989, 246, 369. b) Zhou, X.; Yang, D.; Frindt, R.F. J. Phys. Chem. Solids, 1996, 57, 1137. c) Kosidowski, L.; Powell, A.V. Chem. Commun., 1998, 2201. a) Divigalpitiya, W.M.R.; Frindt, R.F.; Morrison, S.R. J. Mater. Res., 1991, 6, 1103. b) Ruiz-I-Iitzky, E.; Jimenes, R.; Casal, B.; Manriquez, v.; Santa Ana, A.; Gonzalez, G. Adv. Mater., 1993, 5, 738. c) Kanatzidis, M.G.; Bissessur, R.; DeGroot, D.C.; Schindler, J.L.; Kannewurf, C.R. Chem. Mater., 1993, 5, 595. d) Bissessur, R.; Kanatzidis, M.G.; Schindler, J.L.; Kannewurf, C.R. J. Chem. Soc., Chem. Commun., 1993, 1582. e) Lemmon, J.P.; Lerner, M.M. Chem. Mater., 1994, 6, 207. 1) Wang, L.; Schindler, J.; Thomas, J.A.; Kannewurf, C.R.; Kanatzidis, M.G. Chem. Mater., 1995, 7, 1753. g) Oriakhi, C.O.; Nafshun, R.L.; Lerner, M.M. Mater. Res. Bull, 1996, 31, 1513. Tagaya, H.; Hashimoto, T.; Karasu, M.; Izumi, T.; Chiba, K. Chem. Lett., 1991, 2113. 30 58. 59. 61. 62. 63. 65. 67. 68. Divigalpitiya, W.M.R.; Frindt, R.F.; Morrison, S.R. Mater. Res. Symp. Proc., 1992, Boston. a) Yang, D.; Frindt, R.F. J. Phys. Chem. Solids, 1996, 57, 1113. b) Tsai, H.-L.; Heising, J.; Schindler, J.L.; Kannewurf, C.R.; Kanatzidis, M.G. Chem. Mater., 1997, 9, 879. . a) Wang, L. Ph.D. Dissertation, Michigan State University, 1999. b) Tsai, H.-L.; Kanatzidis, M.G., tO be published. a) Danot, M.; Manson, J.L.; Golub, A.S.; Protzenko, G.A.; Fabritchnyi, P.B.; Novikov, Yu.N.; Rouxel, J. Mater. Res. Bull., 1994, 29, 833. b) Golub, A.S.; Payen, c.; Protzenko, G.A.; Novikov, Yu.N.; Danot, M. Solid State Commun., 1997, 102, 419. c) Gash, AE.; Spain, A.L.; Dysleski, L.M.; Flaschenreim, C.J.; Kalaveshi, A.; Dorhout, P.K.; Strauss, S.H. Environ. Sci. Tech, 198, 32, 1007. d) Golub, A.S.; Shumilova, I.B.; Zubavichus, Y.V.; Jahncke, M.; 81155- Fink, G.; Danot, M.; Novikov, Y.N. J. Mater. Chem, 1997, 7, 163. e) Gee, M.A.; Frindt, R.F.; Joensen, R; Morrison, S.R. Mater. Res. Bull., 1986, 21, 543. Golub, A.S.; Protsenko, G.A.; Burnileva, L.V.; Buyanovskaya, A.G.; Novikov, Yu.N. Russ. Chem. bull, 1993, 42, 632. Golub, A.S.; Shumilova, I.B.; Novikov, Yu.N.; Mansot, J.L.; Danot, M. Solid State Ionics, 1996, 91, 307. Nakagaki, S.; Mangrich, A.S.; Wypych, F. Inorg. Chim. Acta, 1997, 254, 213. Taguchi, Y.; Kimura, R.; Azumi, R.; Tachibana, H.; Koshizaki, N.; Shimomura, M.; Momozawa, N.; Sakai, H.; Abe, M.; Matsumoto, M. Langmuir, 1998, 14, 6550. . Ollivier, P.J.; Kovtyukhova, N.I.; Keller, S.W.; Mallouk, T.E. Chem. Commun., 1998, 1563. Yang, D.; Sandoval, S.J.; Divigalpitiya, W.M.R.; Irwin, J.C.; Frindt, R.F. Phys. Rev. B., 1991, 43, 12053. Qin, X.R.; Yang, R.; Frindt, R.F.; Irwin, J .C. Ultramicroscopy, 1992, 42-44, 630. 31 69. 70. 71. 72. 73. 74. 75. 76. 77. a) Wypych, F.; SchOllhorn, R. J. Chem. Soc., Chem. Commun., 1992, 1386. b) Wypych, F.; Weber, Th.; Prins, R. Chem. Mater., 1998, 10, 723. a) Wilson, J.A.; DiSalvo, F.J.; Mahajan, 5. Adv. Phys, 1975, 10, 1379. b) Wilson, J.A.; DiSalvo, F.J.; Mahajan, S. Phys. Rev. Lett., 1974, 32, 882. Dance, 1.; Fisher, K. Prog. Inorg. Chem, 1994, 41, 637. Fenske, D.; Ohmer, J.; Hachgenei, J.; Merzweiler, K. Angew. Chem. Int. Ed. Engl., 1988, 27, 1277. a) Pope, M.T. Heteropoly and Isopoly Oxometallates. Springer- Verlag, New York, New York, 1983. b) Fleming, P.B.; Mueller, L.A.; McCarley, B.E. Inorg. Chem, 1967, 6, l. c) Koknat, F.W.; Adaway, T.J.; Erzerum, S.I.; Syet, S. Inorg. Nucl. Chem. Lett., 1980, 16, 307. Cotton, F.A.; Wilkinson, G. Advanced Inorganic Chemistry, 5th Ed. John Wiley and Sons, New York, 1988, 215-217; 813-819. a) Kohlschiitter, H.W.; Hantelrnann, P. Z. Anorg. All. Chem, 1941, 248, 319. b) Broset, C.; Biederrnann, G.; Sillén, L.G. Acta. Chem. Scand., 1954, 8, 1917. a) Johansson, G.; Lundgren, G.; Sillen, L.G.; Soderquist, R. Acta. Chem. Scand., 1960, 14, 769. b) Johansson, G. Acta. Chem. Scand., 1960, 14, 771. c) Tsai, P.P.; Hsu, P.H. Soil Sci. Soc. Am. J., 1985, 49, 1060. a) Brindley, G.W.; Sempels, R.E. Clay Miner., 1977, 12, 229. b) Lahav, N.; Shani, U.; Shabtai, J. Clays Clay Miner., 1978, 26, 107. c) Singh, S.S.; Kodama, H. Clays Clay Miner., 1988, 36, 397. d) Schoonheydt, R.A.; van den Eynde, J.; Tubbax, H.; Leeman, H.; Stuyckens, M.; Lenotte, 1.; Stone, W.B.E. Clays Clay Miner., 1993, 41, 598. e) Coelho, A.V.; Poncelet, G. Appl. Catal., 1991, 77, 303. f) Bukka, K.; Miller, I.D.; Shabtai, J. Clays Clay Miner., 1992, 40, 92. g) Malla, P.B.; Komarneni, S. Clays Clay Miner., 1993, 41, 472. 32 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. a) Vaughan, D.W. E.; Lussier, R.J.; Magee, J .S. US. Patent 4,176,090 (1979). b) Vaughan, D.W. E.; Lussier, R.J.; Magee, J.S. US. Patent 4,271,043 (1981). c) Shabtai, J.; US. Patent 4,238,364 (1980). a) Bottero, J.Y.; Cases, J.M.; Flessinger,F.; Poirier, J.B. J. Phys. Chem, 1980, 84, 2933. b) Akitt, J.W.; Farthing, A.; J. Chem. Soc., Dalton Trans, 1981, 1617. c) Plee, D.; Borg, F.; Gatineau, L.; Fripiat, L]. J. Am. Chem. Soc., 1985, 107, 2362. a) Lerf, A.; Lalik, E.; KOlOdziejski, W.; Klinowski, J. J. Phys. Chem, 1992, 96, 7389. b) Nazar, L.F.; Liblong, W.S.; Yin, X.T. J. Am. Chem. Soc., 1991, 113, 5889. c) Nazar, L.F.; Yin, X.T.; Zinkweg, D.; Zhang, Z.; Liblong, S. Mat. Res. Soc. Symp. Proc., 1991, 210, 417. Nazar, L.F.; Jacobson, A.J. J. Mater. Chem, 1994, 4, 149. Mori, M.; Isobe, T.; Senna, M. Solid State Ionics, 1995, 81, 157. a) Hong, M.; Huang, 2.; Lei, X.; Wei, G.; Kang, B.; Liu, H. Inorg. Chim. Acta, 1989, 159, l. b) Jiang, F.; Huang, X.; Cao, R.; Hong, M.; Liu, H. Acta Cryst. C., 1995, 1275. c) Hong, M.; Huang, 2.; Lei, X.; Wei, G.; Kang, B.; Liu, H. Polyhedron, 1991, 10, 927. a) Stuczynski, S.M.; Kwon, Y.-U.; Steigerwald, M.L. J. Organomet. Chem, 1993, 449, 167. b) Steigerwald, M.L.; Siegrist, T.; Stuczynski, S.M. Inorg. Chem, 1991, 30, 2256. c) Cecconi, F.; Ghilardi, C.A.; Midollini, S.; Orlandini, A.; Bencini, A. J. Chem. Soc., Dalton Trans, 1996, 3991. d) Cecconi, F.; Ghilardi, C.A.; Midollini, S.; Orlandini, A. Inorg. Chim. Acta, 1983, 76, L183. e) Cecconi, F.; Ghilardi, C.A.; Midollini, 8.; Orlandini, A.; Zanello, P. Polyhedron, 1986, 5, 2021. Delmon, B. Prepr. Am. Chem. Soc. Div. Pet. Chem, 1977, 22, 503. a) Chevrel, R.; Sergent, M.; Prigent, J. J. Solid State Chem, 1971, 3, 515. b) Chevrel, R.; Hirrien, M.; Sergent, M. Polyhedron, 1986, 5, 87. c) McCarty, K.F.; Anderegg, J.W.; Schrader, G.L. J. Catal., 1985, 93, 375. a) Bissessur, R.; Heising, J.; Hirpo, W.; Kanatzidis, M.G. Chem. Mater., 1996, 8, 318. b) Bissessur, R. “Synthesis and Characterization 33 88. 89. 90. 91. 92. 93. 94. of Novel Intercalation Compounds of Molybdenum Trioxide and Molybdenum Disulfide.” Ph.D. Dissertation, Michigan State University, 1994. a) Holm, R.H. Chem. Soc. Rev., 1981, 10, 455. b) Coucouvanis, D. Acct. Chem. Res., 1981, 14, 201. Steigerwald, M.L.; Siegrist, T.; Gyorgy, E.M.; Hessen, B.; Kwon, Y.- U., Tanzler, S.M. Inorg. Chem, 1994, 33, 3389. a) Cecconi, F.; Ghilardi, C.A.; Midollini, S.; Orlandini, A. Inorg. Chim. Acta, 1997, 254, 387. b) Agresti, A.; Bacci, M.; Cecconi, F.; Ghilardi, C.A.; Midollini, S. Inorg. Chem, 1985, 24, 689. c) Cecooni, F.; Ghilardi, C.A.; Midollini, S.; Orlandini, A.; Zanello, P. J. Chem. Soc., Dalton Trans, 1987, 831. d) Cecooni, F.; Ghilardi, C.A.; Midollini, S. J. Chem. Soc., Chem. Commun., 1981, 640. Goddard, C.A.; Long, J.R.; Holm, R.H. Inorg. Chem, 1996, 35, 4347. Nazar, L.F.; Jacobson, A.J. J.Chem. Soc., Chem. Commun., 1986, 570. Fenske, D.; Ohmer, J. Angew. Chem. Int. Ed. Engl., 1987, 26, 148. a) Cecconi, F.; Ghilardi, C.A.; Midollini, s. Inorg. Chem, 1983,22, 3802. b) Ghilardi, C.A.; Midollini, S.; Sacconi, L. J. Chem. Soc., Chem. Commun., 1981, 47. 34 CHAPTER 2 Encapsulation and Rietveld Structural Characterization Of the [AI,,O,.(OH),,(H,O),,]7+ cluster into MOS, and WS,. 35 2.0 Abstract The synthesis Of [Al,,O4(OH),4(H,O).2]xMS, (x=0.02-0.05, M=MO, x=0.02-0.055, M=W) was accomplished by a precipitative encapsulation technique using single layers of MS,. The products were characterized by powder X-ray diffraction, energy dispersive spectroscopy (EDS), thermogravimetric analysis (TGA), 27A1 Magic Angle Spinning Nuclear Magnetic Resonance (MAS-NMR), High Resolution Transmission Electron Microscopy (HRTEM), room temperature electrical conductivity measurements, and surface area measurements. Powder X-ray diffraction patterns show an expansion of approximately 9.9A. 27Al MAS-NMR indicates that the cluster is intact between the MS, layers. One-dimensional electron density mapping and Rietveld refinement performed on the powder diffraction data shows that the cluster is oriented with its C3 symmetry axis perpendicular to the layers. The samples exhibit conductivity values from 3-14 S/cm. TGA shows that the layers remain expanded to 100°C, and partially expanded to 300°C. Surface area measurements suggest that the space between the clusters is not accessible. 36 2.1 Introduction Since Barrer and MacLeOd reported the first pillared species in 1955‘, much effort has gone into the development of microporous materials for use as size selective molecular sieves and as catalysts. Zeolites and pillared clays have been studied extensively for these applications”; however, these materials are insulators. Because WS, and MOS, have interesting electronic and optical properties, pillared WS, and MOS, may be useful for unique applications similar to but distinct from the applications for zeolites and pillared clays. The availability and relatively low cost of MOS, make it an especially desirable chalcogenide host. The practical applications of MOS, itself are quite diverse. These include use as a solid lubricant“, a catalyst’, and as a host material for solid state batteries.6 MOS, has been intercalated with a variety of compounds which include polymers”, small organic molecules9, and inorganic complexes such as ferrocene10 and the cobalt clusters CO,,Q,(PR,)6 (Q=S, Se, Te; R=alkyl)”. WS, is isostructural to MOS,, with similar intercalation chemistry, although less well explored.‘2 Despite the considerable intercalation history of layered dichalcogenides, microporous pillared chalcogenides have not been reported to date. The intercalation of TaS, with [A1,,O,,(OH),,,(H,O),,]7+ and the iron cluster [Fe68,(PEt,)6]2‘° has been achieved, but no information about the 37 porosity of these materials is reported.13 Intercalation of the C068,,(PPh3)6 clusters into MOS, increase the surface area three or fourfold, but TEM studies show that the lamellar expansion due to the clusters is localized.14 We have chosen to explore [A1,3O,,(OH),,,(H,O),,]7+ as a pillaring agent for MOS, because of its successful history as a pillaring agent for clays” and because alumina-supported MOS, is used as a catalyst for hydrodesulfurization.5 2.2 Experimental 2.2.1 Synthesis MOS, was purchased from Cerac and WS, from Alfa Aesar. LiBH4, n-Butyl lithium, hexane, and AlCl3 were purchased from Aldrich. BaCl,'2H,O, Na,SO,,010H,O, and NaOH were purchased from J .T. Baker. All compounds were used as received except hexane, which was dried over CaH, (Aldrich). LiMS,”. LiMoS, was prepared by a) stirring 2H-MOS, in threefold excess of n-butyl lithium in dry hexane for 2 days or b) heating with LiBH4 at 300°C for 2 days. LiWS2 was prepared by the latter method at 350°C. MS, aqueous suspension. LiMS, was exfoliatedl7 in deionized, deoxygenated H,O via a redox reaction, generating single layers, lithium hydroxide, and hydrogen gas. 38 LIMSZ(S) ‘1' H200) ---> M82 (single layers) '1' LlOH(aq) + 0.5H2 (g) Eq. (2.1) Eq. 2.1 has been proposed for the exfoliation of MS, by Divigalpitiya et al.9 The exfoliated suspension (pH>12) was centrifuged for 0.5 hr and rinsed with deionized, deoxygenated H,O three times in order to reduce the pH of the solution to about 7 (pH ~12, pH ~9, and pH ~7 after each rinse). The product was then re-suspended in H,O and stirred 0.5 hr prior to use. [A1,,O4(OH),,,(H,O),,]CI7 solution.18 The [A1,3O4(OH),4(H,O),,]7* was prepared by the slow addition Of 500 ml of NaOH (0.24M) to 500 ml of an aqueous AlCl3 solution (0.1M) at 80°C with stirring. It was then crystallized as the sulfate salt Na[Al,3O4(OH),4(H,O),,](SO4), in order to isolate it from any other soluble Al-containing species. It was then redissolved by stirring in excess BaCl, solution, and the BaSO4 precipitate removed by filtration. This method results in excess Ba2+ ions also present in the solution. [Al,,O4(OH),4(H,O),,],MS,. The aqueous MS, suspension was added to various amounts of a 0.03M solution (approximately) of [A1,_,,O,,(OH),4(I-I,O),,]Cl7 in H,O.l9 After stirring for several hours the product was isolated via centrifuge, rinsed, and dried on a glass slide. The product was a shiny film which could be scraped off the glass and ground to a fine black powder. 39 2.2.2 Characterization Powder X-ray diffraction patterns were recorded using a Rigaku- Denki/RW400F2 (Rotaflex) rotating anode X-ray diffractometer using Cu- Ka radiation. HRTEM micrographs were acquired using a J EOL 4000EX II at Argonne National Laboratory, operating at 400kV. Energy Dispersive Spectroscopy (EDS) was used to determine the ratios of Al to M0 (or W) on a JEOL-JSM-6400V at an accelerating voltage of 20kV with samples mounted on non-aluminum stubs using carbon paint or carbon tape, equipped with a Tracor Northern 5500 X-ray microanalysis attachment. Solid state 27Al MAS-NMR measurements were performed using a Varian 400MHz instrument tuned to 130.4 MHz pulse frequency, and at MAS frequencies of 4 and 6.2 MHz. Therrnogravimetric measurements were Obtained with a Shimadzu TG 50 instrument using oxygen or nitrogen flow and a heating rate of 50C or 20C/min. BET surface area measurements were performed on a Quantasorb Jr. Sorption System using ultra-pure nitrogen gas as the adsorbate and ultra-pure helium gas as the carrier. The surface areas of the samples were calculated using the BET equation.20 Conductivity measurements of the materials at room temperature were obtained for a pressed pellet using the four prong probe and a Keithley 236 source measure unit. 40 2.3 Results and Discussion 2.3.1 Evidence Of Cluster Encapsulation The predominance of the 001 reflections in the X-ray powder diffraction pattern indicate that the layers are well oriented, with a d- spacing of about 16 A (see Figure 2.1). Restacked MOS, has a d-spacing of 6.2 A, so the expansion of the layers is about 9.9 A. This value is consistent with the dimensions of the cluster, which are about 10.2 A along the C3 symmetry axis, 12.8 A along one C, symmetry axis and 11.5 A along the other C, axis (average diameter 11.5 A) (see orientations (A), (B), and (C), respectively, in Figure 2.2). Our observed expansion suggests that the cluster is oriented with its C3 axis perpendicular to the layers. Loadings of the cluster were calculated using Energy Dispersive Spectroscopy (EDS) to quantitate an Ale0 (or Al:W) ratio. Conversion to oxides by heating to 650°C under a stream of O, prior to quantitation was necessary for MOS,-containing samples because the X-ray emission lines (for EDS) of Mo and S overlap completely, hindering quantitative analysis. The Observed values ranged from [A1,,O4(OH),4(H,O)1,]O,O,MOS, to [A1,304(OH),4(H,O)1,]0_05MOS,. The theoretical maximum loading of the 41 (A II1 TII III III IIr I71 II ,_ F I l l T I O O 002 fiTII'IIIjTII'III lllllllllllllllllllllllllll Intensity (arbitrary units) bIIII'III'II E0 lllllJllllJlLllllllll 4o 60 80 100 120 140 29 (degrees) C N O (B) IIIFTI—ITIIIIIIIIIIIIIIIII I F o o Intensity (arbitrary units) illllllllllllLlllll CIIIIlITIIITIIIITIII h— .1 lllllllllllelllJJlll o 20 4o 60 80 100 120 140 26 (degrees) Figure 2 . l . X-ray diffraction patterns of (A) [M1304(OH)24(H20)12]o.osM082 and (B) [M13O4(OI'I),4(H,O)1,]0055WS, used for electron density maps and Rietveld refinements. Strong preferred orientation of the samples leads to the predominance of the 001 reflections. 42 The gray circles represent Al atoms, open Figure 2.2. Three possible orientations of the [Al1,0,,(OH),,,(I-I,O),,]7+ cluster between the layers. circles are 0 atoms. H atoms are omitted for clarity. 43 cluster (if one assumes hcp packing of the spheres) is [A1,304(OH),4(H,O),,]0_06MOS,. Conversion to oxide was not necessary for samples containing WS,, and comparable stoichiometries (up to [A1,3O4(OH),4(H,O)1,]0_055WS,) were observed by EDS. The stoichiometry of selected samples was confirmed by elemental analysis. Lithiated and exfoliated MOS, was presumed to be neutral because it is possible to intercalate neutral molecules.9 One surprising discovery was that no chloride is detected in the samples by EDS. It appears that this cation intercalates without its chloride anion. The strong intensity of reflections, the relatively high degree of order, and the ease of encapsulation of the cation seems to indicate that the MOS, and WS, have some negative charge. This would explain the driving force for the intercalation. We have observed intercalation of other cations without anions, consistent with a negative charge on the layers.21 The observed loadings of the cluster, presuming no change in its charge, suggest that the negative charge on the layers is between 0.1 and 0.4. This wide range Of possible negative charge suggests that the charge on the aluminum cluster may vary, depending on the loading. Another surprising result is that the excess Ba2+ ions found in solution with the cluster do not appear to be encapsulated. The layers appear to favor the soft [A1,,O4(OH),4(H,O),,]7* ion over the hard, electropositive Ba2+ ion. Studies to probe the ionicity of the layers in detail are described in Chapter 5. 44 MAS-NMR 27Al spectra of several samples with various loadings of cluster exhibited a peak at about 64 ppm corresponding to the tetrahedral aluminum at the center of the cluster (Figure 2.3). This peak has been observed by others who have intercalated the cation into various hosts and is found in the sulfate salt of the unintercalated cluster.'0'ls The spectra also exhibit a broad peak centered around 0 ppm which can be attributed to octahedral aluminum. Intermediate coordination environments may contribute to the broad signal found between the two peaks in [A113O4(OH),4(H,O)1,]0.0,MOS,. This suggests that, although some of the cluster is clearly intact, there may be some decomposition products encapsulated as well. An effort was made to image the encapsulated clusters, or at least the lamellar expansion, by HRTEM. HRTEM studies of thin sectioned [Al,,O4(OH),4(I-I,O),,]7*-pillared clays: revealed a lamellar expansion (the cluster, as it is comprised of light atoms, apparently does not scatter sufficiently to be directly visible); however, it was smaller than the expansion indicated by the powder X-ray diffraction patterns.22 The authors attributed this to dehydration of the cluster in the high vacuum of the electron microscope. Rather than thin sectioning, [A1,,O4(OH),,,(H,O)1,],03MOS, and [Al,,O4(OH),4(I-I,O),,]0_03WS, were dusted on a carbon supported copper grid and images were acquired at the edges 45 W I IIII IIII III IIIIIIIIIIIII IIFI FIIII- I71 200 150 100 so 0 -50 -10'0 .150 -260 ppm IWWTIVI r rII1TT Ti 1' .r.lwx rl'T #71: . 1.11.11. .ir 200 150 100 50 o -50 -160 -15'0 ado ppm Fi re 2.3. Solid State MAS-NMR 27A1 s ctra of (A) [A 1304(0H)24(H20)12]o.03M0$2 and (B) a[A11304(OH)24(H20)12](SO4)2. The peak at 64 ppm (asterisk) rs attributed to the tetrahedral Al3+ atom at the center of the cluster; the broad peak at 0 ppm (circle) is attributed to octahedral Al3+. 46 Figure 2.4. HRTEM Micrographs of (A) [A113O4(OH)2,4(H20)12]o,03M082 and (B) [A11304(0H)74(H20)12]o_04WSZ. Lamellar regions are visible in both; however, d-spacings are the same as the collapsed material, suggesting a "calzone"—like structure. 47 of the particles. A view parallel to the layers is visible in Figure 2.4. In contrast to the powder X-ray diffraction pattern, a more reliable method which probes the bulk of the material, the layers do not appear to be expanded. Figure 2.4b contains regions in which the layers appear to be locally expanded, but scattering centers due to the cluster are not visible, and regions of unexpanded WS, are also found in the micrograph. This suggests that the interior of the particle is pillared, but at the edges the layers are mostly collapsed, resembling a “calzone”. The electrical conductivity of thin films of the materials, dried on glass microscope slides, were measured. If measured within twenty four hours of the exfoliation, the conductivity of the cluster intercalated MOS, ranges from 3-14 S/cm. In contrast, the room temperature conductivity of 2H-MOS, is only ~10'5 S/cm. The metallic properties of these samples can be attributed to a phase transition in MOS, upon lithiation.23 48 2.3.2 One-Dimensional Electron Density Mapping In order to probe the structure and orientation of the encapsulated cluster, one-dimensional electron density mapping calculations were carried out on samples Of the cluster in MOS, and in WS, prepared using excess cluster to generate particularly well ordered materials (see Figure 2.1).24 The integrated intensities (I) of the 001 reflections in the X-ray powder diffraction patterns were extracted and converted to structure factors by the following relation: |F(l).b.