SYNTHESIS AND CHARACTERIZATION OF NOVEL COMPLEX IRON OXIDES WITH LAYERED AND TUNNEL STRUCTURES By Shaun R. Bruno A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Chemistry—Doctor of Philosophy 2014     ABSTRACT SYNTHESIS AND CHARACTERIZATION OF NOVEL COMPLEX IRON OXIDES WITH LAYERED AND TUNNEL STRUCTURES By Shaun R. Bruno Novel cathode materials for lithium ion batteries were synthesized via lithium for sodium ion exchange from the known compounds, ß-NaFeO2 and NaFeTiO4. The resulting lithium analogs of these known sodium compounds, T-LiFeO2 and LiFeTiO4, contain tunnel-like structures that were characterized using Rietveld refinement of Powder X-ray diffraction, electrochemical measurements, Mössbauer spectroscopy, thermogravimetric analysis, and inductively coupled plasma spectroscopy. Similarly characterized, α- and ßNaFe2O3, with a double layered rock salt structure, were synthesized for the first time as a bulk powder using an oxygen pressure regulation method that provided the appropriate conditions for the two polymorphs to form. Further, investigation into T-LiFeO2 and NaFe2O3 by doping other transition metals into the iron position, to control specific properties of the two materials was performed with success. T-LiFeO2 and the parent phase, ß-NaFeO2, were doped with up to 0.1 and 0.15 parts of cobalt per formula unit respectively. NaFe2O3 was successfully doped with cobalt up to 0.5 moles with pure phases of both the α-NaFe1.5Co0.5O3 and ß-NaFe1.5Co0.5O3 forming. Manganese doping into NaFe2O3 also showed the formation of the α- phase.     Probing the Fe3+/4+ redox potential of both LiFeO2 and LiFeTiO4 resulted in the decomposition of each. The cobalt doped LiFeO2 though did show a greater possibility of cycling at Fe3+/4+ redox potential, but also resulted in a reaction with the organic electrolyte. Chemical deintercalation of LiFeO2 and LiFeTiO4 were performed with resulting in the decomposition of LiFeTiO4. LiFeO2, indicated successful lithium deintercalation with preliminary Mössbauer results illustrating Fe4+ formation. Both LiFeTiO4 and LiFeO2 successfully cycled electrochemically at the Fe2+/3+ redox potential, with the new calcium ferrite structure polymorph LiFeTiO4 cycling 17 % higher capacity than the previously reported spinel and rock salt structure compounds.     Copyright by SHAUN R. BRUNO 2014     This thesis is dedicated to my family; Ashley and Ryan Bruno, Lisa and David Heagle, Barbara Skodak, Dave and Mary Sebesta, John and Betty Sebesta, Robert and Gloria Carpenter, Shane, Dane, Lindsey and Kailey. v    ACKNOWLEDGEMENTS I would like to thank my advisor; Professor Viktor Poltavets for his advice and guidance during my Ph.D. research, my colleague and lab mate Dr. Colin Blakely for the conversations and teamwork we had throughout our tenure at MSU, Dr. Richard Staples for his help with X-ray diffraction and instrumentation, and my Ph.D. committee Professor Hamann, Professor Weliky, and Professor Mahanti for the help and insight they provided when I needed it. vi    TABLE OF CONTENTS LIST OF TABLES .....................................................................................................ix LIST OF FIGURES ......................................................................................................x Chapter 1. Introduction......................................................................................1 Chapter 2. Instrumentation.................................................................................4 Powder X-ray Diffraction.................................................................4 Inductively Coupled Atomic Emission Spectroscopy......................9 Mössbauer Spectroscopy................................................................10 Electrochemistry.............................................................................15 Thermogravimetric Analysis and Differential Scanning Calorimetry.....................................................................................16 Scanning Electron Microscopy......................................................16 Transmission Electron Microscopy................................................19 REFERENCES...............................................................................25 2.1 2.2 2.3 2.4 2.5 2.6 2.7 Chapter 3. 3.1 3.2 3.3 3.4 3.5 3.6 Chapter 4. 4.1 4.2 4.3 4.4 4.5 4.6 4.7 Synthetic Methods..........................................................................27 Pechini (Sol-Gel)............................................................................27 Solid State (Ceramic).....................................................................28 Reduction........................................................................................28 Partial Oxygen Pressure Regulation...............................................30 Moisture and Oxygen Sensitive Techniques..................................32 Chemical oxidation and reduction .................................................34 REFERENCES...............................................................................36 Synthesis of NaFe2O3 Using Partial Oxygen Pressure Regulation