| = (I/Lp)“2 Eq- (2.2) in which F(l)obs is the structure factor for each reflection and Lp, a Lorentz-polarization correction, is Lp = (1 + c05220)/(sin20c080) Eq. (2.3) Sixteen 001 reflections were used in the calculations for the sample containing MOS, and eighteen 001 reflections for the sample containing WS,. In order to calculate an electron density map it is necessary to know the signs [(+) or (-) for a centrosymmetric structure] as well as the magnitudes of the F(l)obs values. The recovery of these signs is the essence of (centrosymmetric) single crystal structure determination. In order to recover the signs of the F(1)obs values, an assumption is made that the contribution of the [Al,3O.,(OH),,,(I-I,O),,]7+ cluster to each structure factor 49 is small relative to the contribution from the MOS, and WS, layers. Therefore, we can calculate the structure factors for the expanded, empty layers and apply the signs of these structure factors (F(l)c,,c) to the |F(l)obs| values. This is a reasonable assumption because the cluster is comprised of light atoms which will not scatter nearly as well as the heavy atoms found in the MOS, and WS, layers. Subsequent calculations including the cluster (computed for modeling purposes) resulted in no changes in the signs of the structure factors. Calculation of F(l)calc for the expanded, empty layers is more tedious than calculation of |F(l)obs| because one must first calculate the atomic scattering factors (f) for MO (or W) and S by the following equation: 4 f(9) = [2a,e<-b.-/12 sinz 6') + Cid-198m2 9/ 12) Eq. (2.4) i=1 where a,, bi, and c are a series of 3 nine coefficients which describe an exponential decay that is specific to each element (in this case Mo, W, or S), 0 is the Bragg angle, A is the wavelength of the radiation (1.54184 A for Cu K,), and B is the temperature factor (B = 2 in these calculations). The value of (f) for each element varies as a function Of 0, and as such will be different for each F(1)c,,c. Once f(0) for each atom is known, each F(l)calc can be calculated: 50 N F(l)calc = Zij COS(27Tle) Eq. (2.5) i=1 where, if there are N atoms in the structure, fj is the scattering factor Of atom j (at the particular 0 value associated with each F(l) ), l is the 001 calc index, and zj is the fractional coordinate of atom j on the c-axis. The signs Of these F(l)c,,c are then applied to the corresponding |F(1)obsl values, and electron density (p) is calculated: p(z) = (1 / c)[22 F(l)0,,_, cos(27tlz)] Eq.(2.6) l where c is the c-axis, l is the 001 index, and z is a fractional coordinate along the c-axis (varied from -0.2 to 1.2 in increments of 0.01 in the calculation to create a “map” of the electron density). In order to interpret a one-dimensional electron density map, it is necessary to compare it to at least one structural model. The [A1,3O4(OH),4(H,O),,]7+ cluster is a polyhedral species with well-defined 66 ‘99 faces; as such, it may prefer to Slt in certain orientations over a completely random arrangement in the gallery. Three plausible orientations for the cluster are depicted in Figure 2.2. Orientation (A), in which the C3 axis of the cluster is perpendicular to the layers, seems the most plausible on the basis of the Observed d-spacing in the X—ray powder 51 Table 2.1. 113(1),) and Fa)calc for [A1,,O,(OH),,(H,O),,],,,Mos,. (th1 d... (A) d... (A) F... F....(A) F....131,),],,,ws, (A = 11.4 ). Both exhibit a predominance of 001 reflections and greater order than the patterns for [Co,S,(PPh,)6],,6MOS,,but less order than for [All,O4(OH),4(H,O)1,]0.O,MOS,. 84 er ._, Figure 3.5. HRTEM micrograph Of [Fe583(PEt3)x]y]o,2WS2, showing highly ordered in-plane structure. Large dark grey areas could be attributed to aggregated [Fe683121’ cores. 85 The triethyl phosphine ligands are more flexible than the triphenyl phosphine ligands because they are less hindered sterically; hence, a wider range of d—spacings are possible for any given orientation of the cluster. However, given the dimensions of the cluster, it is impossible to Observe these loadings in WS, (5 W atom per cluster) without loss of some of the triethyl phosphine ligands. It is barely possible to retain them all in the loading observed in MOS, (10.4 MO per cluster). The number of phosphines ligated to the clusters will be discussed in conjunction with the thermal analysis of the materials. Room temperature conductivity measurements on thin films of the materials indicated metallic behavior, with conductivity values around 0.4 S/cm. A HRTEM micrograph acquired at the edge of [Fe68,(PEt,),]o_,WS, particles dusted on a carbon-supported copper grid shows the in-plane structure, Figure 3.5. The in-plane structure in this sample is significantly more ordered than the structure visible in Figure 3.3b, and can be indexed as 110 planes. NO discrete scattering centers due to [FeGS,,]2+ cores are observed; however, larger dark gray regions are visible which may be due to aggregation of [Fe,,S,,]2+ cores. Aggregation would be consistent with the higher loading of [Fe,,S,,(PEt,),,]2+ clusters found in this sample (0.2 per W atom) than Co,,S,(PPh,)x cores in MOS, (0.02 per Mo atom). 86 [Ni,S,(PEt,)6],MS,. X-ray diffraction patterns of Ni,S,,(PEt,),,2+ encapsulated in MOS, and WS, are found in Figure 3.6. As with the other cationic clusters, the patterns are more well ordered than those from samples containing the neutral Co,,S,,(PPh,)6 cluster. Fourteen 001 reflections are Observed for are [Ni,S,(PEt,)6]2+ in MoS,, which exhibits an expansion of 14.2 A; and nine 001 for [Ni,S,(PEt,)6]2+ in WS,, which is expanded by 14.0 A. The dimensions of the cluster are approximately 13 A perpendicular to and 11.5 A parallel to the Ni,S, planes (Figure 3.7). EDS measurements indicated that the chemical formulas of the compounds shown in Figure 3.6 are [Ni,S,(PEt,)6]0mMoS, and [Ni,S,(PEt,)6]0m7WS,. There is no evidence of co-intercalation of the C10,; ion. These loadings represent the upper limit; lower loadings (down to ~0.04) have been observed. The maximum charge per Mo/W atom is 0.14/0.15, which is lower than that observed for samples containing the [Fe,,S,(PEt,)6]2+ and the [Al],O,,(OH),,,(H,O),,]7+ clusters. The dimensions, coupled with the observed loadings (13-14 Mo/W atoms per cluster), suggest that the clusters may be packed in the gallery space with the Ni,S, planes parallel to the layers. It is possible to accommodate the cluster in this loading without loss of phosphine ligands. At lower loadings smaller d-spacings have been observed (10.5 A - 13 A). 87 (A) 1- N O O O O A U) 1‘: C 3 S :3 CD ._ o [x e o 3 CU Vt!) " 88 2.2 - a ,3: Coco'- Pro . CO C v- V 8 ° 8 ° 3 a) o 0 fl 5 1111111111111141111111194411’111 10 20 30 40 50 60 70 29 (degrees) '- N 0 1° 3 A (D .2 C 3 Z‘ .2 E N 3 8 8 no 0 E‘ ,9 ° 8 a o ° m 0 fl 5 llllLJlLALLLJllllll lLJJlllllli 51015 20 25 30 35 40 29 (degrees) Figure 3.6. X-ray diffraction patterns of (A) [Ni,S,(PEt,)6]omMOS, (A = 14.2 A) and (B) [Ni,s,(PE1,),],,,,ws, (A = 14.0 A). Again the predominance and number of observed 001 reflections indicates a greater degree of order than in [C068,(PPh,)6]o.,6MOS,. 88 (B) Figure 3.7 . A view of the [Ni,,S,,(PEt,),,]2+ cluster with its Ni,S, planes (A) parallel to and (B) perpendicular to the MS, layers. Orientation (A) gives rise to a larger expansion which is consistent with the d—spacing Observed in the X-ray diffraction pattern. 89 3.3.2 Thermal Analysis [Co,S,(PPh,)6]o.o,,MOS,. The TGA of unintercalated Co,,S,(PPh,)6 (Figure 3.8a) shows a clean weight loss, beginning at about 250°C, corresponding to the departure of the triphenyl phosphine ligands (expected weight loss, 72%; observed, 70.6%). The identity of the volatile product was confirmed by direct probe mass spectrometry, and the nonvolatile decomposition product is CoS.'8 In contrast, the TGA of [C0,,S,(PPh,)6]0.0,,MOS2 shows a broad, continuous weight loss, beginning at about 220°C, to 650°C. Our hope was that the triphenyl phosphine ligands could be removed from the cluster by thermal treatment, resulting in a microporous sulfide (Figure 3.9). If all six phosphine ligands are still attached to the cluster core, the expected weight loss due to the departure of triphenyl phosphine ligands is 16.6%; the Observed weight loss is 18.9%. Although additional weight loss might be explained by co-encapsulated solvent, direct probe mass-spectrometry indicates that the volatile product is actually triphenyl phosphine sulfide. Predicted weight loss for triphenyl phosphine sulfide is 18.7%. In fact, when the selenide analog Co,,Se,,(PPh,)6 is encapsulated the volatile product is still triphenyl phosphine sulfide. This surprising result suggests that the phosphine ligands are extracting sulfur from MOS, as they depart (Figure 3.10). 90 (A) 120 100 - 80- 60*- °/e weight loss 40L lllll[Ill—LlJllllllllllell—LJ 20 “4‘ 0 100 200 300 400 500 600 Temperature (°C) (3) 105 100 - 90+ % weight loss 80 l I l l l l l l l l l l l l l l l l I l l l l L l l l l l l l l 0 100 200 300 400 500 600 Temperature (°C) Figure 3.8. TGA of (A) CO6S,(PPh,)6. The PPh, ligands depart in a clean step just below 300°C (expected weight loss, 72%; observed, 70.6%). (B) [Co,S,(PPh,)6]0,,MoS,. The PPh, ligands begin to depart at a comparable temperature, but the weight loss is gradual to 650°C (predicted weight loss due to PPh, departure, 16.6%; observed, 18.9%). 91 HEAT TREATMENT Figure 3.9. An illustration of the original plan to create a microporous sulfide: encapsulation of the cluster with phosphine ligands acting as “spacers”, followed by thermal treatment to remove the phosphine ligands. 92 HEAT TREATMENT Figure 3.10. An illustration of the observed data: encapsulation of the cluster with phosphine ligands acting as “spacers”, followed by thermal treatment. The phosphine ligands attack the layers as they depart, forming SPR3 species and possible 82' anion vacancies. 93 Intensity (arbitrary units) 20 (degrees) Figure 3.11. X-ray diffraction patterns upon thermal treatment of [CO6S,(PPh,)6]0,4MOS,. [C068,(PPh,)6]0,4MOS, becomes amorphous upon heating to 250°C, then collapses to a d-spacing only slightly larger than restacked MOS, by 650°C (b). 94 X-ray diffraction studies upon thermal treatment show that the material becomes amorphous after onset of the removal of the phosphine ligands; if heated to 650°C, the material collapses to a product with a d- spacing only slightly larger than restacked MOS, (Figure 3.11). It is likely that, as vacancies are created in the MOS, layers by the departing phosphine ligands, sulfur atoms are transferred from the [C0688] cores: [CO6S,(PPh,)6],MoS, --> [C06S,]XMOS,,Y --> [C0,,s,,_,],Mos2 (Eq. 3.1) [FeGS,(PEt,),],MS,. The TGA of [Fe68,(PEt,)6](BPh4), (Figure 3.12a) shows weight loss, beginning shortly after 100°C, due to the departure of the BPh,‘ anions as well as the triethyl phosphine ligands (expected weight loss, 69.5%; observed, 69.2%). The majority of the triethyl phosphine appears by mass spectrometry to depart as a non-sulfide product. [Fe6S,(PEt,),]o.o,6MoS, and [Fe6S,(PEt,),]o,WS,, however, show a gradual weight loss starting at about 140°C for the MoS,-containing sample and 150°C for the WS,-containing sample (Figure 3.12b). No BPh,‘ anions were detected by mass spectrometry. If all six phosphine ligands were present in each sample, and the volatile product was only triethyl phosphine, the predicted weight loss for each compound would be 23.9% and 28.0%, respectively; the observed values are 12.7% for [FeGS,(PEt,),]O.096MOS, and 18.7% for [Fe6S,(PEt,)x]o_,WS,. It seems that, as 95 predicted based on the loadings and dimensions of the cluster, the [Fe,,S,]2+ core is not fully ligated. Furthermore, direct probe mass spectrometry revealed that the majority of the volatile product is triethyl phosphine sulfide, analogous to the behavior of the triphenyl phosphine ligand in [Co,S,(PPh,)6]O.O,,MoS,. For [Fe6S,(PEt,)x]o_096MOS,, if there are two triethyl phosphine ligands still attached to the cluster the predicted weight loss (due to SPEt, loss) is 12.0%. For [FeéS,(PEt,),]O_,WS,, if there are three triphenyl phosphine ligands still attached to the cluster the predicted weight loss (again due to SPEt, loss) is 20.7%. Loss of three phosphine ligands was Observed upon encapsulation Of the [Fe,,S,(PEt,)6]2+ cluster in TaS,.20 X-ray diffraction patterns probing the thermal stability of [FeGS,(PEt,),]O.o,6MOS, are found in Figure 3.13. The material remains pillared up to 100°C (13b), the temperature at which MOS, converts from an octahedral geometry about the ”Mo atom to a trigonal prismatic coordination environment. At 150°C, after which weight loss has begun, the powder pattern is amorphous (13c). By 200°C the pillars have collapsed, resulting in Fe,MoS,, a material with a d-spacing only slightly larger than restacked MOS2 (13d). [Fe6S,(PEt,),]0,WS, exhibits slightly better thermal stability. At 150°C the material is still expanded; however, by 210°C partial decomposition to Fe,WS, has taken place. 96 110 100 90 80 70 60 50 40 30 % weight loss 105 100 95 90 % weight loss 85 80 Figure 3.12. TGA of (A) [FeGS,(PEt,)6](BPh4),, which loses weight due to the departure of the PEt, ligands and the BPh4 anion (expected, 69.5%; observed, 69.15%) 1111111111111114 111_.J ' 0 100 200 300 400 500 600 Temperature (°C) (3) \ r \ \ __ .1....1141.I....1....1.LA11~.: 0 100 200 300 400 500 600 Temperature (°C) and (B) [Fe6S,(PEt,),]O_O,6MOS, (solid line) and [FeGS,(PEt,),]0,0WS, (dashed line). The volatile species, SPEt,, is lost more gradually than the PEt, ligands in the unintercalated cluster. 97 Intensity (arbitrary units) 1 CD > 5 1 0 1 5 2 0 29 (degrees) Figure 3.13. X-ray diffraction patterns upon thermal treatment of [Fe6S,(PEt,),]0.0,6MoS,. The room temperature pattern (a) resembles the pattern upon heating to 100°C (b), but becomes amorphous by 150°C (c) and collapses to FexMOS, (6.3 A) by 200°C ((1). 98 [Ni,S,(PEt,)6]o.o,.,MOS,. The TGA of unintercalated [Ni,S,(PEt,)5](ClO4), is found in Figure 3.14a. The material loses most of its weight in a sharp step beginning at approximately 240°C, but continues to lose weight gradually until 650°C. Theoretical weight loss due to departure of triethyl phosphine ligands is 41%; the observed weight loss is 56.6%. Direct probe mass spectrometry indicates that the volatile product from the thermolysis of the unintercalated cluster is predominantly triethyl phosphine sulfide; predicted weight loss due to loss of triethyl phosphine sulfide is 52.3%, much closer to the experimentally observed value. [Ni9S9(PEt,)6]O.047MoS, also loses weight gradually all the way to 650°C; weight loss begins sooner but the material exhibits a significant drop around 160°C, a slightly lower temperature than the sharp step observed for the unintercalated cluster. Direct probe mass spectrometry indicates that the volatile product from the thermolysis of the unintercalated cluster is also triethyl phosphine sulfide (predicted weight loss due to SPEt,, 13.4%; observed, 16.5%). The sulfur source in this case is probably the [Ni,S,]2+ core, as even the unintercalated cluster loses its phosphine ligands as SPEt,, leading to layer collapse. Additional weight loss may be due to co-encapsulated solvent. 99 (A) 100r 110 % weight loss llllllllll O 100 200 300 400 500 600 Temperature (°C) (B) 105 % weight loss .1 (D co o o 01 o I l I 06 (II I lllll 0 100 200 300 400 500 600 Temperature (°C) on O i" L l— i- l— b Figure 3.14. TGA of (A) [Ni,S,(PEt,)6](ClO,),. Volatile product, lost in a clean step before 250°C, is SPEt, (expected weight loss due to SPEt,, 52.3%; observed, 56.6%). (B) TGA of [Ni,S,(PEt,)6]0_0,,WS,. The volatile product is also SPEt, (expected loss, 13.4%; observed, 16.5%). 3.3.3 HDS Activity HDS, or hydrodesulfurization, is a catalytic process used to remove sulfur-containing impurities from crude oil. The removal of these impurities prevents the formation of sulfur oxides upon combustion, which are precursors to acid rain. For example, an efficient HDS catalyst converts dibenzothiophene to biphenyl: The effectiveness of an HDS catalyst depends not only on its activity, but also on its selectivity toward certain products. In the desulfurization of dibenzothiophene, the desired product is biphenyl. An undesirable product is cyclohexylbenzene: xs H, + ———> 0—0 + H,S 5 Eq. (3.3) because it consumes additional H, gas. The transition metal sulfides are a class of compounds which exhibit good HDS activity, with rhenium, osmium, and iridium sulfides being among the most effective.26 Unfortunately, those particular metal sulfides are rather expensive. It has been known for many years that mixed-metal sulfide catalysts such as Co or Ni promoted MOS, or WS, (over an alumina 101 support) can provide comparable or better activity” Although the more reasonable cost of these mixed-metal sulfides has resulted in widespread commercial use, the mechanism for their activity is not well understood, and a variety of models have been developed to try to explain their effectiveness” 29' 30' 3‘ As mentioned previously, the Co,,S,,(PPh,),5 and [Ni,S,,(PEt,),,]2+ clusters were chosen as pillaring agents due to the similarity in chemical composition of the pillared products to the mixed-metal HDS catalysts. Most Of the models place the active sites on the surfaces or edges of the particles; however, a material such as ours, in which the MS, layers have been propped apart with these clusters, may allow catalysis to take place in the interior of the particles, thereby enhancing the efficiency of the catalyst. Accordingly, the products were tested for HDS activity by reaction with dibenzothiophene (Eq. 3.2). The activities for a number Of reference materials were collected as well. The results are summarized in Table 3.1. The selectivity of each catalyst toward biphenyl is reported in parentheses (as a percentage) next to the overall activity. The results indicate that [Co,S,(PPh,)6]o.o,MOS, exhibits comparable catalytic activity to Crosfield 465 (Co-Mo) at 400°C. At 300°C its activity (38%) is actually better than the commercially available catalyst (20%). 102 Table 3.1. HDS catalytic activity and surface area (SA) of selected samples and references. Reported (as a function Of temperature) are the percent conversion of dibenzothiophene to a non-sulfur containing product, and the percentage of biphenyl in the desulfurized products (in parentheses)“ Catalyst 400°C 350°C 300°C 200°C SA (mi/g) [C068,(PPh,),,]O.O,MoS2 85 38 15 (7) 26.8 (50) (61) Crosfield 465 (Co-Mo) 82 56 20 <5 (80) (91) (85) 0.2[COGS,(PPh,)6] + 25 10 <5 <5 9.3 restacked MOS, (73) (76) [Ni989(PEt3)6]0.04MOS2 58 35 32 (24) (17) (16) 1Ni.S.(PEt.>.1...WS. 44 41 39 (8) (21) (16) Crosfield 504 (Ni-MO) 98 58 27 10 (44) (46) (76) (55) [A1,,O,(OH),4(I-I,O),,]O_O,MOS, 6 (89) <5 <5 <5 10.0 restacked MOS, 22 9 (78) <5 <5 10.0 (75) 2H-MoS, 8 (88) <5 <5 <5 5.8 aRun-to-run conversion reproducibility = :1: 15% (e.g., 20 :l: 3%). Amount of “active” metals (Co + Mo) in each run = 0.1 g in all cases. 103 Unpromoted 2H-MOS, is clearly an inferior catalyst, with an activity of only 8% at 400°C. Restacked MOS,, although better than 2H—MOS, (22%), is still not as effective as [C06S,(PPh,)6]0,O,MOS,. In addition to these reference samples, a material consisting of a physical mixture of restacked MOS, and the Co,,S,(PPh,),5 cluster in a ratio comparable to the stoichiometry of [CO6S,(PPh,)6]O_O,MoS, was prepared and tested (0.2[Co,S,(PPh,)6] + restacked MoS,). The results are only slightly better than restacked MOS, (25%), verifying that the increased catalytic activity in [CO6S,(PPh,)6]0.O,MoS, requires the insertion of the cluster between the layers. Unfortunately [Co,S,(PPh,)6]o_0,MOS, is not as selective toward biphenyl as Crosfield 465. As discussed previously, the structure of [Co,S,(PPh,)6]0.O,MoS, under HDS conditions is ambiguous. HRTEM micrographs acquired after the catalysis suggest that some regions remain pillared whereas other regions have collapsed (Figure 3.15), but this method cannot probe the structure of the bulk of the sample. If in fact the sample is still pillared, as the micrograph suggests, the increase in activity may be due to the larger surface area of the sample (26.8 mZ/g) as compared to the surface areas of pristine MOS, (5.8 mzlg), restacked MOS, (1.0.0 mz/g), [Al,,O,,(OI-l),,,(I-I,O),,]0_(,,MOS2 (10.0 m2/g), and (0.2[COGS,(PPh,)6] + restacked MoS,) (9.3 mzlg). Also, if one compares the surface areas of the 104 . . ) , , . , a, . A, Figure 3.15. HRTEM micrograph of [C06S3(PPh3)6]o_02M082 after HDS. Although some regions do not appear to be expanded (black arrow), other regions are still separated (white arrow). The observed d-spacing, 14.7 A, corresponds to an expansion of 8.5 A, consistent with the dimensions of the C0683 core. 105 samples to their selectivities toward biphenyl, it appears that higher surface area coincides with lower selectivity. This possible relationship between surface area and selectivity supports the rim-edge model31a of the active sites. In this model all catalysis takes place on the exterior of the particle, not on the basal planes but at the perimeter of each layer. There are two different kinds of active sites: the “edge” sites and the “rim” sites. To understand the difference between the two, consider a stack of MOS, layers (Figure 3.16a). The perimeters of the layers at the top and bottom of the stack are the “rim” sites (much like the rim of a soup can). The perimeters of the interior layers, which comprise the bulk of the active sites in Figure 3.16a, are called “edge” sites. Unlike the “edge” sites, which are sterically hindered on both top and bottom by the presence of other layers, the “rim” sites are unrestricted on one side. According to the authors of this model, selective HDS takes place at “edge” sites, whereas HDS combined with hydrogenation takes place at the “rim” sites. The exfoliation/flocculation process used to synthesize restacked MOS, (and [C0688(PPh3)6]0'0,MOS,) decreases the particle size (and increases the surface area) relative to 2H- MOS,, resulting in more “rim” sites, which may explain why restacked MOS, shows higher activity (but less selectivity) than 2H-MOS,. Furthermore, by propping the layers apart with the Co,,S,,(PPh,)6 cluster, 106 (A) “rim” site “rim” site 1 K _ — ‘— M82 layer “edge” — sites — — _ — “rim" site‘ “rim” site (B) “rim” site “rim” site \A .. - .._ . . Q—MSzlayer “rim” sites[ _, 2 _ 1 ° ‘-——clusters — “rim” site‘ ~‘rim” site Figure 3.16. Illustration of the “rim-edge” model for (A) unpillared and (B) pillared MOS,. In unpillared MoS,, the “rim” sites are less selective toward biphenyl. In pillared MOS, the propping apart of the layers results in a higher number of “rim” sites, reducing the selectivity of the catalyst. 