Method............................................................................................38 Introduction....................................................................................38 Synthesis of NaFe2O3 via controlled oxygen pressure...................41 Oxidation via sodium deintercalation.............................................45 Transition Metal Doping................................................................46 Results and Discussion...................................................................47 Conclusion......................................................................................72 Future Direction..............................................................................72 REFERENCES...............................................................................76 vii    Chapter 5. 5.1 5.2 5.3 5.4 5.5 5.6 5.7 Chapter 6. 6.1 6.2 6.3 6.4 6.5 6.6 6.7 Chapter 7. 7.1 7.2 Synthesis of Tunnel Polymorph LiFeO2 and Characterization of the Electrochemical Performance of Cobalt Doping............................80 Introduction....................................................................................80 Synthesis of ß-NaFeO2 and NaFe1-xCoxO2.....................................86 Ion exchange of NaFe1-xCoxO2 to LiFe1-xCoxO2 (0≤x≤0.15)..........86 Lithium intercalation and deintercalation of LiFeO2 via chemical reduction and oxidation..................................................................87 Results and Discussion...................................................................87 Conclusion....................................................................................136 Future Direction............................................................................136 REFERENCES.............................................................................140 Synthesis of the Novel Polymorph, LiFeTiO4, with Calcium Ferrite Structure.......................................................................................143 Introduction..................................................................................143 Synthesis of NaFeTiO4.................................................................145 Ion Exchange of NaFeTiO4 to LiFeTiO4......................................146 Chemical intercalation and deintercalation..................................147 Results and Discussion.................................................................148 Conclusion....................................................................................175 Future Direction............................................................................175 REFERENCES.............................................................................182 Mössbauer Data Analysis.............................................................185 Introduction..................................................................................185 Data Fitting...................................................................................185 REFERENCES.............................................................................232 viii    LIST OF TABLES Table 4.1 Crystallographic Data for the Rietveld Refinement of ß-NaFe2O3 and ßNa0.78Fe2O3 from SXPD data. (Goodness of Fit Parameters: 2 = 5.87, Rwp = 10.6%).........................................................................................55 Table 4.2 Phase composition of nominal “NaFe2O3” stoichiometry at different partial O2 pressures at 850 ºC.................................................................56 Table 4.3 NaxFe2O3 Cell parameters a, c and volume determined from Le Bail fit from PXD................................................................................................67 Table 5.1 Common transition metal oxide cathode materials.................................82 Table 5.2 Polymorphs of LiFeO2, structure, preparation, preparation procedure, and best electrochemical performance....................................................83 Table 5.3 Cell parameters and volume of cobalt doped ß-NaFeO2 determined from Le Bail fit................................................................................................93 Table 5.4 Cell parameters and volume of cobalt doped T-LiFeO2 determined from Le Bail fit..............................................................................................101 Table 5.5 Mössbauer parameters of the chemically deintercalated Li0.42FeO2 fit with three sites......................................................................................131 Table 5.6 Mössbauer parameters of the as prepared LiFeO2 fit with one site.........................................................................................................133 Table 5.7 Mössbauer parameters of the chemically intercalated Li1.57FeO2 fit with three sites..............................................................................................135 Table 6.1 Crystallographic data for the Rietveld refinement of LiFeTiO4 (space group Pnma), a=8.9206(5) Å, b=2.9595(3) Å, c=10.7103(7) Å...........163 Table 6.2 Crystallographic data for the Rietveld refinement of Li2FeTiO4 (space group Pnma), a=9.0525(3) Å, b=2.9566(2) Å, c=10.7575(5) Å...........165 Table 6.3 Mössbauer spectroscopy of LiFeTiO4 and lithium deintercalated Li0.33FeTiO4...........................................................................................174 ix    LIST OF FIGURES Figure 2.1 Mössbauer schematic illustrating the decomposition of 57Co into 57Fe resulting in the emission of energy into an absorber, in this case an iron source, at which point the absorber emits non-resonant, resonant, conversion, Auger electrons, and x-rays with other various energies.....12 Figure 2.2 Schematic of Mössbauer isomer shift, quadrupole splitting and hyperfine splitting illustrating the effect each has upon the Mössbauer spectrum..14 Figure 2.3 The relationship between kI, kD, K, θ, and λ. The initial wavefront(green), the top of the triangular shape is the point at which the wavefront is diffracted by the plane of atoms, resulting in the diffracted wavefront (blue), and the change in the vector due to diffraction (black) is K..........................................................................................................20 Figure 4.1 Structure models of - and -NaFe2O3. The FeO6 octahedra are highlighted. For - NaFe2O3 atoms in the shared Na/Fe position are shown as large blue spheres and small red balls represent O ions. For NaFe2O3 large green spheres correspond to Na ions..............................39 Figure 4.2 Experimental setup for the synthesis of α- and ß-NaFe2O3 under controlled partial O2 pressure..................................................................45 Figure 4.3 PXD of the resulting product of γ-NaFeO2 (red ticks) and Fe3O4 (green ticks) after annealing the mixture of α-NaFeO2, Fe powder and Fe2O3 under flowing house nitrogen at 850 °C.................................................48 Figure 4.4 PXD of the resulting product of ß-NaFe2O3 (red ticks), γ-NaFeO2 (green ticks), and Fe3O4 (blue ticks) after annealing the mixture of α-NaFeO2, Fe powder and Fe2O3 at a pO2 below ∼1.0×10−10 atm at 850 °C............49 Figure 4.5 PXD of the resulting product of ß-NaFe2O3 (red ticks) after annealing the mixture of α-NaFeO2, Fe powder, and Fe2O3 between the a pO2 of ∼1.0×10−10 and ∼1.0×10−19 atm at 850 °C............................................50 Figure 4.6 PXD of the resulting product of α-NaFe2O3 (red ticks) after annealing the mixture of α-NaFeO2, Fe powder, and Fe2O3 below pO2 ∼1.0×10−19 atm at 850 °C...........................................................................................51 x    Figure 4.7 Schematic section of phase composition vs. partial O2 pressure diagram at 850 °C for Na/Fe=1/2..........................................................................52 Figure 4.8 Observed (crosses), calculated (solid line), difference (bottom solid line) profiles and Bragg reflections (tick marks) for the final Rietveld refinement of ß-NaFe2O3 and ß-NaxFe2O3, from SXPD data. The tick marks correspond to the positions (from top down) of ß-NaxFe2O3, ßNaFe2O3, and Fe3O4. The section at higher angles (>16° 2θ) is enlarged by a factor of 20 for clarity.....................................................................54 Figure 4.9 Difference Fourier map for α-NaFe2O3 indicating large residual electron density (green) around the Na/Fe and O3 positions. Atomic positions are denoted for this layered structure and Fe–O octahedra are highlighted. Blue and red spheres represent atoms in the Na/Fe and O positions, respectively.............................................................................................59 Figure 4.10 PXD of unwashed sodium deintercalated ß-NaxFe2O3 (red ticks) with the byproduct NaBr (green ticks)..................................................................63 Figure 4.11 Cell parameter a as a function of moles of sodium in NaxFe2O3 Le Bail refined from PXD....................................................................................64 Figure 4.12 Cell parameter c as a function of moles of sodium in NaxFe2O3 Le Bail refined from PXD....................................................................................65 Figure 4.13 Volume as a function of moles of sodium in NaxFe2O3 Le Bail refined from PXD................................................................................................66 Figure 4.14 NaFe1.5Co0.5O3 synthesized at 850 °C at a pO2 of about 10-23 atm resulting in the formation of the α-NaFe2O3 structure type (red ticks)...68 Figure 4.15 NaFe1.75Co0.25O3 synthesized at 850 °C at a pO2 of about 10-23 atm resulting in the formation of the ß-NaFe2O3 structure type (red) and γNaFeO2 (green). .....................................................................................69 Figure 4.16 PXD of NaFe1.85Co0.15O3 synthesized at 850 °C at a pO2 of about 10-23 atm resulting in the formation of the ß-NaFe2O3 structure type (red), γNaFeO2 (green), and Fe3O4 (blue)..........................................................70 Figure 4.17 NaFe1.5Mn0.5O3 synthesized at 850 °C at a pO2 of about 10-23 atm resulting in the formation of the α-NaFe2O3 (red), structure type γNaFeO2 (blue), and Fe3O4 (green)..........................................................71 xi    Figure 4.18 Magnetic