107 the surface area increases and even more “rim” sites are generated, which also results in a decrease in the selectivity (Figure 3.16b). The rim-edge model does not address the effects of CO- or Ni- containing species on the activity, however. It is interesting to note that, although we considered the attack of the layers by the departing PPh3 ligand to be a serious liability, in fact it may promote the activity of the material. In the “anion vacancy” model, the HDS active sites are sulfur vacancies in MOS, which are stabilized by the promoter species. Although sulfur vacancies induced by phosphine ligand evolution could explain increased activity in [CO6S,(PPh,)6]o_o,MOS,, it does not explain the increase in the activity of restacked MOS, relative to 2H- MOS,. The activities of [Ni,S9(PEt,)6]0.0,MOS, and [Ni9S9(PEt,)6]0_0,WS, were disappointing in comparison to the activity of [CO6S,(PPh,)6]0.0,MoS,. Although they are somewhat increased relative to restacked MOS, (58% and 44% over 22%), they are far lower than the activity of the commercially available catalyst Crosfield 504 (Ni-Mo) (98%) and their selectivity is the worst of all the catalysts. This behavior can be understood in terms of several factors. First, at HDS temperatures the encapsulated cluster has degraded (as discussed previously), most likely resulting in a material which is no longer pillared. Second, because the sulfide source of the 108 volatile triethyl phosphine sulfide is the [Ni,S,]2+ cluster core, no additional sulfur vacancies are created by the departing ligand upon thermal treatment. The fact that there is any increase in the activity relative to restacked MOS, is more of a reflection on the effects of the Ni- containing species on the activity than it is on the rim-edge model or the anion vacancy model. 3.4 Conclusions In conclusion, CO6S,(PPh,)6, [Fe6S,(PEt,)6]2‘° , and [Ni,S9(PEt,)6]2° can be encapsulated in MOS, and WS, and exhibit expansions of 14.0 - 15.4 A, 10.5 - 11.4 A, and ~14 A, respectively, which are consistent with the dimensions of the clusters. X-ray diffraction patterns from samples containing the neutral Co,,S,(PPh,)6 species are weaker than the patterns from samples containing the cationic clusters. Although X-ray diffraction and TGA studies suggest little loss of phosphine prior to thermal treatment in [CO6S,(PPh,)6]O_0,MoS,, HRTEM micrographs show loss of PPh, ligands, perhaps due to the ultrahigh vacuum or the limited applicability of the technique, which allows only the examination of the edges Of thin particles. Furthermore, expansion of the layers appears to be localized. TGA studies indicate the loss of PEt3 ligands prior to thermal treatment in [FeGS,(PEt,),]O.096MoS, and [FeGS,(PEt,),]0,WS,, and HRTEM micrographs 109 suggest some aggregation of the clusters. The orientation of the Co,,S,,(PPh,)6 or the [Fe,,S,(PEt,),]2‘° clusters in the layers is not known; it may change with loading and/or degree of phosphine coordination remaining in the encapsulated cluster. The orientation of the cluster in [Fe,,S,(PEt,),],,,,,TaS,20 is with its C, axis perpendicular to the MS, layers; this kind of arrangement makes sense, particularly at lower loadings, because it minimizes the Observed expansion. The rather large expansion Observed for [Ni,S,(PEt,)6]omMOS, and [Ni,S,(PEt,)6]0_o,,WS,, coupled with the high loadings of the cluster and TGA studies revealing no loss of PEt3 ligands prior to thermal treatment, imply that the clusters are packed with their Ni,S, planes parallel to the MS, layers. Metallic conductivity of the samples verifies that the layers have not structurally transformed from the octahedral coordination found in LiMoS, and restacked MS, to the trigonal prismatic coordination found in 2H-MS,. The ability of MOS, and WS, to encapsulate all three of these clusters is both remarkable and perplexing. What is the driving force for the reaction, particularly the encapsulation of neutral species such as the Co,S,,(PPh,)6 cluster? This ability of MOS, (and WS,) to accept neutral molecules initially convinced some scientists that the layers were in fact completely oxidized in the exfoliation process.9 However, increasing evidence of facile encapsulation of cationic species, Often with a high degree of order, suggests that the layers retain some negative charge. The 110 driving force for the encapsulation of cationic species would simply be an electrostatic interaction. The degree of charge remaining on the layers is difficult to determine from these samples, as a range of stoichiometries for all clusters can be Observed. Furthermore, residual negative charge on the layers may explain the encapsulation of species such as the [Fe,,S,(PEt,),,]2+ and the [Ni,S,(PEt,),,]2+ clusters, but it still does not explain the ability of these layers to also incorporate neutral species. Admittedly, the encapsulation of the neutral cluster is more difficult, requiring sonication and two days of stirring, and is sometimes difficult to reproduce. Encapsulation can be expedited by bubbling oxygen or air through the reaction mixture, or by (cautiously) introducing a neutral oxidizing agent such as 1,. This indicates that the layers may require further oxidation, either with time or with additional oxidizing agents in the reaction, before neutral species can be encapsulated. In addition, it is probable that small amounts of Li“ cations, which are extremely difficult tO detect, are co- encapsulated with the Co,,S,(PPh,)6 cluster. However, the driving force for neutral cluster encapsulation remains a mystery. Thermal treatment combined with mass spectrometry reveals that all phosphine ligands depart from the samples as phosphine sulfide. In [CO6S,(PPh,)6]o_o,MoS, and [Fe6S,(PEt,)y],MS, the phosphines appear to attack the layers as they depart; even the unencapsulated [Ni,S,(PEt,)6]2° 111 cluster decomposes to SPEt,, indicating that the sulfur source in all samples is the [Ni,S,]2+ core. The HDS activity of [C0658(PPh3)6]0.O,MOS, is comparable to the commercially available catalyst Crosfield 465 (Co-Mo); however, it is less selective toward desirable products such as biphenyl. Although the poor selectivity prohibits it from becoming commercially viable, the correlation between surface area and activity seems to support the “rim-edge” model to describe a mechanism for selectivity. In addition, the attack of the layers by the departing phosphine ligand does not seem to inhibit catalytic activity; in fact, it may help promote it. 112 References 1. Barrer, R.M. “Zeolites and Clay Minerals as Sorbents and Molecular Sieves,” Academic Press, New York, 1978, p407. 2. a) Pinnavaia, T.J. “Characterization of Catalytic Materials,” Wachs, I.E. (Ed.), Manning Publications Co, Greenwich CT, 1992, p149. b) Johnson, I.D.; Werpy, T.A.; Pinnavaia, T]. J. Am. Chem. Soc., 1988, 110, 8545. c) Pinnavaia, T.J. Science, 1983, 220, 365. 3. Kresge, C.T.; Leonowizc, M.E.; Roth, W.J.; Vartuli, J.C.; and Beck, J .5. Nature, 1992, 359, 710. 4. a) Hanko, J.A.; Kanatzidis, M.G. Angew. Chem. Int. Ed., 1998, 37 , 342. b), Stephan, H.-O.; Kanatzidis, M.G. J. Am. Chem. Soc., 1996, 118, 12226. c) Chou, J.-H., Kanatzidis, M.G. Chem. Mater., 1995, 7, 5. d) Chou, J.-H., Kanatzidis, M.G. Inorg. Chem, 1995, 33, 1001. e) Parise, J.B. Science, 1991, 251, 293. 5. a) Yaghi, O.M.; Sun, 2.; Richardson, D.A.; Groy, T.L. J. Am. Chem. Soc., 1994, 116, 807. b) Tan, K.; Darovsky, A.; Parise, J.B. J. Am. Chem. Soc., 1995, 117 , 7039. c) Ahari, H.; Garcia, A.; Kirkby, S.; Ozin, G.A.; Young, D.; and Lough, A.J. J. Chem. Soc., Dalton Trans, 1998, 2023. 6. a) Bedard, B.L.; Wilson, S.T.; Vail, L.D.; Bennet, J.M.; Flanigen, E.M. Zeolites, Facts, Figures, Future. Jacobs, P.A.; van Santen, R.A. (Eds) Elsevier Science Publishers 3.; Amsterdam, the Netherlands, 1989, 375-387. b) Jiang, T.; Lough, A.; Ozin, G.A.; Bedard, R.L.; Broach, R. J. Mater. Chem, 1998, 8, 721. c) Jiang, T.; Lough, A.; Ozin, G.A.; Bedard, R.L. J. Mater. Chem, 1998, 8, 733. d) Ahari, H.; Bowes, C.L.; Jiang, T.; Lough, A.; Ozin, G.A.; Bedard, R.L.; Petrov, 8.; Young, D. Adv. Mater., 1995, 7, 375. e) Ahari, H.; Ozin, g.A.; Bedard, R.L.; Petrov, S.; Young, D. Adv. Mater., 1995, 7, 370. f) Bowes, C.L.; Petrov, S.; Vovk, G.; Young, D.; Ozin, G.A.; Bedard, R.L. J. Mater. Chem, 1998, 8, 711. 7. a) Dhingra, S.; Kanatzidis, M.G. Science, 1992, 258, 1769. b) Chondroudis, K.; Kanatzidis, M.G. J. Am. Chem. Soc., 1997, 119, 2574. c) Sheldrick, W.S.; Braunbeck, H.-G. Z. Naturforsch, 1990, 45b, 1643. d) Marking, G.A.; Kanatzidis, M.G. Chem. Mater., 1995, 113 7, 1915. e) McCarthy, T.J.; Tanzer, T.A.; Kanatzidis, M.G. J. Am. Chem. Soc., 1995, 117, 1294. 8. Eisenmann, B.; Jackowski, J.; Schafer, H. Rev. Chim. Min., 1983, 20, 295. 9. Bowes, C.L.; Ozin, G.A.; Adv. Mater., 1996, 8, 13. 10. Li, H.; Laine, A.; O’Keefe, M.; Yaghi, O.M. Science, 1999, 283, 11. 12. 13. 14. 15. 16. 17. 18. 1145. a) Wachhold, M.; Rangan, K.K.; Billinge, S.J.L.; Petkov, V.; Heising, J.; Kanatzidis, M.G. submitted. b) MacLachlan, M.J.; Coombs, N.; Ozin, G.A. Nature, 1999, 397, 681. a) Whittingham, M.S.Prog. Solid St. Chem, 1978, 12, 41-99. h) Rouxel, J. “Layered Metal Chalcogenides and their Intercalation Chemistry.” Comprehensive Supramolecular Chemistry, Alberti, G.; Bein, T., Eds. Vol. 7, Pergamon Press, New York, 1996, p77-105. (a) Bissessur, R.; Kanatzidis, M.G.; Schindler, J.L.; Kannewurf, C.R. J. Chem. Soc. Chem. Commun. 1993, 1582. (b) Kanatzidis, M.G.; Bissessur, R.; De Groot, D.C.; Schindler, J .L.; Kannewurf, C.R. Chem. Mater. 1993, 5, 595. (c) Wang, L.; Schindler, J.L.; Thomas, J.A.; Kannewurf, C.R.; Kanatzidis, M.G. Chem. Mater., 1995, 7, 1753. Lemmon , J .P.; Lerner, M.M. Chem. Mater., 1994, 6, 207. (a) Divigalpitiya, W.M.R.; Frindt, R.F.; Morrison, S.R. Science 1989, 246, 371. (b) Divigalpitiya, W.M.R.; Frindt, R.F.; Morrison, S.R. J. Mater. Res. 1991, 6, 1103. Tigaya, H.; Hashmoto, T.; Karasu, M.; Izumi, T.; Chiba, K. Chem. Lett. 1991, 2113. Heising, J.; Bonhomme, F.; Kanatzidis, M.G. J. Solid State Chem, 1998, 139, 22. a) Bissessur, R.; Heising, J.; Hirpo, W.; Kanatzidis, M.G. Chem. Mater., 1996, 8, 318. b) Brenner, J.; Marshall, C.; Ellis,L.; Tomczyk,N.; Heising, J.; Kanatzidis, M.G.Chem. Mater., 1998, 10, 124-4. c) Bissessur, R. “Synthesis and Characterization Of Novel 114 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. Intercalation Compounds of Molybdenum Trioxide and Molybdenum Disulfide.” Ph.D. Dissertation, Michigan State University, 1994. a) Rudorff, W. Chimia 1965 19, 489. b) Bayer, E.; Rudorff, W. Z. Naturforsch 1972, 27b, 1336. c) SchOllhom, R.; Bethel, J.; Paulus, W. Rev. Chim. Miner., 1984, 21, 545. d) Miremadi, B.K.;. Morrison, S.R. J. Appl. Phys. 1988, 63, 4970. (a) Nazar, L.F.; Jacobson, A.J. J. Chem. Soc. Chem. Commun. 1986, 570. (b) Nazar, L.F.; Jacobson, A.J. J. Mater. Chem, 1994, 4, 149. Hong, M.; Huang, 2.; Lei, Z.; Wei, G.; Kang, B.; Liu, H. Polyhedron 1991, 10, 927. Agresti, A.; Bacci, M.; Cecconi, F.; Ghilardi, C.A.; Midollini, S. Inorg. Chem. 1985, 24, 689. Cecconi, F.; Ghilardi, C.A.; Midollini, S. Inorg. Chem. 1983, 22, 3802. Murphy, D.W.; DiSalvo, F.J.; Hull, G.W. Jr.; Wasczak, J.V. Inorg. Chem. 1976, 15, 17. Py, M.A.; Haering, R.R.. Can. J. Phys. 1983, 61, 76. a) Pecoraro, T.A.; Chianelli, R.R. J. Catal., 1981, 67, 430. b) Harris, 8.; Chianelli, R.R. J. Catal., 1984, 86, 400. c) Ledoux, M.J.; Michaux, O.; Agostini, G.; Panissod, P. J. Catal., 1986, 102, 275. Massoth, F.E. Adv. Catal., 1978, 27, 265. a) Lipsch, J.M.J.G.; Schuit, g.C.A. J. Catal., 1969, 15, 179. b) Massoth, F.E. J. Catal., 1977, 50, 190. a) Farragher, A.L.; Cossee, P. “Proceedings, 5th International Congress on Catalysis, Palm Beach, 1972”, Hightower, J.W., Ed., North-Holland, Amsterdam, 1973, 1301. b) Dungey, K.E.; Curtis, M.D.; Penner—Hahn, J.E. J. Catal., 1998, 175, 129. Topsoe, H.; Clausen, B.S. Catal. Rev. -Sci. Eng., 1984, 26, 395. 115 31. a) Daage, M.; Chianelli, R.R. J. Catal., 1994, 149, 414. b) Del Valle, M.; Cruz-Reyes, J .; Avalos, Borja, M.; Fuentes, S. Catal. Lett., 1998, 54, 59. 116 CHAPTER 4 Structure Of Restacked MOS, and WS, Elucidated by Electron Crystallography 117 4.0 Abstract There has been a lot of confusion about the nature of restacked MOS, and WS,. The structure has been proposed to be trigonal TiS, type with octahedral M4+ and called 1T-MoS,. The presence of a distortion in the metal plane which gives rise to a superstructure has been suggested. We have performed electron crystallographic studies on small (sub micron) single crystal domains of restacked WS, and MOS, in order to solve their superstructure. We find that what initially seems to be a trigonal crystal is actually a “triplet” of three individual orthorhombic crystals. Using two- dimensional hk0 data from films for both “triple” and “single” crystals we calculated corresponding Patterson projections, which reveal a severe distortion in the Mo/W plane, forming infinite zigzag chains. The projection of the structure suggests M-M distances of 2.92 A and 2.74 A for MOS, and WS,, respectively. Least squares refinement from the single crystal data gives R,=13.3% for WS, and R,=15.3% for MOS,. Therefore, we submit that restacked MOS, and WS, are not lT-form but rather WTe, type. 118 4.1 Introduction Due to a unique combination of valuable structural, electronic, and optical properties, the layered dichalcogenides have been studied and used for a litany of practical applications.”4 One Of the most versatile members of this class of compounds is MOS,. Found in nature in its 2H form as the mineral molybdenitez, its inexpensiveness and availability have permitted its use as a solid lubricant, a catalyst for hydrodesulfurization3, a host for intercalation chemistry and an electrode material for solid state batteries“. Many layered transition metal chalcogenides can be treated with n- Butyl lithium to form a reduced species in which lithium occupies the space between the layerss. (This ability to undergo reduction and accept lithium is one important reason why these materials have been investigated for solid state batteries.) The redox properties of these reduced dichalcogenides vary from one compound to the next, but some have the remarkable ability to form suspensions in water, much like clays. The chemistry of LiMoS, (and LiWS,) in water is particularly fascinating because, after undergoing a redox reaction with water to form H, (g) and LiOH, a suspension of colloidally dispersed single layers is formed, and the layers can remain separated from one another in water for days. The material can be “restacked” by filtration, precipitation, centrifugation, or 119 evaporation. Due to this remarkable ability, many guest species have been encapsulated between the layers.“8 There has been a lot of confusion about the nature of restacked MOS, and WS,. The presence of a distortion in the metal plane which gives rise to a superstructure has been suggested. The reduction of the 2H form appears to induce a structural transformation from trigonal prismatic coordination about the metal to octahedral coordination? The structure has been proposed to be trigonal TiS, type10 with octahedral M‘” and called 1T- MoS,. Originally, lT-MoS, was synthesized by the oxidation of K,(H,O),MOS, instead of the exfoliation/restacking method, but the products of both synthetic methods exhibit metallic conductivity and an exothermic transition around 90-100°C which suggests that they are identical.” The layered dichalcogenides are prone to a wide variety of structural distortions caused by Charge Density Waves (CDWs), resulting in significant changes in the properties of the materials.'2 The source of confusion is that several superlattices have been reported for lT-MOS, (Figure 4.1). 120 Figure 4.1. Schematic illustrating the proposed superstructures of restacked/lT-MOS,. The superstructures derive from bonding associations of metal atoms. (A) An ideal undistorted lattice, (B) a tetramerization, (C) a trimerization, and (D) zig-zag chain formation. Black circles represent Mo, gray circles, S. 121 EXAFS (Extended X—ray Absorption Fine Structure) analyses have indicated M-M associations, but are not able to provide a structural model. X-ray and electron diffraction studies of the restacked MOS, (and WS,) have indicated a 2a x 2a superstructure in the ab plane. A tetramerization model was proposed based on the X-ray diffraction studies.”"“ The electron diffraction studies on restacked WS,, however, found that the hexagonal symmetry was preserved, but the proposed tetramerization model is incompatible with this finding.15 To add to the confusion, another electron diffraction study on LiMoS, found the same 2a x 2a superstructure, but the authors attributed it to lithium ordering between the layers.16 Later, an STM study of the surface layer of restacked MOS, suggested a 2a x a superstructure caused by the formation of zig-zag chains.17 The original lT-MOS,, however, is reported to exhibit a fla x [[321 superstructure, which would be best described by a trimerization model?” In an effort to clarify the structure, we have performed 2— dimensional electron crystallographic studies on restacked WS, and MOS,. This method has enabled us to probe the bulk material, not just the local environments of the atoms or the structure at the surface. Furthermore, it has allowed us not only to directly determine the structure, but to refine it as well. 122 X-rays are scattered by the electron shells of atoms; electrons, on the other hand, are scattered by the electrostatic potential due to the atomic nuclei and their electron clouds. Despite the different origins of scattering, much of the theory developed for X-ray crystallography can be applied to electron crystallography and thus, in principle, the two techniques provide similar structural infonnation.‘9’2° Most experimentalists have avoided electron diffraction as a technique for structure solution, however, because electrons interact more strongly with the sample than X-rays; hence, multiple scattering events (secondary and dynamic scattering) may take place in electron diffraction.”20 Because a TEM (Transmission Electron Microscope) is capable of high magnification, it is possible to acquire data for many samples which cannot be investigated by single crystal X-ray methods?“20 Furthermore, the wavelength of an electron (at an accelerating voltage of l20kV) is 0.0335 A; as compared to 0.71 A for Ka Mo radiation, which (theoretically) can lead to. higher resolution data. Because the Ewald sphere is much flatter, many reflections can be observed simultaneously for one orientation of the crystalm"22 We have discovered that, despite the presence of Mo/W in our samples, quasi-kinematical data are acquired, from which a plausible projection of the structure can be obtained and refined. 123 4.2 Experimental LiMS, (M=Mo, W) was synthesized by reacting 2H-MS2 with excess LiBH4 at 300-350°C for 3 days. Restacked MS,“'13 was synthesized by reacting LiMS, with H,O, rinsing several times to remove the LiOH generated in the exfoliation process, and depositing the solution on a copper carbon coated grid. The grids were examined at 120kV on a JEOL 120CX TEM. Suitable crystals were located and their diffraction patterns captured on film. The negatives of the diffraction patterns were scanned into the computer at 600 dpi. The patterns were indexed to an orthorhombic cell with a =5.56A and b=3.21A, related to the subcell by the relation 3a,“, x am. The “Gel Plotting Macro” in NIH Image 1.60 was used to extract the integrated intensities (Im) from the patterns.23 Accurate cell parameters a and b were determined from powder diffraction data using a Rigaku-Denki/RW400F2 (Rotaflex) rotating anode powder diffractometer. The data were converted into .hkl file format for use in SHELXTL programs.24 The two-dimensional Patterson maps were calculated from the data by Equation 4.1: P(u,v) = %2:2|(hk)|2 cos 27t(hu + kv) Eq (4 l) h k . . 124 Where P corresponds to electron density overlap in the structure (i.e. the Patterson function), A is the unit cell area, and l(hk)l = 1“,.19 Least squares refinement of the structure in SHELXTL (version 5)24 were carried out after the coefficients for the electron scattering factors were obtained by fitting the sine/2t curves using the program Curve Expert.” 