response of NaFe2O3 as a function of temperature. The measurement was preformed from room temperature to 4 K at 5000 Oe............................................................................................................73 Figure 4.19 Thermopower (green) and electrical conductivity (black) measurements of ß-NaFe2O3 with electrical conductivity showing transitions occurring at about 250 K and at about 100 K (black).............................................75 Figure 5.1 PXD of pure ß-NaFeO2 synthesized at 1000 °C in air, red ticks represent expected reflections of pure phase..........................................................88 Figure 5.2 Cell parameter a of ß-NaFeO2 as a function of cobalt doping......................................................................................................89 Figure 5.3 Cell parameter b of ß-NaFeO2 as a function of cobalt doping......................................................................................................90 Figure 5.4 Cell parameter c of ß-NaFeO2 as a function of cobalt doping......................................................................................................91 Figure 5.5 Cell volume of ß-NaFeO2 as a function of cobalt doping......................................................................................................92 Figure 5.6 Thermogravimetric analysis (black) and differential scanning calorimetry (blue) of the lithium for sodium ion exchange. The endothermic peak beginning at about 190 °C and exothermic peak at about 212 °C, (exothermic is positive µV/mg), are indicative of melting and ionic diffusion respectively..............................................................94 Figure 5.7 PXD of the unwashed lithium for sodium ion exchange of ß-NaFeO2 and LiNO3 to produce LiFeO2 (red tick marks) and NaNO3 (green tick marks)......................................................................................................96 Figure 5.8 Cell parameter a of T-LiFeO2 as a function of cobalt doping......................................................................................................97 Figure 5.9 Cell parameter b of T-LiFeO2 as a function of cobalt doping......................................................................................................98 Figure 5.10 Cell parameter c of T-LiFeO2 as a function of cobalt doping......................................................................................................99 volume of T-LiFeO2 as a function of cobalt Figure 5.11 Cell doping....................................................................................................100 xii    Figure 5.12 PXD and Rietveld refinement of LiFeO2 Rwp=1.49. Observed (black crosses), calculated (solid red line), difference (solid blue line) profiles, and Bragg reflections (tick marks) for the final....................................102 Figure 5.13 Rietveld Refinement of PXD Li0.57FeO2 Rwp=3.90. Observed (black crosses), calculated (solid red line), difference (solid blue line) profiles, and Bragg reflections (tick marks) for the final Rietveld refinement...103 Figure 5.14 Synchrotron diffraction of Li1.57FeO2 Rwp=5.73 and χ2=1.969. Observed (black crosses), calculated (solid red line), difference (solid blue line) profiles and Bragg reflections (tick marks) for the final Rietveld refinement.............................................................................................106 Figure 5.15 Thermogravimetric analysis of T-LiFeO2 in oxygen up to a final temperature of 1000 °C.........................................................................107 Figure 5.16 PXD of the resulting powder from the oxygen TGA of T-LiFeO2, red ticks are indicative α-LiFeO2 with the rock-salt structure....................108 Figure 5.17 Thermogravimetric analysis of Li0.42FeO2 in oxygen up to a final temperature of 725 °C...........................................................................109 Figure 5.18 PXD of the resulting powder from of Li0.42FeO2 after TGA in oxygen red tick are indicative of the compound Fe0.873(Li0.365Fe1.2915)O4.........110 Figure 5.19 Thermogravimetric analysis of T-LiFeO2 in nitrogen up to a final temperature of 1000 °C with a final mass loss of about 2 %................111 Figure 5.20 PXD of the resulting powder from the nitrogen TGA of T-LiFeO2, red ticks are indicative Fe3O4, but is likely to be lithium doped Fe3O4......112 Figure 5.21 Thermogravimetric analysis of ß-NaFeO2 in nitrogen to a final temperature of 250 °C for analysis to determine structural stability for high temperature electrochemical cycling............................................113 Figure 5.22 PXD of the resulting powder from the nitrogen TGA of ß-NaFeO2, red ticks are indicative ß-NaFeO2...............................................................114 Figure 5.23 Thermogravimetric analysis of ß-NaFeO2 in oxygen to a final temperature of 350 °C for analysis to determine structural stability for high temperature electrochemical cycling............................................115 Figure 5.24 PXD of the resulting powder from the oxygen TGA of ß-NaFeO2, red ticks are indicative of