4.3 Results and Discussion 4.3.1 Data Collection Because restacked MOS, and WS, are layered compounds with plate- like morphology, they exhibit preferred orientation which causes the ab plane to be perpendicular to the electron beam, and the diffraction patterns consistently contained only hk0 data. Our initial efforts to locate a suitable crystal resulted in pictures similar to Figure 4.2a-4.2b. The patterns suggested a 2a x 2a superstructure. Upon consideration, however, we recognized that the exfoliation/flocculation process, involved in the synthesis of restacked MOS, and WS,, would result in turbostratic materials, or at least materials susceptible to stacking faults, and that this could result in pronounced twinning phenomena. We then targeted 125 Figure 4.2. (A) Bright field image and (B) selected area electron diffraction pattern from a "triple" WS2 crystal, giving rise to an apparent 2a x 2a hexagonal superstructure. (C) Bright field image and (D) selected area electron diffraction pattern from a "single" WS2 crystal, giving rise to a Zn x a or ‘13a x a superstructure. 126 a o 0 2a ° ° ‘3' o o . . . . a o o . o O . o . . g . o B O . o - O ......... ........... ............. ...... . . .. ...... ... ...... ... ...... ... ...... . . .... ............. ........... o.......... ......... Figure 4.3. (A) Relationship between the 2a x a and flu x a lattice. (B) Illustration of how three flu x a patterns can be overlapped to form a 2a x 2a pattern. Note that only the sublattice reflections overlap. 127 extremely thin crystals in our investigations and found that, in fact, restacked WS, and MOS, have a 2a x a superstructure (Figure 4.2c-4.2d), which is consistent with the results of the STM studies of restacked MOS,.l7 This 2a x a superstructure can actually be more simply described by an orthorhombic [Xi—a x a cell (Figure 4.3a). Note that a = a but a* ¢ a* in the two unit cells because of the change from a hexagonal to an orthorhombic crystal system. The 2a x 2a superstructure is in fact a “triplet” of three 1/3—a x a crystals rotated by 120 degrees relative to each other. This causes their diffraction spots to be aligned with respect to the sublattice reflections but not the superlattice reflections (Figure 4.3b). 4.3.2 Structure solution from the “triplet” crystal Of WS, Because a “single” crystal pattern of sufficient quality was not initially available, we proceeded with the structure determination using the twinned crystal depicted in Figure 4.2a-4.2b. Intensities of the reflections were extracted from the diffraction pattern using the “Gel Plotting Macro” of NIH Image 1.60.” This Macro was developed to calculate concentrations in gel electrophoresis experiments by a comparison of the 128 bands in a sample to the same bands in reference samples of known concentrations. In order to make this comparison it is necessary to consider not only how dark (or light) the band, but also the area occupied by the band. Similarly, the intensity of a reflection in an electron diffraction pattern is a function of the area of the spot as well as its optical density. NO external standard is required because only the “relative” intensities of the reflections are necessary for structure determination. Accordingly, the diffraction pattern was indexed to the three orthorhombic cells and the relative intensities of the reflections were extracted from the pattern (Figure 4.4). First the macro creates a one-dimensional set of peaks from the data; then the width and baseline of the peaks are specified and a numerical value for the integrated intensity is extracted. A 2-D rolling ball background subtraction suggested in the “Gel Plotting Macros” manual was applied, as well as a Lorentz correction recommended by early electron crystallographers.‘9 Intensities of reflections from all three of the “triplet” cells were extracted, but only the cell with the strongest reflections was used. The intensity of the subcell reflections were corrected for the twinning by dividing by three. 129 Figure 4.4. Illustration of the "Gel Plotting Macro" in NIH Image 1.60. (A) Lanes of reflections are specified, then (B) converted to a set of one- dimensional peaks from which the integrated intensities of the reflections can be extracted. 130 b axis a axis Figure c-axis: (A) ideal 1T-M82 (from simulated data) (B) restacked W82 with background subtraction and Lorentz correction (C) restacked WSz with only the Lorentz correction (D) restacked WS2 with no corrections. All experimental patterns indicate a deviation in the W—W atom vector. 131 A two-dimensional Patterson contour plot generated from the extracted data clearly indicates a substantial deviation in the metal atom position from the ideal position in the lT-TiS, structure type, which gives rise to a short M-M distance (Figure 4.5b). In an ideal 1T structure the electron density overlap at the center of the Patterson map, which is indicative of the M-M vector in the structure, is circular (Figure 4.5a). This translation of the metal atom along the /3_a axis can only be explained by the formation of zig-zag chains. This kind of distortion has been observed in WTe,,26 another layered dichalcogenide with (distorted) octahedral coordination about the metal atom, and represents a significant departure from the ideal structure, TiS,.lo Distortion of layered octahedral ML, (12 systems to form zig-zag chains due to a charge density wave has been predicted by Rovira and Whangbo.27 This kind of distortion gives rise to changes in the band structure of the material which are consistent with the properties we have observed.” The Patterson map in Figure 4.5b contains some strange features which had to be addressed before refinement of the structure was possible, however. The M-S vector peaks are very weak, and there are other puzzling features which seem to indicate short atom-atom vectors in the structure. These vectors became more pronounced with the elimination of higher order reflections. They disappeared completely, however, when the 132 2-D rolling ball background subtraction was eliminated from the data extraction procedure. Electron diffraction patterns have a rather high, sloping background due to diffuse scattering which occurs in all samples. The peaks closest to the main beam are most severely affected. A background subtraction seemed an ideal manner to address this problem; however, use of the background subtraction method in the “Gel Plotting Macro” Obviously results in perturbation of the data, particularly the data closest to the main beam, because the contribution of the background is less significant in gel electrophoresis samples than in electron diffraction patterns. The background can be removed manually (with much better results) during the step in which the baseline is specified for each peak. Also, data from two different camera lengths were combined. Use of a larger camera length allows reflections to be collected further away from the main beam on the negative, whereas the smaller camera length allows higher order reflections (which provide better resolution) to be collected. The modified data set gave rise to the Patterson map found in Figure 4.5c. Least squares refinement of this data set in SHELXTL (version 5)24 against a WTe, type model gives an R, value of 36.8% (Tables 4.1 and 4.2). Structures with comparable R, values solved by electron diffraction have actually been published by early electron crystallographers.19 Such a high R, value is considered acceptable for many systems due to the 133 Table 4.1. Crystal data and structure refinement for restacked WS,. Empirical formula Formula weight Temperature Wavelength Crystal system Plane group Unit cell dimensions Area (A2) z 6 (000) 20 range for data collection Data resolution Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on (1)2 R indices (all data)“ we, 248.00 293 K 0.0335 A Rectangular pg (#4) a=556A b=321A y = 90 deg. 17.295 2 46 0.36 to 4.20 deg. 0.09 to 1.1 sin0/A -12<=h<=10, -5<=k<=5 173 99 Full-matrix Is. on (I) 2 9910/7 0.874 R, = 0.3683 “R, = 201601 - K1¢,11)/ 21601 134 Table 4.2. Preliminary atomic coordinates ( x 10°) and equivalent isotropic displacement parameters (A2 x 103) for restacked WS,. x y U(eq) Occ. W( 1) -2022 0 l 1 S( 1) 423 1 0 724 1 8(2) 997 0 1 100 1 U(eq) is defined as one third of the trace of the orthogonalized Uij tensor. susceptibility of the electron diffraction data to perturbation by multiple scattering events, as mentioned previously. In order to improve the refinement, we attempted a two-beam dynamic scattering correction. 135 4.3.3 Dynamic Scattering: Two-Beam Approximation Dynamic scattering has inhibited the widespread use of electron diffraction data for structure determination, particularly in samples containing heavy elements. A schematic depicting the perturbation of the electron diffraction data by dynamic scattering is found in Figure 4.6. In the ideal situation (found in X-ray diffraction), called “kinematic” scattering, the radiation source is diffracted only one time as it passes through the sample. Due to the strong interaction of the electron beam with the sample, however, electrons can be diffracted many times as they pass through the sample. The thicker the sample, the greater its susceptibility to multiple scattering events. One can then consider the intensity of each diffracted beam to be “contaminated” with intensity from the other diffracted beams, and also “contaminated” due to the loss of some of its original intensity to other reflections: I = (Ikin - Ilost ) + I,gained Eq’(4'2) dyn Historically, correction for this phenomenon has been attempted in more than one way. The two-beam approximation tries to simplify the problem by treating the data as if the pattern comes from a “mosaic” of small crystals, each with a unique and perfect orientation such that only that one 136 A main beam sample diffracted diffracted beams beams B main beam sample diffracted diffracted beams beams Figure 4.6. Schematic illustrating (A) kinematic scattering, in which the electron beam is diffracted only once as it passes through the sample and (B) dynamic scattering, in which the beam is diffracted multiple times, resulting in “contamination” of the intensity of all the reflections. 137 reflection is produced per crystal. Each reflection therefore functions completely independently of all the other reflections in the pattern.'9'”'35 In this case, the intensity of each diffracted beam can be conceptualized as 1d,“: Ikin - 1,0,, Eq.(4.3) If one imagines I,ost tO be a constant fraction (say, 30%) Of the original intensity of each reflection, it becomes clear that the strongest reflections will be most noticeably affected. The mathematics are actually somewhat more complicated, but the net effect is a homogenization of the diffraction pattern. In order to adjust for this ‘homogenization’, a method has been developed to correct the experimental data.” It operates under the assumption that most reflections scatter kinematically, but a few of the most intense reflections exhibit “deviations toward dynamic scattering.”19 The following equation applies in a kinematically scattering crystal: I = Q24 Eq.(4.4) in which Q is a normalized structure factor”0 and .4 is a Lorentz correction (in this case, to divide each reflection by its d-spacing). In a dynamically scattering crystal the equation is slightly modified: I=Q°A3K(A3Q)-4 Eq.(4.5) such that it now includes a term for the thickness of the crystal (A,) and a function K(A,Q), shown in Figure 4.7, which describes dynamic scattering 138 as a function of the product of the thickness and the normalized structure factor. 0:1 . . . . . . A . 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 A30 Figure 4.7. Graph of the function K(A,Q), which describes dynamic scattering as a function of the thickness (A,) of a crystal and the normalized structure factor (Q) according to the two-beam approximation. In a kinematically scattering crystal K = 1. After a preliminary refinement has been conducted, the calculated structure factors are used to determine an experimental K value for each reflection: (I.../ 49692.... = K... Eq.(4.6) If Kexp < 1 the reflection is considered to be decreased in intensity due to dynamic scattering. The purpose of calculating Kexp is purely diagnostic, and Kexp < 1 values are only significant for reflections of large intensity. Once the reflections in need of correction have been determined, (I)calc and the thickness (A,) of the crystal can be used to calculate a theoretical A,Q value for each reflection, and the graph in Figure 4.7 is 139 used to find a Km value. The corrected intensity can then be calculated by the following relation: Icorr = (Iobs/ A/Kcon’ Eqfl(47) Two major drawbacks of this method (aside from the fact that the 1’,ained term is neglected) are that some knowledge of the structure is required to correct the data, and the thickness of the crystal must be determined. Unfortunately the thickness of the “triplet” crystal used in this structure determination was not known. Several different X-ray diffraction patterns containing 001 reflections were acquired, and Debye-Scherrer line broadening31 was used to estimate an average thickness for the particles: th, = 0.9 x A, x 57.3 [3,,, cos0 Eq.(4.8) where B,,, is the full width at half maximum of the 001 reflection. The results indicated a thickness range of 150 A to 220 A. Km values were calculated for 12 reflections at 150 A and up to 34 reflections at 220 A, new Patterson maps were generated, and refinements of the corrected data were conducted. 140 Table 4.3. Refinement for WS, as a function of data correction methods. R,a Goodness of Fit (GOF) Original data 0.3683 0.874 (no background subtraction) Dynamic scatterin correction 0.3580 0.871 (2 beam) t= 150 Dynamic scatterin correction 0.3161 0.756 (2 beam) t = 220 Lorentz correction eliminated 0.2454 1.301 1R1 = 201601 - 106,10 / 216,1 The results (Table 4.3) indicated little improvement in the refinement at either thickness (from 36.8% to 35.8% and 31.6%). This led us to re-assess the quality of our integrated data prior to the dynamic scattering correction. As mentioned previously, only data from the strongest supercell was used in the refinement, and subcell reflections were divided by three to correct for the twinning. In reality, this correction results in an underestimation of intensity values of the subcell reflections. The relative intensities of the three different supercells indicated that 50% of the intensities of the subcell reflections are due to that particular supercell. Accordingly the data was adjusted, but the refinement was not much improved (33.66%). 141 Next the Lorentz correction suggested by Vainshtein was eliminated. This correction, which is simply to divide the intensities of the reflections by their d-spacings, is a crude method to compensate for the curvature of the Ewald sphere (Figure 4.8). Because most crystals examined by TEM are very thin, their reciprocal lattice points are more “cigar”-shaped than spherical along the axis parallel to the electron beam. At the center of the negative the Ewald sphere intersects the lattice points through the center of the cigars, capturing their maximum intensity. Towards the edge of the negative, when the Ewald sphere has begun to curve away from the plane, the sphere no longer intersects the lattice points through the center of the cigars. The end result is that the intensities of the reflections at the limit of the Ewald sphere, which have the smallest d-spacings, are underestimated. In the early days of electron diffraction, smaller accelerating voltages (which give rise to smaller Ewald spheres) were used; hence the effect of the curvature was more significant and a d-spacing correction was more important. Most modern electron microscopists no longer apply this correction, however.” With the elimination of the Lorentz correction the intensities of the vectors in the Patterson contour plots generated from the uncorrected data increased, (Figure 4.5d) and the refinement improved to 24.54% (Table 4.3). 142 -_‘_ _—“‘ Figure 4.8. The intersection of the Ewald sphere with the reciprocal lattice at an accelerating voltage of (A) ~50kV and (B) ~100kV. at 100kV the Ewald sphere is much flatter, and the effects Of its curvature become significant at smaller d-spacings than at 50kV. 143 1 \ ii 011‘ -< B O o O 0 ° 0 O . O O 0 ° 0 O O . O O O o ° 0 O O . O . O 0 ° C O O O O o O . O . O . O O O 0 ° 0 ° 0 ° C O O 0 ° 0 ° C 0 ° 0 ° 0 9 O 0 ° 0 ° 0 Figure 4.9. (A) The limit of detectable reflections (in a perfectly oriented crystal) as a function of the thickness (t) of the sample (q = 1/t), the radius of the Ewald sphere (r), and angle 7. (B) An illustration of secondary scattering, in which the diffracted beams from one crystal act as “main beams” as they pass through a second crystal, resulting in the appearance of reflections with artificially small d-spacings. 144 Based on the accelerating voltage and the d-spacing of the highest order reflections Observed it is possible to estimate the thickness (t) of the crystal (Figure 4.9a) through the following relationships: d*/2r = sin'y tan'yd* = q Eq.(4.9) where d* is the reciprocal of the d-spacing of the outermost observed reflection, r is the radius of the Ewald sphere (29.8 A"), and q is 1/2 the length Of the cigar and is equal to the reciprocal of the thickness (t). Based on this calculation, the “triplet” crystal would have to be 4 layers thick or less. This immediately seemed suspicious because in order to observe these higher order reflections it was necessary to examine twinned crystals of “intermediate” thickness, indicating that the crystal is probably more than 4 layers thick. Therefore it seems that, in addition to dynamic scattering, the data is perturbed by secondary scattering. Secondary scattering, like dynamic scattering, involves multiple scattering of the electrons as they pass through the sample. Unlike dynamic scattering, however, which is a multiple scattering event within a single crystal domain (or, in the approximation, a “mosaic” of independently scattering crystals forming a nearly single crystal domain), secondary scattering involves multiple scattering events encompassing two or more crystals which are significantly dislocated relative to one another. It is possible that the two crystals could even be two different compounds. 145 Consider two stacked layers rotated by 120 degrees relative to one another, as in the “triplet” crystal. The electron beam is scattered at least once as it passes through the first crystal; but before the pattern reaches the negative the undiffracted and diffracted beams must pass through the second crystal, where they can act as “main beams” and be scattered again. These new “main beams” each have their own Ewald sphere that is slightly shifted relative to the original main beam, which results in a pattern which appears to have higher resolution data than is possible (Figure 4%) given the thickness of the sample and the curvature Of the Ewald sphere associated with the original main beam. Secondary scattering can be treated mathematically; however, when one combines correction of the data for secondary scattering as well as dynamic scattering the reliability of the experimental data becomes questionable. For this reason, we decided to concentrate our efforts on patterns which were relatively “single” crystal. 4.3.4 Structure solution from “single” crystals of WS,. Data was extracted from the diffraction pattern of restacked WS, in Figure 4.2d, and also the patterns from restacked WS, in Figure 4.10, using the Gel Plotting Macro as described previously. All Patterson maps generated from these data sets indicate a deviation in the M-M vector from 146 the ideal position (Figure 4.11). The other peaks in the Patterson maps, which contain information about the M-S vectors, are elongated because there are two non-equivalent metal atom positions. Least squares refinement Of the data from Figure 4.2d in SHELXTL (version 5)” gave rise to an R, value of 13.3% (Tables 4.4,4.5). The structure is non-centrosymmetric, belonging to the plane group pg (#4). Due to a convention regarding the cell parameters in this plane group, which contains a glide plane along one axis, l/Ta = am, and a = b,,,,,. The structure, shown in Figure 4.12, contains the W atoms in distorted octahedral coordination, forming zigzag chains. The structure of WTe,, which is very similar, is also shown.” It is interesting to note that, in WTe,, the W atoms are also slightly distorted along the c-axis of the structure. Because the electron diffraction data contains only hk0 reflections, the refinement of the structure of ,WS, produces only a projection of the layer structure. The projection of the short W-W distance in the structure is 2.74 A, significantly shorter than the 3.16 A distance found in the ideal 2H-WS, structure. 147 Figure 4.10. Slected area eelctron diffracotin pattern from "single crystal" restacked WS2 used to generate Patterson contour plots. 148 Figure c-axis: (A) ideal lT-M82 (from simulated data) (B) restacked WS2 from the data in Figure 4.2d (C) restacked WS; from the data in Figure 4.10a (D) restacked W82 from the data in Figure 4.10b. All patterns indicate a deviation in the W-W atom vector. 149 Table 4.4. Crystal data and structure refinement for crystals of WS,. Title Empirical formula Formula weight Temperature Wavelength Crystal system Plane group Unit cell dimensions Area (A2) Z (I) (000) 20 range Data resolution Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on (D2 R indices (all data)“ “single” WS, 248.00 293 K 0.0335 A Rectangular Pg (#4) a = 5.56 A b = 3.21 A y: 90 deg. 17.295 2 46 0.36 to 2.52 deg. 0.09 to 0.66 sine/A --7<=h<=7, 3<=k<:3 78 43 [R(int) = 0.1556] Full-matrix 1.8. on (1)2 43/0/5 1.305 R, = 0.1328 “triple” ws, ws2 248.00 293 K 0.0335 A Rectangular Pg (#4) a = 5.56 A b = 3.21 A Y = 90 deg. 17.295 2 46 0.36 to 4.20 deg. 0.09 to 1.1 sine/A —12<=h<=10, -5<=k<=5 173 99 [R(int) = 0.0870] Full-matrix 1.8. on (1)2 99/0/7 1.301 R, = 0.2456 “R, = 2(ll,| - 106,11) / 216,1 150 Table 4.5 . Atomic coordinates ( x 10°) and equivalent isotropic displacement parameters (A2 x 103) of restacked “single” and “triple” WS,. x y U(eq) Occ. “single” WS, W(1) -2021 0 2 1 S(1) 4290 0 8 1 8(2) 950 0 8 1 “triple” WS, W( 1) -2066 0 4 1 8(1) 4182 0 105 1 8(2) 883 0 127 1 U(eq) is defined as one third of the trace of the orthogonalized Uij tensor. 151 A Z‘9 ' zag . H“? O chain Twfifi? tat-st b axis 7%“;gfg,‘ may. .Ilk.’ 1‘. O . . B zig - zag ‘ chain .‘V’A‘Y’AV’. ii b axis .’/3%‘ 741% We$. E M82 layer Figure 4.12. (A) Two-dimensional projection of a restacked MS, layer (M=MO,W) and one WTe, layer. (B) View of WTe, parallel to c-axis. Dark Circles are M; gray circles, S; open circles, Te. 152 4.3.5 Dynamic Range Correction One of the challenges in the data collection for this system is balancing the dynamic range in the intensities of the reflections with the limited capability of the film. In the electron diffraction pattern depicted in Figure 4.2d, a number of the strongest peaks (usually sublattice peaks) are overexposed, creating a “flat top” on the negative (Figure 4.13a), but the exposure is a good length of time to capture most of the weaker reflections (Figure 4.13b). Conversely, if a shorter exposure time is used as in the electron diffraction patterns in Figure 4.10, the strongest peaks will not have a “flat top” (Figure 4.13c), but the weaker peaks, which encompass most of the superlattice reflections, are so weak that it is extremely difficult to evaluate their relative intensities (Figure 4.13d). For this reason, the data collected at longer exposures were corrected for this “flat top” problem. The intensities of the (1020), (13110), and (1300) reflections (8 total, 5 unique) were increased by 30%. Least squares refinement against corrected and uncorrected data sets revealed no differences in positional parameters and only minor changes in thermal parameters; the net result was a virtually identical solution with slightly lower Rm, (Am=0.5%); R, (Am=1.6%); and wR, (AM=8%) values. 153 Figure 4.13. Surface plot of the (3-10) (A) and the (-2-20) (C) reflections in the electron diffraction pattern in Figure 4.2d. The (3-10) has a "flat top" due to overexposure. Surface plot of the (3-10) (B) and the (-2-20) (D) reflections in the electron diffraction pattern in Figure 4.1 la. The (3-10) is not overexposed, but the (-2-20) is very weak. p—i 54 4.3.6 Structure Solution from “single” crystals Of MOS,. Data was also collected for restacked MOS, (Figure 4.14). The lighter MO atom does not scatter as well as W; hence, it was more difficult to find “single crystal” patterns which diffracted well enough for a refinement. Patterson maps calculated from both diffraction patterns in Figure 4.14 indicate a distortion in the M-M vector along the WT: axis, forming an oval (Figure 4.15). In restacked WS, the distortion is more pronounced, as the M-M vector has split into two resolvable peaks which correspond to two different M-M vectors. Least squares refinement in SHELXTL (version 5)24 was conducted on the data shown in Figure 4.15b, and gave rise to an R, value of 15.3% (Tables 4.6, 4.7). The short MO—Mo distance is 2.92 A, longer than the value found for restacked WS, (2.74 A) but still significantly shorter than the ideal MO-MO distance in 2H-MOS, (3.16 A). 155 Figure 4.14. Bright field images (A,C) andselected area electron diffraction patterns (B,D) of restacked M082 , giving rise to 21 2a x a or a 13a x a superstructure. 156 a axis Figure 4.15. Two-dimensional Patterson projections along the c-axis: (A) restacked M082 calculated from the pattern in Figure 4.14b. (B) restacked MOS2 calculated from the pattern in Figure 4.14d. 157 Table 4.6. Crystal data and structure refinement Of restacked MOS,. Empirical formula Formula weight Temperature Wavelength Crystal system Plane group Unit cell dimensions Area (A2) Z (I) (000) 20 range for data collection Data resolution Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on (D2 R indices (all data)“ MOS, 160.00 293 K 0.0335 A Rectangular pg (#4) a = 5.47 A b = 3.16 A y = 90 deg. 17 .295 2 40 0.36 to 2.78 deg. 0.09 to 0.73 sine/2t -7<=h<=5, -3<=k<=3 73 46 [R(int) = 0.2120] Full-matrix 1.8. on (I)2 46/0/7 1.267 R, = 0.1525 “R, = 201601 - 106,11) / 2|°| 158 Table 4.7. Atomic coordinates ( x 10°) and equivalent isotropic displacement parameters (A2 x 103) in restacked MOS,. x y U(eq) Occ. Mo( 1) -225 1 0 9 1 8(1) 4240 0 38 1 8(2) 925 0 54 1 U(eq) is defined as one third of the trace of the orthogonalized Uij tensor. The M-M distances have been measured by EXAFS to be 2.74 A and 2.8 A for WS2 and MoS,, respectively.” Shorter M-M distances than those observed by EXAFS might be explained by displacement of the metal atom along the c-axis. As mentioned previously, this displacement is actually Observed in WTe,. Longer M-M distances than those observed by EXAFS may be an artifact of dynamic scattering: we have observed by simulation that the M-M distances appear to increase with increasing thickness (vide infra).33 Despite these difficulties, electron crystallography is a more powerful tool for the structure elucidation of these compounds than EXAFS. The latter probes only the local structure, while electron diffraction provides a direct structure determination. 159 4.3.7 Dynamic scattering: n-beam approximation. Although including the effects of dynamic scattering in the refinement should improve the results, it is non-essential for determination of the structure of these materials. This suggests that the use of electron diffraction for structure determination of inorganic materials may not be as unrealistic as previously supposed. Nonetheless, an effort was made to address the effects of dynamic scattering in these data sets. The two-beam approximation is severely limited by the assumption that each diffracted beam functions independently of all the other diffracted beams in the diffraction pattern. The assumption is particularly false when large accelerating voltages are used, resulting in very large Ewald spheres which can excite many reflections at one time. Therefore, alternate approximations, most notably the n-beam approximation, have been developed.”'3°'” The n-beam approximation permits the calculation of the dynamic contribution to each reflection from multiple beams in a series of slices through the crystal.”5 Although more accurate, it is significantly more complicated than the two-beam approximation, and solving a structure using a dynamical treatment of the electron diffraction data is a mathematically challenging procedure. Some researchers have combined HRTEM (High Resolution TEM) images with the electron diffraction data in order to obtain phase information and then used direct methods.37 160 Others have developed software which enables one to combine least-squares refinement with a multislice calculation.” As part of an effort to include a correction for dynamic scattering, the models which resulted from kinematic refinement in SHELX programs were constructed using the molecular modeling program CERIUSZ.33 This program contains a HRTEM module which employs the multislice n-beam calculation to simulate high resolution images. One step in this process involves the generation of electron diffraction patterns. Simulated diffraction data for varying thickness of the WS, crystal are shown in Figure 4.16. Figure 4.16a contains essentially a kinematic pattern, as it comes from a crystal which is only one layer thick. Note that the ( 200) is less intense than the (300). At three layers thick (Figure 4.16b) the sublattice reflections have increased in intensity, and at six layers thick (Figure 4.16c) the intensity of the (200) is equal to the (300). At eight layers thick (Figure 4.16d), the hexagonal sublattice reflections are the most intense reflections in the diffraction pattern. It is interesting to note that, using the two-beam approximation, one might have predicted that the sublattice reflections, which are generally more intense than the superlattice reflections, would have been decreased in intensity with increasing thickness rather than increased. One can qualitatively use the values of the (200) and the (300) to estimate the degree of dynamic 161 A i I B 1'07 3 . 1.01 i 053 O O O O . O O .0.5_ o O . T . O . O O O . O O O O O . . . O O O o . O O O . o 0.0 O O O 0.0 O O O 0 o O O O O . 0 O Q a O . . . 3 . . . . i . . . '0.57 O O . é . O O r05" O O . ‘ . O O -1.0— . ° , , ‘ . -1.0— 3 -110 05 0T0 05 {To -1i0 -015 050 015 110 1.0-C 3 1.0-D , 3 , o 0 § 0 o O O O . 05- ' ' 9 ' 05- ° ' 9 ' ° ' O O O O O ' O O ¢ O ‘O O o . O . O O O . O . o O 0.0 O O O 0.0 O O O O o . O . o O O o . O . o O as: O O O O O O O ‘ O O -0.5‘ . . O . e '05‘ . . O . . . o O O O o O O O -1.0- ' -1.0- b* -110 0.5 010 015 110 -110 -015 010 015 110 Figure 4.16. Electron Diffraction patterns of restacked W82 simulated using the multi-slice dynamic scattering calculation employed in the CERIUS2 HRTEM module: (A) one layer thick (kinematic scattering) (B) three layers thick (C) six layers thick and (D) eight layers thick. As thickness increases, the relative intensities of the hexagonal sublattice reflections increases. 162 scattering present in the sample. In this case, the data in Figure 4.2d is the most free of dynamic scattering. Simulations of the diffraction patterns from the models of MOS, were comparable, with the exception that the (200) surpassed the (300) before 5 layers thick rather than after 6 layers thick, as in the simulations involving WS,. One can also see that in both data sets for MOS, the (200) is more intense than the (300), suggesting contamination due to dynamic scattering. The intensities were extracted from these simulated diffraction patterns (an option in CERIUS”), and Patterson contour plots were generated (Figure 4.17). In the kinematic pattern (Figure 4.17a), the distortion in the W—W vector can be resolved as two distinct peaks, and the W-S vectors are elongated and fairly intense. At three layers thick (Figure 4.17b) the W-W vectors are slightly more blurred and the W-S vectors are more rounded. At six layers thick (Figure 4.17c) the W-W vectors are no longer resolvable as two separate peaks, and the W-S vectors are significantly reduced in intensity. At eight layers thick (Figure 4.17d) there is almost no intensity in the W-S vectors and the distortion in the W- W vector is only evident as a large, oval peak. If the data from this last Patterson map is refined against the model that was used to create it the W- W distance increases and the temperature factors of all the atoms become very large. Based on the fact that the simulated electron diffraction 163 patterns for MOS, are comparable, would expect it to follow a similar trend, but the deterioration of the data with thickness would be more rapid. This loss of resolution of the M-M vector with increasing dynamic scattering led us to conduct a kinematic simulation in which the W-W distance was varied from 2.65 A to 2.90 A (Figure 4.18). The W-W vector starts to blur between 2.70 and 2.80 A (Figures 4.18b and 4.18C), and blurs completely into one oval peak between 2.80 and 2.90 A (Figure 4.18d). The W—S vectors become more rounded due to the decrease in the distortion of the W atoms. The oval which describes the W-W vector in these Patterson contour plots is smaller than the oval in the plots generated as a function of thickness, but the most significant difference is that the W- S vectors are not decreased in intensity with increasing W-W distance. One might expect this feature to manifest itself as elevated temperature factors for the S atoms, which is observed to a small degree in the MOS, data. The CERIUS2 program is not designed to conduct least squares refinements in conjunction with multislice n-beam calculations. In order to try to evaluate the degree to which our data is contaminated by dynamic scattering, R, values were calculated as a method to compare the simulated dynamic data” with the experimental data: 211 6.1—K16. 11 —Ak 1” ' 20>. hk Eq.(4.10) 164 >b axis a axis Figure . c-axis calculated from the data in Figure 4.16: (A) one layer thick (kinematic scattering) (B) three layers thick (C) six layers thick and (D) eight layers thick. As thickness increases, the W-W vectors become blurred and the W-S vectors decrease in intensity. 165 b axis a axis Figure c-axis calculated from simulated electron diffraction data (kinematic) as a function of WW distance: (A) 2.65 A (B) 2.70 A (C) 2.80 A and (D) 2.90 A. As the W-W distance increases, the W-W vectors become blurred, but the W-S vectors retain their initial intensity. 166 Table 4.8. A comparison of simulated dynamic electron diffraction data to the experimental data for M08, and WS, as a function of layer thickness. Layer thickness R, for MOS, (%) R, for WS, (%) 1 15.38 13.27 2 14.75 13.21 3 15.44 14.77 4 19.03 19.58 5 23.15 24.98 6 26.19 29.92 7 28.14 34.76 8 30.13 39.81 Unfortunately the simulations did not result in significant improvements over kinematic least squares refinement: the best R, value calculated for the “single” W8, data set is only ~0.1% lower than the value from the kinematic refinement, and the best value for M08, is only ~0.9% lower. Part of the problem is that the atomic coordinates and the thermal parameters have already been minimized in a kinematic refinement in SHELX; if we had access to a program which would allow those parameters to refine against a dynamic model the R, value might 167 improve.38 The lack of improvement also suggests that there are other sources of perturbation of the data besides dynamic scattering. For example, the correction which was applied to compensate for the limited dynamic range of the electron microscope film may not have been adequate. Another possible source of perturbation Of the data is the flatbed scanner used to digitize the negative, as it is also somewhat limited in dynamic range. Despite these difficulties, meaningful structural information has been extracted from the electron diffraction patterns. 4.3.8 Electron Crystallographic Studies Of an alternate superlattice in M08, Recently, STM studies on the original lT-MOS,, prepared by the oxidation of K,(H,O),Mos2 with 1,, indicated that it has a 3a x fl. superstructure”, as proposed in the original publication reporting the material.11 It is becoming clear that restacked M08, is not the same as IT- MOS,. Other studies have indicated that restacked MoS,, previously believed to be neutral because of its ability to encapsulate neutral species, has some residual negative charge."0 The charge balancing species could be Li°, H“, or H,O". This residual negative charge apparently stabilizes the 168 structure of restacked MOS,. STM studies of K,(H,O) yMOS, suggest that it has the same superstructure as restacked MOS,.°l As will be discussed in the next chapter, it is possible to produce this alternate fl; x 1/3_a superlattice in restacked MOS, upon oxidation of the sample with Br,. It is very unusual to observe this lattice without the presence of the orthorhombic lattice solved in this chapter, but a few “single crystal” patterns were obtained (Figure 4.19). We attempted to elucidate the structure of this alternate lattice from the electron diffraction patterns. Figure 4.20a contains the Patterson map of an idealized octahedral species with a 1/3_a x l/3_a lattice. The M-M vectors can be found at (1/3, 2/3) and (2/3, 1/3). Figures 4.20b and 4.20c are the Patterson maps calculated from the experimental data in Figure 4.19. Although the M-M vector is not spherical, there does not appear to be a splitting or even an elongation of the M-M vectors as seen in the orthorhombic Patterson maps Actually it seems that, for this system, even if there is a distortion in the M-M distance it will not be visible in the Patterson map. In Figure 4.21a a Patterson map has been simulated from a model (created using the distortion model shown in Figure 4.1c) with a short M-M distance of 2.8 A. Although the M-M vectors are no longer spherical, there is no evidence of a pair of M-M vectors representing the short and long distances in the 169 structure, even though the map was generated from a model which contained such distances. In fact, if one distorts the M-M distance to a ridiculously short value, such as 2.2 A, the M-M vector disappears almost completely and the overall intensity of the peaks in the pattern is reduced (Figure 4.21b). The problem is that the Patterson map is not a plot of the electron density in the structure, but a plot Of the atom-atom vectors in the structure. The M-M distortion generates distortion in the M-8 vectors as well, which in this case complicates the Patterson map significantly more than in the orthorhombic cell and results in an overall increase in the background noise in the map. When the 8 atoms are eliminated from the model, the distorted M-M vectors are easily observed (Figure 4.21c). Although the Patterson did not provide much information about the structure, we attempted unsuccessfully to refine the structure using the proposed trimerization model shown in Figure 4.10, which belongs to the plane group p31m (#15). This unit cell is larger than the orthorhombic cell, and consequently contains more atoms. Even with maximum constraint Of the parameters the data:parameter ratio was 4:1, which is insufficient for a refinement. Data collection at higher accelerating voltages, or employing tilt studies to collect higher order reflections, are required for the determination of the structure of this compound from electron diffraction data. 170 A A frctoals of A .u .91. Seede electron draction pattern restacked M082 which have been oxidized with Br2, giving rise to a 43a x 43a superstructure. (A) ideal 1T-MS2 (computed from simulated data) (B) from the electron diffraction data shown in Figure 19a (C) from the data shown in Figure 19b. Although the M-M vector is not spherical, the map does not reveal a short M-M vector. 172 Figure 4. 21. Two- dimensional Patterson projections along the c- axis calculated from simulated data: (A) a trimerization with a short M- M distance of 2.8 A (B) a trimerization with a short M-M distance of 2.2 A and (C) the same model without 8 atoms. 173 4.4 Conclusions In summary, the structure of restacked WS, and M08, has been determined from electron diffraction data. The 2-D Patterson projection indicates unequivocally that the metal atoms are distorting to form zigzag chains with a short W-W distance of 2.74 A and a short Mo-Mo distance of 2.92 A. These distortions are in agreement with earlier theoretical predictions for layered octahedral ML, (12 systems.27 The results are consistent with those from EXAFS studies, and they are more informative because they provide direct structural information. Structure refinement has been conducted to R,=l3.3% for W8, and R,=15.3% for MOS,. We conclude that restacked M08, and WS, are not lT-TiS, type, but rather WTe, type, and that quasi-kinematical electron diffraction data are sufficient for meaningful structure elucidation. 174 References l. Fleischauer, P.D. Thin Solid Films, 1987, 154, 309. 2. Dickinson, R.G.; Pauling, L. J. Am. Chem. Soc., 1923,45, 1466. 3. a) Harris, 8.; Chianelli, R.R. J. Catalysis, 1984, 86, 400. b) Brenner, J.; Marshall, C.L.; Ellis, L.; Tomczyk, N.; Heising, J.; Kanatzidis, M.G. Chem. Mater, 1998, 5, 1244. 4. Julien, C.; Saikh, S.I.; Nazri, G.A. Mater. Sci. Eng., 1992, B15, 73. 5. Murphy, D.W.; DiSalvo, F.J.; Hull, G.W.,Jr.; Waszczak, J.V. Inorg. Chem, 1976, 15, 17. 6. a) Divigalpitiya, W.M.R.; Frindt, R.F.; Morrison, 8.R. Science, 1989, 246, 369. b) Divigalpitiya, W.M.R.; Frindt, R.F.; Morrison, S.R. J. Mater. Res., 1991, 6, 1103. c) Gee, M.A.; Frindt, R.F.; Morrison, 8.R. Mater. Res. Bull., 1986, 21, 543. 7. a) Bissessur, R.; Heising, J.; Hirpo, W.; Kanatzidis, M.G.; Chem. Mater., 1996, 8, 318. b) Wang, L.; Schindler, J.L.; Thomas, J.A.; Kannewurf, C.R.; Kanatzidis, M.G. Chem. Mater., 1995, 7, 1753. c) Kanatzidis, M.G.; Bissessur, R.; DeGroot, D.C.; Schindler, J.L.; Kannewurf, C.R. Chem. Mater., 1993, 5, 595. 8. a) Lemmon, J.P.; Lerner, M.M. Chem. Mater.,. 1994, 6, 207. b) Dungey, K.E.; Curtis, M.D.; Penner-Hahn, J.B. J. Catalysis, 1998, 175, 129. 9. Py, M.A.; Haering, R.R. Can. J. Phys, 1983, 61, 76. 10. Chianelli, R.R.; Scanlon, J.C.; Thompson, A.H. Mater. Res. Bull., 1975, 10, 1379. 11. Wypych, F.; SchOllhorn, R. J. Chem. Soc., Chem. Commun., 1992, 1386. 12. a) Wilson, J.A.; DiSalvo, F.J.; Mahajan, 8. Adv. Phys, 1975, 24, 117. b) Wilson, J.A.; DiSalvo, F.J.; Mahajan, 8.; Phys. Rev. Lett., 1974, 32, 882. 175 13. Yang, D.; Frindt, R.F. J. Phys. Chem. Solids, 1996, 57, 1113. 14. Yang, D.; Sandovals, S.J.; Divigalpitiya, W.M.R.; Irwin, J.C.; Frindt, R.F. Phys. Rev. B. , 1991, 43, 12053. 15. Tsai, H.L.; Heising, J.; Schindler, J.L.; Kannewurf, C.R.; Kanatzidis, M.G. Chem. Mater., 1997, 9, 879. 16. Chrissafis, K.; Zamani, M.; Kambas, K.; Stoemenos, J.; Economou, N.A.; Samaras, I.; Julien, C. Mater. Sci. Eng., 1989, B3, 145. 17. Qin, X.R.; Yang, D.; Frindt, R.F.; Irwin, J.C. Ultramicroscopy, 1992, 42-44, 630. 18. Wypych, F. Weber, Th. Prins, R. Chem. Mater., 1998, 10, 723. 