ß-NaFeO2...........................................................116 xiii    Figure 5.25 Thermogravimetric analysis of the sodium deficient ß-Na0.79FeO2 in nitrogen to a final temperature of 975 °C..............................................117 Figure 5.26 PXD of the resulting powder from the nitrogen TGA of ß-Na0.79FeO2, ßNaFeO2 (red) and Na3Fe5O9 (green) were the resulting phases............118 Figure 5.27 Thermogravimetric analysis of the sodium deficient ß-Na0.79FeO2 in oxygen to a final temperature of 725 °C...............................................119 Figure 5.28 PXD of the resulting powder from the oxygen TGA of ß-Na0.79FeO2, ßNaFeO2 (red) and Fe2O3 (green) were the resulting phases..................120 Figure 5.29 Thermogravimetric analysis of ß-NaFe0.9Co0.1O2 in nitrogen to a final temperature of 1000 °C.........................................................................121 Figure 5.30 PXD of the resulting powder from the nitrogen TGA of ßNaFe0.9Co0.1O2, ß-NaFeO2 (red) and ß-Na1-xFeO2 (green) were the resulting phases.....................................................................................122 Figure 5.31 Thermogravimetric analysis of the ß-NaFe0.9Co0.1O2 in oxygen to a final temperature of 1000 °C.........................................................................123 Figure 5.32 PXD of the resulting powder from the nitrogen TGA of ßNaFe0.9Co0.1O2, ß-NaFeO2 (red) and ß-Na1-xFeO2 (green) were the resulting phases.....................................................................................124 Figure 5.33 Cyclic voltammogram of T-LiFeO2 between 4.0 and 4.6 V probing the Fe3+/4+ redox potential showing the reaction of T-LiFeO2 with the electrolyte..............................................................................................125 Figure 5.34 Integrated cyclic voltammogram of T-LiFeO2 illustrating what would normally be the deintercalation of lithium, but this is the reaction of TLiFeO2 with the electrolyte, this is also seen by absence of lithium intercalation...........................................................................................126 Figure 5.35 High voltage cycling of LiFe0.9Co0.1O2 between 3.5 and 4.6 V probing the Fe3+/4+ redox couple, the lithium deintercalation peak is present at about 4.4 V and where a distinct intercalation peak does not exist due to the compound’s reaction with the electrolyte.......................................127 Figure 5.36 Low voltage cyclic voltammogram of LiFe0.9Co0.1O2 probing the Fe2+/3+ redox couple between voltages of 1.5 and 3.0 V, with lithium intercalation occurring at about 1.8 V and deintercalation occurring at about 2.0 V............................................................................................128 xiv    Figure 5.37 Mössbauer spectrum of Li0.42FeO2 after chemical lithium intercalation fit with 3 sites the total fit (black) and the three sites correlating to isomer shifts are 0.129 (red), -0.09 (blue), and 0.53 (green)............................130 Figure 5.38 Mössbauer spectrum of the as prepared LiFeO2 with one site..............132 Figure 5.39 Mössbauer spectrum of Li1.57FeO2 after chemical lithium intercalation fit with 3 sites, the total fit (black) and the three sites correlating to isomer shifts are 0.161 (green), 0.438 (red), and 0.494 (blue)..........................134 Figure 5.40 Proposed structure of tunnel NaCoO2 resulting from the ion exchange using NaNO3 with the known compound KCoO2.................................138 Figure 5.41 Proposed structure of tunnel LiCoO2 resulting from the ion exchange using LiNO3 with the known compound KCoO2..................................139 Figure 6.1 Single crystal diffraction of Anosovite, Ti3O5, formed during initial attempt to synthesize NaFeTiO4 from literature preparation. Space group Cmcm, a=3.7097(2)Å, b=9.7308(4)Å, c=9.9716(5)Å, and volume 357.96(3)ų from a molybdenum source single crystal diffractometer........................................................................................148 Figure 6.2 The mixture containing NaFeTiO4 (green ticks) and Na0.79Fe0.8Ti1.2O4 (red ticks)..............................................................................................149 Figure 6.3 The mixture containing NaFeTiO4 (green ticks) and Na0.75Fe0.75Ti0.25O2 with the α-NaFeO2 structure type (red ticks)........................................150 Figure 6.4 The PXD pattern of NaFeTiO4 (green ticks) was determined to be a pure phase sample.........................................................................................151 Figure 6.5 Thermogravimetric analysis of calcium ferrite polymorph LiFeTiO4 in nitrogen to a final temperature of 1050 °C............................................153 Figure 6.6 PXD of LiFeTiO4 after TGA in nitrogen to 1050 °C. Red ticks illustrate peak reflection locations of the spinel structure LiFeTiO4...................154 Figure 6.7 Thermogravimetric analysis of calcium ferrite polymorph LiFeTiO4 in oxygen to a final temperature of 1050 °C.............................................155 Figure 6.8 PXD of LiFeTiO4 after TGA in oxygen to 1050 °C. Red ticks illustrate peak reflection locations of the spinel structure LiFeTiO4...................156 xv    Figure 6.9 SEM images of as prepared LiFeTiO4 illustrating the particle size of the compound in the micrometer regime, scale bar represents 5 µm..........157 Figure 6.10 SEM images of as prepared LiFeTiO4 illustrating the particle size of the compound in the micrometer regime, scale bar represents 5 µm..........158 Figure 6.11 SEM images of as prepared LiFeTiO4 illustrating the particle size of the compound in the micrometer regime, scale bar represents 500 nm......159 Figure 6.12 Observed (black crosses), calculated (red line), difference (blue) profiles, and Bragg reflections (tick marks) for the final Rietveld refinement of LiFeTiO4 from PXD data, Rwp=4.42 and χ2=0.260...............................162 Figure 6.13 Observed (black crosses), calculated (red line), difference (blue) profiles, and Bragg reflections (tick marks) for the final Rietveld refinement of Li2FeTiO4 from PXD data, Rwp=5.50 and χ2=0.183.............................164 Figure 6.14 Cyclic voltammogram of LiFeTiO4 between 2.8 and 3.7 V decomposing after the initial cycles as also illustrated by the low current.................166 Figure 6.15 Integrated cyclic voltammogram illustrating the voltage as a function of capacity of LiFeTiO4 in the attempt to probe the Fe3+/4+ oxidation states resulting in the decomposition of the tunnel structure..........................167 Figure 6.16 Capacity as a function of the cycle number showing the instability of the compound with no indication of a cycling trend...................................168 Figure 6.17 Cyclic voltammogram of LiFeTiO4 between 1.4 and 3.0 V showing relative stable cycling intercalation and deintercalation of the Fe2+/3+ oxidation states......................................................................................169 Figure 6.18 Integrated cyclic voltammogram of the voltage as a function of capacity of LiFeTiO4 probing the Fe2+/3+ oxidation state at the two cycling rates C/5 and C/2.5........................................................................................170 Figure 6.19 Capacity as a function of cycle number of the two cycling rates of C/5 (red) and C/2.5 (black), illustrating the cycling capacity and the capacity fade........................................................................................................171 Figure 6.20 Rietveld refined structure of LiFeTiO4 and of Li2FeTiO4 after electrochemical/chemical intercalation of a lithium ion.......................172 Figure 6.21 Mössbauer spectroscopy of LiFeTiO4 fit with one Fe3+ site.................173 xvi    Figure 6.22 Mössbauer spectroscopy of lithium deintercalated Li0.33FeTiO4 with the total of the fit two sites (red), LiFeTiO4 (green) and the amorphous iron oxide (blue)...........................................................................................174 Figure 6.23 The known structure NaV2O4 with the calcium ferrite structure as a potential starting point for probing new cathode materials...................177 Figure 6.24 Proposed structure of LiV2O4 resulting from the successful lithium for sodium ion exchange of the known compound NaV2O4.......................179 Figure 6.25 Proposed structure of Li2V2O4 after successful chemical lithium ion intercalation or electrochemical lithium intercalation during cyclic voltammetry..........................................................................................180 Figure 6.26 Proposed structure of V2O4 after successful chemical lithium ion deintercalation or electrochemical lithium deintercalation during cyclic voltammetry..........................................................................................181 Figure 7.1 Two-column space separated raw data saved with the ".exp" extension...............................................................................................185 Figure 7.2 Folder containing the required program files, raw data, and converted experimental file....................................................................................187 Figure 7.3 The in_kctl file with highlighted input files and parameters required for CONUSS analysis.................................................................................188 Figure 7.4 The in_kfit file with highlighted parameters for Mössbauer refinement.......................................................................................190 Figure 7.5 The in_kmco file with highlighted parameter files and input files.......191 Figure 7.6 The in_kref with highlighted material data input to be changed in each experiment.......................................................................................193 Figure 7.7 Thickness unit / micron parameter highlighted also in in_kref that will need to be initially refined to optimize data analysis............................194 Figure 7.8 The NAME_in file illustrating the known constant values of iron for Mössbauer spectroscopy.......................................................................196 Figure 7.9 The NAME_in file with highlighted number of sites of the Mössbauer atom.................................................................................................197 