19. Vainshtein, B.K. Structure Analysis by Electron Difiraction; Pergamon Press, MacMillan Co., New York, 1964. 20. Dorset, D.L. Structural Electron Crystallography ; Plenum Press, New York, 1995. 21. Stout, G.H.; Jensen, L.H. X-ray Structure Determination: a Practical Guide (2nd ed.); John Wiley & Sons, New York, 1989. 22. Ladd, M.F.C.; Palmer, R.A. Structure Determination by X-ray Crystallography (3rd ed.); Plenum Press, New York, 1994. 23. Analysis of integrated intensities was performed on a Macintosh Performa 6214CD computer using the public domain NIH Image 1.60 program (developed at the US. National Institutes of Health and available on the Internet at http://7/2/98/rsb.info.nih.gov/nih~image/) using the “Gel Plotting Macros” plug-in. 24. Sheldrick, GM. “SHEL'I'XTM Version 5”, Bruker Analytical X-Ray Instruments, Inc., Madison, WI. 25. Hyarns, Daniel. “CurveExpert Version 1.34”, 1995-1997. Portions copyright (1993) Microsoft Corporation, Seattle, WA. 26. a) Brown, B.E. Acta Crystallogn, 1966, 20, 268. b) Mar, A.; Jobic, 8.; Ibers, J. A. J. Am. Chem. Soc., 1992, 114, 9587. 176 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. a) Rovira, C.; Whangbo, M.-H. Inorganic Chem, 1993, 32, 4094. b) Whangbo, M.-H.; Canadell, E. J. Am. Chem. Soc., 1992, 114, 9587. Bissessur, R.; Kanatzidis, M.G.; Schindler, J.L.; Kannewurf, C.R. J. Chem. Soc., Chem. Commun., 1993, 1582. a) Pinsker, Z.G.; Dvoryankina, G.G. Kristallografiya, 1958, 3, 438. b) Nagakura, 8. Acta Crystallogr., 1957, 10, 601. Q = A |/ Q I, where (I) is the structure factor, Q is the volume of the unit cell, and A is the wavelength of the electrons. Klug, P.; Alexander, L.E. X-Ray Diffraction Procedures for Polycrystalline and Amorphous Materials, John Wiley and Sons, New York, 1954. a) Joensen, P.; Crozier, E.D.; Alberding, N.; Frindt, R.F. J. Phys. C.: Solid State Phys, 1987, 20, 4043. b) Prouzet, E.; Heising, J.; Kanatzidis, M.G. Unpublished results. “CERIUS2 Version 3.5 HRTEM module”, Molecular Simulations, Inc., San Diego, CA. Cowley, J.M.; Moodie, A.F. Acta Cryst., 1957, 10, 609. Cowley, J .M. Difi’raction Physics; Elsevier, New York, 1975. Spence, J.C. H. Acta Cryst., 1998, A54, 7. Sinkler, W.; Marks, L.D.; Edwards, D.D.; Mason, T.O.; Poeppelmeier, K.R.; Hu, 2.; Jorgensen, J .D. J. Solid State Chem, 1998, 136, 145. a) Zandbergen, H.W.; Andersen, S.J.; Jansen, J. Science, 1997, 277, 1221. b) Jansen, J.; Tang, D.; Zandbergen, H.W.; Schenk, H. Acta Cryst., 1998, A54, 91. “CERIUS2 Version 3.8 HRTEM module”, Molecular Simulations, Inc., San Diego, CA. (According to the company, this newer software version contains an “improved” dynamic scattering calculation, 177 particularly with regard to the incorporation of the temperature factors of the atoms in the calculation.) 40. Heising, J.; Bonhomme, F.; Kanatzidis, M.G. J. Solid State. Chem, 1998, 139, 22. 41. Wypych, F.; Weber, Th.; Prins, R. Surf. Sci., 1997, 380, L474. 178 CHAPTER 5 Exfoliated and Restacked M08, and WS,: Ionic or Neutral Species? Encapsulation and Ordering of Hard Electropositive Cations 179 5.0 Abstract The relationship between charge and structure in restacked M8, (M = MO,W) has been probed by cation encapsulation and chemical oxidation and characterized by elemental analysis, electron diffraction, X-ray diffraction, and Differential Scanning Calorimetry. Alkali cations have been encapsulated in M08, and WS, without the presence of a co-intercalated counter-ion, suggesting a negative charge in the range 0.15-0.25 electrons per M atom. Electron diffraction studies show ordering of these cations between the layers. Chemical oxidation with I, or Br, results in a change in the structure of restacked MoS,, giving rise to a l/3_a x /3—a superlattice, whereas no change is observed in the structure of restacked WS,. Differential Scanning Calorimetry studies show an irreversible exothermic transition to 2H-MS, upon heating, which shifts in temperature with oxidation. Thermopower measurements indicate that restacked M08, and WS, are p-type metallic conductors, consistent with an observed structural distortion and residual negative charge. 180 5.1 Introduction 2H-MOS, and 2H—W8, are the most stable members of the family of layered dichalcogenides, and find practical application in such processes as hydrodesulfurization (HDS) due to their availability and relatively low cost. ' Both materials strongly resist attempts to incorporate guest species between their layers.2 Li atoms can be inserted into M08, (and WS,) only upon treatment with a strong reducing agent such as n-butyllithium (n- BuLi) or LiBH,.3'° LiMoS, and LiWS,, however, exhibit a remarkable ability to exfoliate in water by the following redox reaction: LiMS, + H,O --—> (MS,),,,,,, m + LiOH + H, (g) (Eq. 5.1) resulting in a colloidally dispersed suspension of single layers.”5 The layers can be recovered in a restacked form by filtration, centrifugation, or precipitation. The ability of these materials to exfoliate and to be restacked with relative ease has permitted the encapsulation of a wide variety of guest species which include neutral organic molecules,6 polymers,7 metal chalcogenide8 and metal oxide9 clusters, metallocenes”, porphyrins”, and metal cations?”13 The authors who first reported the exfoliation Of MOS, believed that the redox reaction between H,O and M08, resulted in complete oxidation of the layers, returning the molybdenum to a 4+ oxidation state, because many 181 of the molecules which have been encapsulated are neutral.‘5°8'14 A net drift of the suspended layers toward the anode in an electrophoresis apparatus was observed by these authors, but was attributed to OH' ions associated with the layers which could be displaced by the organic molecules.15 Subsequent pH studies seemed to support this hypothesis.“5 The increasing number of examples of cationic species which have been encapsulated have suggested to other authors that the layers may retain some negative charge.“"°“3'l7 It is rather curious that restacked M082 (and W82) can incorporate both cationic and neutral species. The ambiguity of the situation stems from an abundance of both Li* and OH‘ generated in the exfoliation process: Li‘, which could counterbalance a negative charge on the layers when neutral species are intercalated, is a notoriously difficult element to detect; OH’, which could co—intercalate with the cationic species, can be confused with residual co-intercalated H20. In addition to the unresolved issues about the oxidation state of the layers, there has also been confusion about the atomic structure of the layers. Reduction of 2H-MoS2 with n-BuLi or LiBH4 results in a structural transformation in the layers, causing the geometry around the metal atoms to shift from trigonal prismatic to octahedral.18 This results in a change in the conductive properties of the material from semiconducting to 18.19 metallic. This structural change appears to be retained in exfoliated 182 MoSz, with a structural distortion which results in a a x l/3_a orthorhombic superlattice?“21 Restacked MoS2 has also been called 1T— MoS2 because of the octahedral coordination of the Mo atom. Restacked WS2 appears to be analogous to restacked MoSz.“'21 Another material which has been called 1T--MoS2 is prepared by high temperature synthesis and subsequent oxidation of K(,_33(H20)yMoSZ.22 Restacked M082 and lT-MoS2 were thought to be the same material because both appear to have octahedral metal coordination and both undergo an irreversible exothermic transition to 2H-MoS2 at around 100°C.”23 lT-MoSZ, however, exhibits a 1/51 x 1/371 superlattice.”24 The a x 1&1 superlattice found in restacked M082 is due to M-M associations resulting in the formation of zigzag chains, whereas a 1/321 x [[321 superlattice would involve M-M trimerization. Are these layers neutral or cationic? Why do 1T-MoS2 and restacked MoS2 have different superlattices? What is the impact of negative charge on the structure of these materials? If the M082 layers are slightly reduced, how does it affect the conductivity? In this chapter we describe our experimental efforts toward answering these questions. The encapsulation of hard, electropositive alkali metal cations is described, chosen due to their relatively poor affinity for OH,’ which make them perhaps the best 183 species for chemical analyses to determine negative charge. The materials are characterized by X-ray diffraction, electron diffraction, TGA, and elemental analyses. The treatment of LiMS2 and exfoliated MS2 with the oxidizing agent Br2 and concentrated HCl has also been conducted and the products characterized by X-ray diffraction, electron diffraction, and DSC measurements. Thermopower measurements of restacked M082 and W82, which address the effects of structure and charge on the nature of their conductivity, are also presented. 5.2 Experimental 5.2.1 Synthesis LiMoS2 and LiWS2 were synthesized by the LiBH4 method.4 Acetonitrile, RbCl, CsF, and CsI were purchased from Aldrich. KCl and were purchased from J .T. Baker. Br2 and NaCl were purchased from EM Science. I2 was purchased from Mallinckrodt. HCl was purchased from Columbus Chemical Industries. All compounds were used as received. MoS2 aqueous suspension. In a glove box under nitrogen atmosphere LiMoS2 (0.1 g, 0.6 mmol) was placed in a 125 ml Erlenmeyer flask equipped with a stir bar and rubber stopper. The flask was removed from 184 the glove box and 10 ml deoxygenated deionized H20 was added. The mixture was allowed to stir for 0.5 hr, then the contents of the flask were transferred to a centrifuge tube and centrifuged for 0.5 hr. The supernatant was decanted (pH >12) and to the black gooey solid at the bottom of the tube 10 ml fresh deoxygenated deionized H20 was added to rinse away the LiOH. The tube was agitated to re-suspend the product, and the mixture again centrifuged for 0.5 hr. This rinsing was conducted three times, the supernatant having pH ~12, pH ~9, and pH ~7 after each rinse. Then 10 ml more deoxygenated deionized H20 was added, the tube was agitated to re-suspend the product, and the contents of the centrifuge tube were returned to a 125 ml Erlenmeyer flask with stir bar. The suspension was stirred for at least 0.5 hr before further use. WS2 aqueous suspension. In a glove box under nitrogen atmosphere LiWS2 (0.1 g, 0.39 mmol) was placed in a 125 ml Erlenmeyer flask equipped with stir bar and rubber stopper. The flask was removed from the glove box and 10 ml deoxygenated deionized H20 was added. The material was then centrifuged, rinsed, and re-suspended in the same manner as MoSz. Na0.,4(H,O)mMoSZ. NaCl (0.11 g, 1.8 mmol) was dissolved in about 3.5 ml H20. To this solution was added an aqueous suspension of MoS2 (3:1 excess Na”). Flocculation occurred immediately. The reaction mixture was stirred for one half hour, centrifuged and the supernatant decanted. 185 The product was rinsed and centrifuged three times as described in the preparation of aqueous MoS,, then deposited on a glass slide to dry. K013(H20)MMoSZ. KCl (0.13 g, 1.8 mmol) was dissolved in about 3.5 ml H20. To this solution was added an aqueous suspension of MoS2 (3:1 excess K*). Flocculation occurred immediately. The product isolation was analogous to that used for Na0.,4(H20)0.45MoSZ. Rbo.,5(H20)oJMoSr RbCl (0.22 g, 1.8 mmol) was dissolved in about 3.5 ml H20. To this solution was added an aqueous suspension of MoS2 (3:1 excess Rb*). Flocculation occurred immediately. The product isolation was analogous to that used for Nam 4(I-120)0.45M082. Cso.,3(HzO)yMoS,. CsF (0.27g, 1.8 mmol) was dissolved in about 2 ml H20. To this solution was added an aqueous suspension of MoS2 (3:1 excess Rb“). Flocculation occurred immediately. The product isolation was analogous to that used for Na0.,4(H20)0_45MoSZ. Ba,(H,0),Mos,. BaC12'2H20(0.88g, 3.6 mmol) was dissolved in about 5 ml H20. To this solution was added an aqueous suspension of MoS2 (6:1 excess Ba2+ ). Flocculation occurred immediately. The product isolation was analogous to that used for Na0.,4(H20)0_45MoSZ. Na0.,,(H20)M,WS,. NaCl (0.14 g, 2.4 mmol) was dissolved in about 3.5 ml H20. To this solution was added an aqueous suspension of WS,, (6:1 excess Na*). Flocculation occurred immediately. The product isolation was analogous to that used for Na0.,4(H20)0.45M082. 186 K03,(H20)M,W82. KCl (0.18 g, 2.4 mmol) was dissolved in about 5 ml H20. To this solution was added an aqueous suspension of WS2 (6:1 excess K”). Flocculation occurred immediately. The product isolation was analogous to that used for Nao.,4(H20)0.45MoS2. Rbo.24(H,O)oMWSZ. RbCl (0.15 g, 1.2 mmol) was dissolved in about 3.5 ml H20. To this solution was added an aqueous suspension of WS2 (3:1 excess Rb“). Flocculation occurred immediately. The product isolation was analogous to that used for Na0.,4(H20)0.45MoSZ. Cso.,3(H20)yWS,. Method 1. An aqueous suspension of WS2 was added to 5 ml of a concentrated CsF solution in H20. To this solution was added an aqueous suspension of WS2 (7:1 excess Cs*). Flocculation occurred immediately. The product isolation was analogous to that used for Nao,14(HzO)o,45MoSZ. Method 2. C51 (0.73 g, 2.8 mmol) was dissolved in about 5 m1 H20. To this solution was added an aqueous suspension of WS2 (7:1 excess Cs“). Flocculation occurred immediately. The product isolation was analogous to that used for Na0_,,(H20)0_45MoSZ. Baom(H20),WS,. BaC12°2H20 (0.29 g, 1.2 mmol) was dissolved in about 5 ml H20. To this solution was added an aqueous suspension of WS2 (3:1 excess Ba2+ ). Flocculation occurred immediately. The product isolation was analogous to that used for Na0.,4(H20)0.45MoS2. 187 HxMoSZ. In a glove box under nitrogen atmosphere LiMoS2 (0.3 g, 1.8 mmol) was placed in a 125 ml Erlenmeyer flask equipped with a stir bar and rubber stopper. The flask was removed from the glove box and 30 ml cold concentrated HCl was added. The reaction mixture was stirred for a few minutes, then was transferred to a centrifuge tube and centrifuged for a few minutes. The supernatant was decanted, then 30 ml cold H20 was added to the solid to rinse, and the slurry was centrifuged again. This rinsing was repeated two more times, the supernatant after each rinse having pH~1, pH~1, and pH~7, respectively. The solid was then slurried in a few ml of H20 and deposited on a glass slide to dry overnight. H,WS,. In a glove box under nitrogen atmosphere LiWS2 (0.3 g, 1.2 mmol) was placed in a 125 ml Erlenmeyer flask equipped with a stir bar and rubber stopper. The reaction procedure and rinsing was the same as the one used to prepare HxMoSz, with the supernatant after each rinse having pH~1, pH~4, and pH~7, respectively. The solid was then slurried in a few ml of H20 and deposited on a glass slide to dry overnight. Oxidized LiMoS,. Method 1. In a glove box under nitrogen atmosphere LiMoS2 (0.3 g, 1.8 mmol) was placed in a 125 ml Erlenmeyer flask equipped with a stir bar and rubber stopper. The flask was removed from the box, and 50 ml of a 0.18 M solution of I2 in acetonitrile or 50 ml of a 45 mM solution of Br2 in acetonitrile was added to the solid (5:1 excess Ilerz). The mixture was stirred for different amounts of time 188 ranging from 15 minutes to 4 days, then was centrifuged and the supernatant decanted. The product was rinsed by re-suspending it in 15 m1 acetonitrile and re-centrifuging 2 times and then deposited on a glass slide and a saran wrapped transmission XRD slide to dry. Method 2. In a glove box under nitrogen atmosphere LiMoS, (0.15 g, 0.9 mmol) was placed in a 125 ml Erlenmeyer flask equipped with a stir bar and rubber stopper. The flask was removed from the box and an exfoliated suspension of MoS, was prepared as described previously, then centrifuged and rinsed once with acetonitrile. To the solid 50 ml of a 0.18 M MeCN solution of I, or 50 ml of a corresponding 45 mM MeCN solution of Br2 was added (5:1 excess I,/Br,). The mixture stirred for different amounts of time ranging from 15 minutes to a week, was centrifuged and the supernatant decanted. The product was rinsed by re-suspending in acetonitrile, re-centrifuging and decanting the supernatant. The product was then deposited on a glass slide and a SaranTM wrapped transmission X- ray Diffraction (XRD) slide to dry. Oxidized LiWS,. Method 1. In a glove box under nitrogen atmosphere LiWS, (0.3 g, 1.2 mmol) was placed in a 125 ml Erlenmeyer flask equipped with a stir bar and rubber stopper. The flask was removed from the box, and to the solid 50 ml of a 0.12 M MeCN solution of I, or corresponding 75 ml 80 mM MeCN solution of Br, was added (10:1 excess I,/Br,). The mixture stirred for varying amounts of time from ~15 189 minutes to a week, was centrifuged and the supernatant decanted. The product was isolated as described for oxidized LiMoS,. Method 2. In a glove box under nitrogen atmosphere LiWS, (0.15 g, 0.6 mmol) was placed in a 125 ml Erlenmeyer flask equipped with a stir bar and rubber stopper. The flask was removed from the box and an exfoliated suspension of WS, was prepared as described previously, then centrifuged and rinsed with acetonitrile. To the solid 50 ml of a 0.12 M MeCN solution of I, or corresponding 75 ml 80 mM MeCN solution of Br, was added (10:1 excess I,/Br,). The reaction time and product isolation was the same as described above. Method 3. LiWS, (0.15g, 0.6mmol) was used to prepare an exfoliated suspension as described previously. Next 100ml of a saturated aqueous solution of Br, was added. The product was centrifuged immediately for a few minutes, then rinsed with water and centrifuged three times. It was deposited on a glass slide and a SaranTM wrapped transmission XRD slide to dry. 190 5.2.2.Characterization. Selected Area Electron Diffraction patterns were collected using a JEOL 120 CX Transmission Electron Microscope (TEM). Transmission and reflection powder X-ray diffraction patterns were collected using a Rigaku-Denki/RW400F2 (Rotaflex) rotating anode X-ray diffractometer using Cu-Ka radiation. Differential Scanning Calorimetry (DSC) measurements were conducted using a Shimadzu DSC-50 instrument under nitrogen flow in sealed aluminum containers at a heating rate of 5°C/minute. Thermogravimetric Analyses (TGA) employed a Shimadzu TGA-50 instrument under nitrogen flow in quartz containers at a heating rate of 2°C/minute. Energy Dispersive Spectroscopy (EDS) was used to determine the alkali metal to Mo/W ratio, and employed a JEOL-ISM- 6400V Scanning Electron Microscope was operated at an accelerating voltage of 30kV using a Tracor Northern 5500 X-ray microanalysis attachment. Thermopower measurements utilized a MMR Technologies Seebeck System. 191 5.3 Results and Discussion 5.3.1 Cation encapsulated MS,. Addition of exfoliated MOS,./W S, to the alkali halide solution results in immediate flocculation of the suspension. This rapid precipitation, which is not usually observed upon addition of neutral species, strongly suggests that the interaction of the layers with the alkali cations is ionic in nature. Similar behavior is observed upon addition of exfoliated MoS,/W S, to solutions containing Al,,0,,(OH),,,(H,O)1,7+ or tetraphenylphosphonium cations.”25 The samples were characterized by reflectance X-ray diffraction (Figure 5.1a) and were found to be well ordered, particularly compared to the X-ray diffraction patterns of samples containing neutral species such as the cobalt chalcogenide clusters.8 The predominance of the 001 reflections in the pattern is due to the tendency of lamellar materials to adopt a preferred orientation. All samples exhibited interlamellar d-spacings in the range 9.3-9.7 A (an expansion of 3.1 - 3.5 A), indicating co-encapsulation of approximately one monolayer of water with the cations. 192 Intensity (arbitrary units) weight loss (%) 96.5E‘J—‘l‘411111111111r111 50 100 150 200 250 Temperature (°C) Figure 5.1. (A) X-ray diffraction pattern of Nao.,,(H,O)o.45WS,., exhibiting a d-spacing of 9.7 A (A: 3.5 A). (B) TGA of Na0_,,(I-I,O)o.45WS, (dotted line), Ko.21(1'120)o.4owsz (daShed line)’ and Rbo.24(HzO)o.34WS2 (SOIid line). Weight loss corresponds to H,O loss. 193 The flocculated products were found to have the following formulas: Nao,|4(1'120)xMOS,, K0,,(H,O),MoS,, Rbo.,5(H,O),MoS,, Cso_,3(H,O),MoS,, Nao.13(HzO),WS,, Ko_,,(H,O),WS,, Rb0_,4(H,O),WS,, Cso_,3(H,O),WS,, and Ba0.08(H,O),WS,. . TGA (Thermal Gravimetric Analysis), performed in order to determine the amount of co-encapsulated water, indicates a range of 03-045 for x (Figure 5.1b). Higher x values were found for samples containing Na+ and K*; lower values for those containing Cs” and Rb“. No halide counter-ions were detected. Although the presence OH- ions cannot be completely discounted, the pH of the suspension (~7) and the relatively poor affinity of the alkali cations for OH- ions make their co-encapsulation unlikely. 5.3.2 Electron diffraction studies of (cation),MS,. The cation intercalated M08, and WS, samples were also examined by electron diffraction. Both restacked M08, and WS, give rise to electron diffraction patterns which appear to be more simple than they actually are.21 In Chapter 4, we solved and refined the structure of restacked M08, and WS, from electron diffraction data.“ The materials exhibit a superlattice caused by metal-metal associations to form zig-zag chains, similar to those observed in W'I'e,.26 This zig-zag distortion results in the 194 doubling of one axis of the hexagonal sublattice. As the hexagonal symmetry is destroyed, the new lattice is best described by an I/3_a x a orthorhombic cell (Figure 4.2d, Figure 4.3). Due to their turbostratic nature and the disorder introduced by the exfoliation/flocculation process, the diffraction patterns of exfoliated and restacked MoS, and WS, are often a “triplet” of these orthorhombic cells, forming an apparent 2a x 2a superlattice (Figure 4.2b). Failure to recognize this twinning can lead to incorrect interpretations of the M-M interactions. The cation intercalated samples are represented by an array of diffraction patterns (Figure 5.2) which contain an additional superlattice due to ordering of the electropositive cations in the gallery. The high vacuum found in a TEM (~10'7 torr) probably results in the removal of most of the co-encapsulated water from the samples. The majority of the patterns appear at first glance to contain a weak 6a x 6a superlattice relative to the hexagonal sublattice, with distinct variations in the reflections present and the intensity distribution from one pattern to the next. Because of the propensity of these materials for twinning, however, the diffraction patterns are actually quite complicated. 195 u ‘n #14, I . ‘ mi ' -’ ' '4. Figure 5.2. SAED patterns of (A) Cso,23M082 and (B) Cso,13W82. Super- lattice reflections caused by Cs+ ordering indicate a tripling of the a axis and the 43a axis, indicating a 3430 x 30 supercell. Twinning of the lattice projects a 6a x 6a unit cell. 196 Considering only the orthorhombic lattice caused by the formation of zig-zag M-M chains in the host material,21 Figures 5.2a and 5.2b contain one dominant flh x a orthorhombic cell with only weak twinning. In the interest of clarity, this /3_a x a orthorhombic lattice will henceforth be called the “host-lattice”. A closer examination of Figure 5.2a reveals that the superlattice caused by the intercalated cations form strong columns of reflections along the 1/3_a* axis of the host-lattice. The superlattice reflections are spaced in thirds between the host-lattice reflections. This suggests that at least one cell parameter should be 3a. The reflections at 1/2 the spacing between the columns of host-lattice reflections can be attributed to the weak twinning of the host-lattice; thus, the same phenomena which give rise to a false 2a x 2a lattice also give rise to a false 6a superlattice parameter. Interestingly, there do not appear to be any superlattice reflections from the cations contained in the columns or rows of host-lattice reflections in Figure 5.2a. Considering one of the columns of superlattice reflections to find the second cell parameter, the smallest distance between two reflections is 1/3 the length of the (010) reflection of the host-lattice; therefore, the superlattice may be minimally described by the orthorhombic cell 3a’ x 3b relative to the host-lattice (a’ = 3a; b = 197 3a), and 31/32: x 30 (orthorhombic) relative to the hexagonal sublattice of the host material (Figure 5.2b). It is logical to conclude that, if the host-lattice is somewhat twinned, the superlattice will also be somewhat twinned. In this manner the variations in the other diffraction patterns (Figures 5.3 and 5.4) can be explained by the degree of twinning of the 1/51 x a orthorhombic host- lattice and the 3a' x 3b cation superlattice. Figure 5.3a is very similar to Figures 5.2a and 5.2b, but comes from K0,,MoS, rather than Cs0,3MoS, (Figure 5.2a) or CstS, (Figure 5.2b). The host-lattice is more twinned than the pattern in Figure 5.2a, but the pattern of cation ordering is essentially the same. Figures 5.3b and 5.3c are from crystals of K0,,3MoS, and Rb0_,4MoS,, respectively, that are twinned but are still relatively thin. The distribution of orthorhombic host-lattice cells is uneven, appearing to be stronger in one direction. Consequently, the cation superlattice in these pictures is still stronger in one direction, but does show evidence of twinning. Unlike the previous patterns, the cation superlattice reflections are now found in the columns of reflections along the l/3_a * axis of the dominant orthorhombic host-lattice (arrows, Figure 198 3'.’ v .- Figure 5.3. SAED patterns of (A) K023MoSZ, which resembles the patterns in Figure 5.2 but the degree of twinning is slightly greater; (B) K033MoSZ and (C) Rbo. tsMoSZ, in which the host-lattice is more twinned but the crystals are still relatively thin; and (D) Ko_23Mosz, from a thick crystal with a very twinned host-lattice and a "halo" of cation superlattice reflections. As twinning of the host-lattice increases, twinning of the cation superlattice increases, leading to a variety of diffraction patterns. 199 . a: ‘ Figure 5.4. SAED patterns from (A) a thick crystal of LixWS2 and (B) a thin crystal of RbonMoSz. Streaking is observed due to disorder of the cations. The cations are more susceptible to disorder along the ‘13a axis. 200 5.3b). This seems to be due to cation superlattices associated with the other (weaker) host-lattice cells. Figure 5.3d, on the other hand, came from a thicker crystal of K0_,3MoS,. In this case, the twinning of the host-lattice is so pronounced that that the host material exhibits the pseudo 2a x 2a hexagonal lattice which is common in thicker crystals. The crystal is thin enough that the superlattice from the intercalated cations is still visible, but extremely weak. It is twinned in the same manner as the host-lattice, creating a hexagonal “halo” around the pseudo 2a x 2a lattice. The systematic absences of the superlattice (for example, the 020, 130, and the 310 ) are completely destroyed due to the twinning, generating the pseudo 60 x 60 cation superlattice. Not all crystals gave rise to well ordered cation superlattices. Figure 5.4a comes from a rather thick sample of exfoliated WS, that contained a large excess of LiOH in solution when it was flocculated; there is evidence of Li” cations ordering in the sample, but the pattern is very weak and, instead of spots, one sees faint streaking. The streaking is likely due to the fact that Li, in addition to being a very light atom, is probably coordinated by four water molecules, which would prevent it from resting easily in the sites created by the S atoms. The pattern best resembles Figure 5.3d. Crystals which contain discrete superlattice spots are somewhat unusual; 201 more common are patterns which contain disorder or streaking along the l/3_a* axis of the host lattice, as in the waMoS, crystal in Figure 5.4b. 5.3.3 Modeling the cation ordering. It is virtually impossible to obtain a diffraction pattern which is completely “single crystal” with respect to the host-lattice and the cation superlattice so that this manner of indexing the patterns might be confirmed; however, it is the best available indexing which can account for the variations in intensities and reflections from one diffraction pattern to the next, all of which can be found within one sample. As such, we have developed a cation ordering model from which an electron diffraction pattern was simulated. It has been designed to resemble the diffraction patterns in Figures 5.2a, 5.2b, and 5.3a, since they are the closest to a “single crystal” pattern (Figure 5). As noted previously, no superlattice reflection is found within a column or row of host-lattice reflections (Figure 5.2a). The model of cation ordering, depicted in Figure 5.5a, gives rise to such an electron diffraction pattern (Figure 5.5b).27 202 A 1 . . . é . 0 O E o o o o O i... O o c _ . g . I: '5 m 0 o . o . o 5.0 . o . o 0 an, . . 5 . g o o o O o 0.0; Q o .0 O o _ O : o 8 m--. ------ Q ------ -o ----- .--no-o-a-u-«En---o-o-o----.----o--o «- .— o O : 0 e o O o. o Q g... o O 0.0 o a“ o E O .5 o o . o o O 0.; o O .0 . o O o 3. I-t — , , V 0 0 DE 0 O o o 9(- o E '6 d 0 Q o . o o E o l l l I i l l l I d* (reciprocal angstroms) Figure 5.5. (A) Structural model and (B) simulated electron diffraction pattern to explain the observed SAED patterns in Figures 5.2-5.4. Filled circles represent cations in one layer; open circles represent cations in a second layer. The projection of the two results in a 3V3a x 30 supercell. 203 The cell contains three chains. Alkali cations are placed along the chains in pockets defined by sulfur atoms, where a cation would be expected to reside in the gallery. One third of the sites are empty in two of the chains; two thirds of the sites are empty in the third chain. So many alkali cations would not normally be in such close proximity to one another within one layer, however, particularly when one considers the elemental analyses of these samples. More likely this picture represents the superposition of two alternating layers of cations, depicted as white and black. Hence, four of the cations have an occupancy of 0.5 and one cation is fully occupied. This results in a stoichiometry of A0.,67MS,, which is consistent with the experimentally determined formula. Note that it is a necessary condition of the systematic absences that all the alkali cations be in the same kind of site, either “down” or “up” relative to the zig-zag chain. The alkali cations occupy only sites which are pointing “up” in the model presented in Figure 5.5a; however, the sites which are pointing “down“ are equally valid sites, so it becomes apparent that this is a possible source of disorder in the samples. The streaking which is observed along the 1/3-a * axis in Figure 5.4b indicates that the cations are more disordered along the 1/33 axis of the host-lattice. This is reasonable when one considers that the M atom distorts from its ideal octahedral position along this axis, causing an accompanying shift of the S atoms and distorting the site on which the 204 alkali cation sits. Also, the fact that there seems to be no disorder along the 0 axis indicates that the cations are better ordered along the zig-zag chains than between zig-zag chains. A few of the electron diffraction patterns did not have this 3a x 3b superlattice, but had a rather different pattern (Figure 5.6) corresponding to a 1/3-a x 1/3—a lattice combined with the pseudo 2a x 2a lattice of the host material, forming a “honeycomb“ pattern. The projection of this 1/33 x 1/3_a unit cell occupies a smaller area than the previous superlattice (31 A2 vs. 161 A2), which is puzzling when one considers that it is most readily observed for samples in which Ba“ is encapsulated. To reconcile such a lattice with the observed loading of cation one must again employ alternating layers of cations and consider their projection onto the ab plane. For example, in Figure 5.7a a model has been constructed with four layers of cations, depicted as white, black, gray, and striped circles. The unit cell per MS, layer is 2 3a x 2 3a (solid line), but the projection is the smaller 3a x [[321 lattice (dashed line). This gives a stoichiometry of A0.09MS,. The simulated diffraction pattern generated from this cation model (Figure 5.7b) resembles the experimental data.27 It is now possible to understand why cations of different charge can have the same diffraction pattern - they can occupy half as many sites and require twice as many 205 layers to project the appropriate unit cell. This is consistent with the fact that the superlattice is weaker in the samples containing Ba2+ than it is in the samples containing N a+ or Rb“. 206 Figure 5.6. SAED patterns of (A) BaoggMOSz and (B) Nao_14MoSZ. The combination of a V3a x ‘13a cation superlattice with a pseudo 2a x 2a host- lattice gives rise to a "honeycomb" pattern. This alternate cation super- lattice, although most common for samples with encapsulated Ba2+, was also observed within samples which exhibited the 3V3a x 3a superlattice. 207 A l C B o E o o . o 3 o . o A '7 5 U2 0 ; o o a o o o : o o O O _ o + o I: o 0 Q o g Q . Q o o {I} i 3.. 0 O 9 O o a o 0 g 0 o a -—----o ------ .. ------ —O ------ .- ------ ...-----.§ ------- O ------ . ------ -o ------ .— ------ on- 8 O O i O o a— o . o . o $ 0 . o . o o_ i Q I o . o . E . o . o 0 cu : b — o O ? O o o o . o o 0 cu. o o E o o c d O 0 Q ‘ O 0 O .1 O . O : O . O O l I o l l I I i l l l l d* (reciprocal angstroms) Figure 5.7. (A) Structural model and (B) simulated ED pattern to explain the SAED patterns in Figure 5.6. Black circles represent one cation layer; open circles a second; grey a third; and striped a fourth. Each layer has a 2V3a x 243a lattice, but the ‘13a x V3a lattice is formed by the projection. 208 5.3.4. Oxidation state of M (M = Mo,W) and the charge Of the layers. The fact that one can see ordering of alkali cations in the MOS, and WS, layers without co-intercalation of the halide counter-ions supports the conclusion that the exfoliated and restacked layers retain residual negative charge. The layers in suspension behave as if they are solvated macroanions, precipitating readily in the presence of most cationic species. The hypothesis that the charge is due to the association of OH' ions with the basal planes seems unlikely because the driving force for the association of the OH' ions with the layers is not obvious. Furthermore, the pH Of the exfoliated suspension has been reduced to ~7 by repeated rinsing. It seems surprising that any remaining OH' ions would not be at least partially displaced in a solution with an excess of halide ions, resulting in materials with co-encapsulated halide ions. Unfortunately the. high degree of disorder and the propensity of the systems for twinning make it impossible to determine the degree of negative charge on the layers based on the cell parameters of the samples with encapsulated cations. The amount of negative charge on exfoliated and restacked MOS, has been studied rather extensively by a group of Russian scientists.”l7 They assert that MOS, which has been freshly exfoliated under an inert atmosphere has a negative charge of approximately 015-029 per Mo atom 209 (Li+ counter-ion), but if the dispersion is allowed to sediment over time the value decreases to 0.07-0.12 per Mo atom.17m This slow, continuous oxidation of MOS, suspensions in solution over time may explain how it is possible to intercalate the neutral cobalt chalcogenide clusters, which require stirring for at least 2 days before encapsulation is observed.8 This same group has also observed that in some transition metal systems it is possible to co—intercalate OH’ anions, thereby increasing the number of cations which may be encapsulated between the layers beyond the range of negative charge attributed to MOS,.'3° The ability of MOS, and WS, to accept co-intercalated OH' ions hinders the quantitation of the negative charge through chemical analyses, as the exfoliation process creates an abundance of OH' ions. These difficulties necessitate the use of “innocent” cations such as alkali metals, as they have poor affinity for OH‘. This is particularly true for the heavier alkali cations Rb+ and Cs,+ and they are easier to quantify than Li“, Na+ or K”. Although the external OH' ions can largely be removed by rinsing, it is difficult to be certain of their total removal without acidification of the solution. This complicates the situation by increasing the ionic strength of the solution and providing protons, a competing cation which is virtually impossible to unambiguously detect. In addition, if the MOS, is in fact slowly oxidized further in solution over time, more OH’ ions will be generated. Even restacked MOS, or MOS, with encapsulated hydrated alkali cations may continue to undergo 210 redox chemistry with intercalated H,O (see Eq. (5.1)). Presumably the cations/OH' pairs would remain trapped between the layers, indicating at least the oxidation state at the time of flocculation, but would not reflect the continuum of oxidation states possible for the material over time. The results of our chemical analyses, which indicate a negative charge in the range 0.10—0.25 per Mo/W atom, are consistent with the ‘3'” and are more results of the analysis of fresh Li+ encapsulated samples reliable due to the use of the heavier alkali metal cations. Other studies involving the selective encapsulation of heavy metal ions in such as Ag+ and Pb2+ into MOS, and WS, have been conducted, and the results indicate a higher negative charge (0.5-0.8 per Mo/W atom).28 These chalcophilic cations, like the alkali cations, have a poor affinity for OH’; however, the samples have been prepared under significantly different experimental conditions, in which LiMS, was exfOliated in the presence of the cations under anaerobic conditions in solutions of moderate ionic strength, resulting in extremely rapid precipitation of the layers. These results suggest that a continuum of oxidation states are possible for MOS, and WS,, depending on the experimental conditions. 211 5.3.5 Oxidation reactions of LiMS,. In light of increasing evidence that these materials have a negative charge, LiMS, and exfoliated MS, were treated with oxidizing agents to see if truly neutral MS, could be obtained by oxidizing with I, or Br,. The oxidized products were characterized by X-ray and electron diffraction studies and their thermal behavior. Figure 5.8a contains the Selected Area Electron Diffraction pattern of LiMoS, oxidized with Br, in acetonitrile (MeCN). If LiMoS, is first exfoliated in water, then rinsed with acetonitrile, forming Li,_ n(MeCN),MoS, (n=0.75-0.85), and subsequently oxidized with Br, in acetonitrile, similar patterns can be observed. The pattern resembles the one found for the sample Ba0.08MoS,, in which a hexagonal array of diffraction spots resembling a “honeycomb” is formed by the superposition of the twinned 3a x a lattice caused by Mo-Mo distortions and a [[371 x 1/51 lattice formed by the ordering of the Ba2+ cations. The SAED pattern from oxidation of MOS, with Br, is caused by a superposition of two types of lattices as well; however, in this case the fit x fizz lattice is not due to Li+ cations ordering between the layers, but different Mo-Mo associations. Wypych and coworkers have reported that, upon treatment with I, or 212 H,Cr,O,, K0_33(H,O)yMoS, is oxidized to a material which contains a 1/3'a x [[371 lattice attributed to Mo-Mo associations.22 This material has a smaller a parameter than restacked MOS,. A closer inspection of Figure 5.8a reveals that the hexagons become increasingly distorted as one moves further away from the center spot (Figure 5.8b), indicating a mismatch in the two lattices. The 1/321 x 1/321 lattice has a smaller a parameter than the 1/321 x a lattice. If Li,,n(MeCN)xMoS, is treated with Br, it is possible to find electron diffraction patterns of a few crystals which contain only the 1/3-a x 1/371 lattice (Figure 5.8c), but the majority of the sample is a decomposition product. The material with the 1/371 x 1/3_a lattice is metastable and appears to be more susceptible to conversion to 2H-MOS, than the material with the 3a x a lattice. If one treats LiMoS, with Br, for a longer period of time or with sonication, the 1/3—a x l/ib lattice disappears, leaving a mixture of 2H-MOS, and material with the 1/3-0 x a lattice (Figure 5.9). 213 The pattern, which resembles those in Figure 5.6, is a combination of a twinned V3a x a orthorhombic lattice and a 43a x ‘13a hexagonal lattice of M082. The two lattices have slightly different cell parameters, which results in a distortion of the honeycomb as one moves away from the center of the pattern (B). 214 Figure 5.9. Oxidation Of Li1-n(H20)xM082 with Br; results in some particles which contain only the V3a x 43a lattice (A). If Li1-n(H20)xMoSZ or LiMoSz is oxidized over a longer period of time, or with sonication, the ‘13a x \/3a lattice disappears, leaving a mixture of 2H-M082 and M082 with the twinned 43a x a orthorhombic lattice (B). 215 B a 8 a 9.‘ <2- 8 V 9 (273a x 273a) d* (reciprocal angstroms) Figure 5.10. SAED pattern of a minority phase formed upon oxidation of LiW82 with Brz in MeCN which exhibits a 2‘13a x 243a lattice. 216 The electron diffraction patterns of WS, samples obtained upon oxidation with Br, are different from those of MOS, samples. Almost all patterns contained only the original l/3_a x a lattice, even with repeated sonication and stirring in excess Br, for a week. After a week a small amount of 2H-WS, could be found, and the amount of diffuse scattering was significantly increased due to decomposition to amorphous product(s). From these results one can conclude that either the product has not been oxidized or restacked WS, does not exhibit an alternate superlattice upon oxidation. Although the majority of the product exhibited no change, in almost every sample examined there was a small minority of crystallites which gave rise to the diffraction pattern shown in Figure 5.10a. The pattern contains the reflections attributed to the twinned 1/3—a x a lattice, but there are additional, extremely weak reflections which could be indexed as a 21/51 x 2/3_a lattice (Figure 5.10b). This suggests that even if WS, is able to have an alternate superlattice upon oxidation it is not the same as the lattice found in oxidized MOS,. 217 5.3.6 X-ray diffraction studies of oxidized MS,. LiMS, and Li,_n(MeCN)xMS, were also examined by transmission X- ray diffraction (Figures 5.11, 5.12). Due to the turbostratic nature of these samples induced by the exfoliation/flocculation process, only hkO reflections are visible from the 1/3_a x a lattice. Interpretation of the X-ray diffraction patterns is more difficult than the electron diffraction patterns in part because the peaks are broad. Furthermore, some hOI reflections of the 2H form of the dichalcogenide overlap with the unique hkO reflections from the 1/3_a x a lattice and the Ifla x 1/3—a lattice. For example, the 200 of the 1/3_a x 1/3-0 lattice overlaps with the 103 of 2H-MS,. In fact, the presence of the l/3_a x z/3—a lattice cannot be reliably determined by powder X-ray diffraction because the only reflection which is not subject to an overlap problem is the 100, which is extremely weak. The presence of the l/fi x a lattice is more easily monitored because the 120, a moderately strong reflection at d=2.10 A, does not overlap with peaks from either of the other two phases which may be present in the sample. 218 m C 71? 0 * a: t c o :3 ' o 2‘ N ‘3 9:: '9 c (U v z: 'k B '7) c o m '- o m E c or F , A llLJJJllllllllbllllllLJlIlllllllllllll 25 30 35 40 45 50 55 60 65 26 (degrees) Figure 5.11. Transmission X-ray diffraction (XRD) pattern of LiMoS, oxidized in MeCN with increasing time. (A) Initially the pattern contains the a x 1/32 lattice with two new peaks at 2.55 A ( solid black circle) and 1.56 A (asterisk). (B) a peak at 2.3 A appears, indexed as either the 200 of the I/3—a x 1/3—a lattice or the 103 of 2H-MOS,. (C) The 120 of the a x 1/371 lattice has disappeared. The pattern is mostly 2H-MOS,. 219 ’0? 9: A r: : 0%., 2‘ co 2 . 9 - o N 8 E - V V a .8— c Q 4—0 5 M L l I 1 L J I l l 1 4L 1 1 l 20 40 60 80 100 29 (degrees) Figure 5.12. Transmission XRD pattern of LiWS, oxidized with Br, in MeCN. The black circle marks a peak at 2.55 A which does not appear to be an hk0 reflection of any of the three expected lattices. The asterisk marks a contribution from the SaranTM wrap substrate. 220 The time dependent behavior of the oxidation of restacked MOS, (and LiMoS,) is illustrated in Figure 5.11. Initially the Viz x a lattice dominates (Figure 5.11a), with the appearance of a shoulder (d = 1.56 A) on the 310 reflection at 20 = ~59 degrees and a new peak at 20 = 35° ((1 = 2.55 A) which is not visible in the electron diffraction patterns and is not readily indexed to any of the 3 phases which may be present. At an intermediate stage a new peak appears at 29 = ~38° (d = 2.35 A) which, as mentioned previously, can be indexed as either the 200 of the V3—a x 1/3?! lattice or the 103 of 2H-MOS2 (Figure 5.11b). The 110 and the 310 reflections begin to shift to smaller (1 spacings (d = 2.78 to 2.71 A and 1.61 to 1.59 A). This is not surprising, because the V3_a x a lattice has the largest a parameter at 3.21 A; the 1561 x V321 lattice is slightly contracted, as evidenced by the electron diffraction patterns; and in 2H-MOS, a = 3.16 A. The 120 peak of the V371 x a lattice at 29 = ~38° (d = 2.10 A) is still present, and the shoulder at 20 = ~59° has become more pronounced. In the final stage the diffraction pattern no longer contains the peak at 29 = 35° (d = 2.55 A) or the 120 peak of the V37: x a lattice at 29 = ~38° (d = 2.10 A), and the 110 and the 310 reflections have shifted to even smaller (1 221 spacings (d = 2.69 A and 1.58 A). The pattern at this stage contains mostly MOS, which has converted to the 2H form. The transmission X-ray diffraction patterns of restacked WS, upon treatment with Br, (Figure 5.12) resemble those of oxidized restacked MOS,. The peak at 29 = 35° ((1 = 2.55 A) appears but, unlike in MoS,, does not disappear with time. Under more rigorous conditions or with time the peak at 29 = ~38° (d = 2.35 A) appears. Given that there is no evidence of a V371 x 1/321 lattice by electron diffraction, that peak is probably best indexed as the 103 reflection of 2H-WS,. The disappearance of the 120 peak of the V51 x a lattice at 29 = ~38° (d = 2.10 A) and the new peak at 29 = 35° (d = 2.55 A) happens upon heating the material, which converts it to 2H-WS, (Figure 5.13). The reflection X-ray diffraction patterns contain the basal 001 reflection at 6.05 A for WS, and 6.21 A for MoS,, indicating the de- intercalation of most or all of the lithium cations. The basal spacing of WS, is slightly contracted (0.1 A ) relative to 2H-MS,, but expands upon heating. Residual negative charge on MOS, is also substantiated by these electron diffraction patterns. The structure of restacked MOS, and WS,, which we recently solved from electron diffraction data and found to be W'I'e, type, consists of metal-metal associations to form zigzag chains.21 222 As mentioned previously, a material called 'lT-MOS,’ has been prepared by the oxidation of K0,33(H,O)yMOS2 and found to have a V3_a x V3—a lattice by X-ray diffraction and also by STM studies.”24 The starting material K0.3,(H,O)YMOS,, however, has also been studied by STM and found to have the same structural distortion as restacked MOS, and WS,.29 Furthermore, it is evident by the electron diffraction studies presented here that restacked MOS, can be at least partially converted to “IT-MOS,” upon treatment with Br, or 1,. The exterior of the particle appears to be oxidized, forming the V3_a x Via lattice, whereas the interior of the particle retains the original V3_a x a structure. Actually in their STM studies of lT-MOS, the authors conceded that occasionally they observed the fit: x a lattice found in their starting material and concluded that it was symptomatic of incomplete oxidation of those samples. 223 ’0? 9:! C 3 B B .2 E m C v .E‘ o m N c '- m . E l - A I11111111111111111111111111111111411111 25 30 35 4O 45 50 55 60 65 26 (degrees) Figure 5.13. Transmission XRD pattern of LiWS, oxidized with Br, (A) before and (B) after heating. The peak at 2.55 A (black circle) disappears, leaving 2H-WS,. 224 5.3.7 DSC studies of oxidized M8,. As mentioned earlier, restacked MOS, and WS, are metastable, and can be converted to 2H—MS, by an irreversible exothermic transition upon heating. The lT- form of MOS, with the V3_a x V3_a lattice has also been reported to be metastable with a comparable conversion.22 Figure 5.14 contains a DSC measurement Of oxidized MoS,, which has been identified by electron diffraction to contain both the fla x a and fl: x V3_a lattices, and a DSC measurement of restacked MOS, for comparison. Restacked MOS, undergoes conversion to 2H-MOS, at 98°C (at a heating rate of 5°C/min). The transition in the MOS, obtained by oxidizing LiMoS, with Br, in acetonitrile occurs at a lower temperature, the broad peak reaching a maximum at 90°C. Although by TEM it is evident that the material contains more than one kind of lattice, it is not possible to resolve two separate transitions. Li1_n(MeCN),MoS, oxidized with Br, converts rapidly, and a DSC measurement of the material could not be obtained. 225 -0.5 98°C 'IIIIIIIIIIITIITTTT llllllllllllll11111111111 IIIITIIII .l 40 60 80 100120140160180 Temperature (°C) Figure 5.14. Differential Scanning Calorimetry (DSC) plots of restacked and LiMoS, oxidized with Br,. Restacked MOS, exhibits an irreversible exothermic phase transition at 98°C, corresponding to its conversion to 2H- MoS,. LiMoS, oxidized with Br, undergoes a similar transition at 90°C. 226 ‘UTIIYI—ITIIV'I 0 50 100 150 200 250 300 350 Temperature (°C) h i- d -15b1111PJAII11111L111L‘ O 100 1 50 200 250 Temperature (°C) Figure 5.15. DSC plots showing the irreversible exothermic phase transition to 2H-WS,. (A) LiWS, (149°C) and Li,,,,(MeCN),WS, (189°C) oxidized with Br, in MeCN (B) Li,,,,(H,O) xWS, oxidized with Br, in H,O (183°C) and restacked WS, (195°C). 227 The situation is different and somewhat more complicated for the oxidation of WS, samples (Figure 5.15). Restacked WS, undergoes conversion to 2H-WS, at 195°C (at a heating rate of 5°C/min). WS, obtained by treating LiWS, with Br, in acetonitrile converts at 149°C, whereas Li,,n(MeCN),WS, treated with Br, in acetonitrile converts at 189°C (Figure 5.15a). By electron diffraction and transmission X-ray diffraction the materials are structurally identical, containing the 1/321 x a lattice. The oxidation was carried out in acetonitrile because aqueous exfoliated WS, decomposes fairly rapidly upon exposure to Br,, but if the aqueous reaction is isolated within a few minutes it is possible to see a shift in the transition to lower temperatures (Figure 5.15b). By transmission X-ray diffraction this material has the V371 x a lattice with a small shoulder at 2.55 A, indicating that it has the same structure as the products of the oxidation of Li,_n(MeCN),WS, and LiWS,. The structural behavior of restacked WS, and LiWS, is clearly different from that of restacked MOS, and LiMoS, upon treatment with Br,. Despite the lack of a M321 x V371 phase in the WS, system, there is a dramatic shift in conversion temperature, from 195°C to 149°C, which is reminiscent of the shift in the transition temperature observed in the MOS, system. This shift suggests that oxidation has in fact taken place. 228 Furthermore, reflection X-ray diffraction indicates no residual cations between the layers (i.e., the d spacing is 6.1 A). The fact that both restacked and oxidized WS, have the V3_a x a lattice implies that, at least in WS,, the shift in conversion temperature is not dependent on a structural rearrangement, but rather the degree to which the sample is oxidized. The discrepancy in transition temperature between Br, treated LiWS, and Br, treated Li,,n(MeCN),WS, is probably due to limited diffusion in Lil, n(MeCN),,WS,. In LiWS, the acetonitrile can solvate the lithium cations, swelling the layers apart and allowing the oxidant to access the layers more intimately, whereas Li,,n(MeCN),WS, does not swell in acetonitrile; hence the interior of the particles is inaccessible to the oxidizing agent and the material remains trapped in the incompletely oxidized state that it achieved upon exfoliation in water. Studies of the structure dependence upon oxidation state in MOS, (and WS,) are complicated in part because 2H—MOS, undergoes a phase transition upon lithiation in which the coordination environment changes from trigonal prismatic to octahedral. If only a small amount of lithium is introduced the material can remain in the 2H-MOS, structure.30 It has also been observed that the material does not convert to the octahedral structure homogeneously with small amounts of lithium, resulting in a two phase system.18a Electrochemical studies which start with 2H-MOS, are difficult 229 to interpret due to this problem.“ The V3_a x a octahedral phase can be used as a starting point, however. Electrochemical studies of Li,MoS,, K,(H,O)yMoS,, and K,(H,O)yWS, which start from the octahedral structure indicate the presence of 46 different phases in the range x = 0 to x = l.'8b' 32 Although some authors have attributed these many phases to cation ordering in the gallery“, others have attributed them to charge density waves.32 The different observed superlattices in MOS, suggest that charge density waves are responsible for at least some of these phases. Wypych and coworkers have proposed a scheme containing three different lattices: a 2a x 2a for KO.,MOS,, V32 x a for K,,,(H,O),MOS,, and a V321 x V3—a for lT-MOS,.24 A comparison of the electrochemical behavior of Kx(I-I,O)yMoS, and K,(H,O),WS, reveals that the two materials behave differently at low x values”, consistent with the different electron diffraction results observed in the reaction of LiMoS, and LiWS, with Br,. 230 5.3.8 Acid restacked M8,. When LiMS, is exfoliated in concentrated acid the product should form a proton bronze, H,MS,. The transmission X—ray diffraction patterns for the reaction of LiMoS, and LiWS, with concentrated HCl (Figure 5.16) show that the materials retain the Viz x a lattice and a peak appears at 2.55 A. The products were examined by electron diffraction and exhibited only the V361 x a lattice with no evidence of an additional reflection at 2.55 A. In this case, the structure of HxMos, is the same as that of PLWS,, exhibiting no evidence of a Viz x V32 lattice. The reflection X-ray diffraction patterns exhibit basal 001 spacings of 6.05 A. If LiWS, is exfoliated directly in concentrated HCl the DSC shows an exothermic transition at 155°C, whereas if it is first exfoliated in H,O and then concentrated acid is added it exhibits a transition temperature of 191°C (Figure 5.17). Presumably the x values in the products of each reaction, H,WS,, are different. The discrepancy in the behavior of LiWS, and Li,,n(H,O),WS, in acidic conditions is similar to the discrepancy upon oxidation with Br,, in which the difference in conversion temperatures 231 110 A (D .".'.' g 0 5 o 9 E S! m e (U V .2; ' A (D C Q) fill E tllllrllerlLUIermrlnllllLtllllrlr 25 30 35 4O 45 50 55 60 65 26(degrees) Figure 5.16. Transmission XRD pattern of (A) LiWS, and (B) I-LMOS, Obtained from LiMoS, exfoliated in concentrated HCl. Black circle marks peak at 2.55 A. 232 I Jillllllllllllll[1111;111111141 .— h -1 bIIILlllIJJIIIIJI11111111]IIII- 0 50 100 150 200 250 300 Temperature (°C) Figure 5.17. DSC plots of HXWS, obtained from the reaction of HCl with (A) LiWS2 (155°C) and Li,,n(H,O),WS, (191°C). 233 could be attributed to a diffusion problem. Exfoliation in concentrated acid provides more oxidizing conditions than exfoliation in H,O before addition of acid. The amount of oxidation during the exfoliation process is important because there is a competition between oxidation and proton encapsulation. Once the material has flocculated its oxidation is limited by diffusion; hence the discrepancy in transition temperatures. LiMoS, exfoliated in concentrated acid does not exhibit a shift in its exothermic transition to a lower temperature. The structure of I-LMOS, samples prepared from the reaction of LiMoS, and Li,_n(H,O),MoS, with acid resembles that of their W analogues more than in Br, solution. There is no evidence of the [[321 x V3—a phase in acidic conditions. This suggests that the Russian scientists’ hypothesis that MOS, is fully oxidized upon treatment with acid is not correct; only that protons are now the encapsulated counter-ion.l7 As mentioned previously, the authors who first reported the exfoliation of MOS, believed that the negative charge of the layers was due to OH' groups associated with the basal planes of the layers. It was believed that exfoliated MOS, (and WS,) has a point of zero charge (PZC) upon acidification. The authors used the different PZC to create a restacked composite material with alternating M08, and WS, layers, forming a superstructure along the 001 axis.16c Their evidence for a 234 superstructure of alternating MOS, and WS, layers was poor, however; a better interpretation of these results is that the introduction of protons to a homogeneous mixture of exfoliated layers resulted in rapid flocculation, creating a homogeneous mixture of MOS, and WS, in the product. They also developed the concept that the PZC occurs at slightly different pH for the edge OH' sites and basal OH‘ sites, and used this to create a House-Of- Cards structure for MOS,.16b Their evidence for this kind of structure was a decrease in the intensity of the 001 reflections in the X-ray diffraction pattern. However, the intensity of the 001 reflections, which is indicative of the degree of order in the flocculated layers, can be influenced by a variety of factors (the ionic strength of the solution, for example), and does not constitute proof of a point of zero charge on the layers. The mysterious peak at 2.55 A in the transmission X-ray diffraction patterns of HXMS2 and LiMS, oxidized with Br, is puzzling, but shifting of the layers relative to one another could allow hOI reflections to appear. A 102 reflection has been observed in some samples of 2H-MOS, at that d spacing. The peak at 2.3 A is almost certainly the 103 of 2H-MS,. An illustration of the different superlattices observed in MOS, and WS, due to M-M associations can be found in Figure 5.18. An ideal 1T- MS, lattice (not experimentally observed) can distort to form zig—zag 235 Figure 5.18. Schematic illustrating (A) an ideal lT-MS2 lattice, (B) a V371 x a lattice with infinite zig-zag metal chains, (C) a V3_a x V3_a lattice with trimers, and (D) a 215a x 2 3a lattice with possible tetramers formed due to M—M distortions. 236 chains, giving rise to the orthorhombic V51 x a lattice found in restacked MOS, and WS,. It can trimerize, giving rise to a V3_a x V371 lattice found in MOS, upon oxidation with Br, or I, (Figure 5.180). Finally, when WS, is treated with Br,, a minority phase with a 2 V3_a x 2 V3—a lattice is formed. This distortion could be due to a number of M-M associations, perhaps a trimerization similar to Figure 5.18c or a tetramerization (Figure 5.18d). 5.3.9 Thermopower measurements. As mentioned previously, the reduction of 2H-MoS, with lithium results in a structural transformation from trigonal prismatic to octahedral metal coordination and a change in properties from semiconducting to metallic. These changes have been explained using. a depiction of the band structures for LiMoS, in both coordination environments, which revealed that there was a net stabilization in energy upon conversion to octahedral that resulted in a half-filled band with predominantly d orbital character.‘8 Restacked MOS, and WS, retain octahedral coordination about the metal atom,21 and their conductivity has been reported elsewhere” to be metallic. If one simply removes electrons from the half filled band which was published for LiMoS,, an idealized octahedral d3 system, one should 237 expect an n-type conductor for the d2 system (Figure 5.19). Thermopower measurements on restacked MOS, and WS,, however, indicate that they are p-type (hole) conductors (Figure 5.20), with room temperature values of 50 and 20 ttV/K, respectively. The structural distortion in restacked MOS, and WS, to form zig-zag chains obviously should result in a change in the band structure. In fact, distortion of octahedral ML, (12 systems to form zig-zag chains has been predicted by Extended Huckel Tight Binding calculations” and, for a fully oxidized (12 system, results in a half filled band (Figure 5.19). If, however, the system is actually d”, as is the case with restacked MOS, and WS,, residual negative charge results in a material with p-type conductivity, consistent with our experimental observations (Figure 5.19). 238 dz2 zig-zag _ b chains Figure 5.19. Schematic band diagram illustrating the distortion of an ideal octahedral system to form zigzag chains. The ideal system describes an n-type conductor whether it is d2 (black fill) or d“ (black fill + stripes). The distorted system is half filled for d2 and more than half filled for d“, resulting in a p-type metallic conductor. 239 60 11111 P 50 IIIIII 40 IIIIIIIIITTI D P —L o ‘> > O O O O O WTI.IIIII b lllllllllllllllllllillll11_LIJ 80 120 160 200 240 280 320 360 400 Temperature (°C) Figure 5.20. Thermopower measurements of restacked MOS, (triangles) and WS, (circles). The relatively small magnitude and slope of the data is consistent with that of a p-type metallic conductor. 240 5.4 Conclusions MOS, and WS, retain some negative charge upon exfoliation and flocculation. Although it is possible to incorporate neutral species, cations can also be encapsulated without a detectable co—encapsulated anion, usually giving rise to very ordered reflectance X-ray diffraction patterns.9"3'l7 In the studies presented here hard, electropositive alkali cations have been encapsulated in MOS, and WS, under neutral synthetic conditions. The choice of these cations and these synthetic conditions minimizes the risk of co-encapsulation of OH’ ions. The cations order in the gallery, forming additional superlattices. The range of negative charge on the layers appears to be 015-025 for both MOS, and WS,, which is consistent with measurements on samples prepared in a similar fashion by other researchers.17 Structural characterization of restacked MOS, also supports a negative charge. Restacked MOS, and WS, have been found by electron diffraction to have the same superlattice as K,(H,O),MOS,.2'2° Treatment of restacked MOS, with Br, results in a change from a V371 x a superlattice to a 1/371 x V321 superstructure, which is analogous to the superstructure observed in “1T-MOS,”, the oxidation product of K(,.,MOS,.22 This change in structure upon exposure to an oxidizing agent implies that restacked 241 MOS, is incompletely oxidized, and is probably better formulated as LixMoS, or Lix(H,O)yMoS,. Restacked WS, does not appear to exhibit a structural change upon exposure to Br,. However, both restacked MOS, and WS2 exhibit irreversible exothermic transitions which are shifted to lower transition temperatures upon treatment with Br,. The shift is quite significant in WS,, from 195°C to 149°C in the Br,-treated sample, indicating that oxidation has taken place. It is somewhat surprising, however, that oxidation of restacked WS, does not result in an alternate superlattice analogous to the Viz x V3_a superstructure observed in oxidized MOS,. The products of all thermal conversions are the semiconducting phase 2H- MS,, which requires an expulsion of the residual negative charge and explains why the oxidation state affects the transition temperature. The expulsion of the residual negative charge likely results in the reduction of residual H,O (or H+ in H,MS,), forming H, gasf"4 242 A 2H-MoSz (1)] H20 + * + HCI + Br2/MeCN l Lio.2M032 (2) i HxMOSZ (2) (130 X01 \ (l/3a Xa) + Br2/ _ M eCN LIO.2M082 (2) N3a Xa) ‘A “1T-MOSQ” (3) (1/3a x V3a) A 98°C A¢9OOC A V V l2H-MoSz (1)| |2H-Mosa (1)| I2H-M082 (1)] B le-ws2 (1 )l H20 + + HCI + Br2/MBCN I UO.2W32 (2)|~ \ . Br2/ 'ws2“ 13) («30 X0) MBCK (‘30 X0) 7 I \ “W82“ (3) (‘30 Xa) Lio.2W82 (2) minority (2V3ax2\/3a) + "we; (3) A 195°C A *14goc