xvii    Figure 7.10 The NAME_in file with highlighted parameters that will be analyzed in Mössbauer refinement.........................................................................198 Figure 7.11 The NAME_in file illustrating electric field gradient as the target for distribution input file........................................................................199 Figure 7.12 The NAME_in file illustrating isomer shift as the target for distribution input file...........................................................................................200 Figure 7.13 The NAME_in file illustrating magnetic hyperfine field as the target for distribution input file.........................................................................201 Figure 7.14 The in_kfit illustrating the input of the % symbol, this opens the parameter for spectrum fitting............................................................204 Figure 715 Terminal screen shot illustrating input of the kctl command................205 Figure 7.16 Terminal screen shot illustrating the reults of the kctl command after fitting background and thickness.........................................................205 Figure 7.17 The main folder with out_kctl file highlighted.....................................206 Figure 7.18 The out_kctl file with the resulting values providing the best χ2 from the kctl command....................................................................................207 Figure 7.19 The in_kref with thickness unit / micron highlighted to be opened for background refinement........................................................................208 Figure 7.20 The resulting spectrum from the background refinement with unchanged parameters........................................................................................209 Figure 7.21 The NAME_in file illustraing the isomer shift, magnetic hyperfine field, and quadrupole splitting parameters open with the "%" for fitting.......210 Figure 7.22 Terminal illustrating kctl command with no convergence after 40 iterations.........................................................................................211 Figure 7.23 The out_kctl with the resulting values providing the best χ2 of the nonconverged data.................................................................................212 Figure 7.24 The NAME_in file illustrating the values from the kctl_out file input into the NAME_in file for further fitting..............................................214 xviii    Figure 7.25 The resulting spectrum of a non-converged kctl command with the isomer shift, magnetic hyperfine field, and quadrupole splitting parameters open.................................................................................215 Figure 7.26 The in_kmco with highlighted number of levels and samples per level.................................................................................................216 Figure 7.27 The kmco set up in the NAME_in file................................................218 Figure 7.28 The terminal illustration of output from the kmco command...............220 Figure 7.29 The highlighted output file as result from the kmco command............221 Figure 7.30 The best sampling as a result of the kmco command highlighted in the out_kmco file......................................................................................222 Figure 7.31 The best sampling values input into the NAME_in file for further refinement...........................................................................................223 Figure 7.32 Number of sites of the MB atom highlighted in the NAME_in file with 1 site..............................................................................................224 Figure 7.33 Number of sites of the MB atom highlighted in the NAME_in file with the changed number of sites from 1 to 2.............................................225 Figure 7.34 The @ &dist2 parameter highlighted after it was copied from the @&dist1 parameter and the 1 changed to 2..........................................226 Figure 7.35 The NAME_in file with defining MB site #2 highlighted after the 1 was changed to 2 and the entire section of defining MB site #1 was copied and pasted beneath the 0 0 0 at the bottom of defining MB site #1......227 Figure 7.36 The final paramater to be renamed, Make Gaussian 80 @&dist2, is highlighted after all of the parameters have been changed in the defining MB site #2.............................................................................................228 Figure 7.37 The weight of sublattice parameter for site 1 within the NAME_in file is highlighted illustrating its change from 1 to 0.5 due to the addition of the second site.............................................................................................229 Figure 7.38 The weight of sublattice parameter for site 2 within the NAME_in file is highlighted illustrating its change from 1 to 0.5 due to the addition of the second site.............................................................................................230 xix    Chapter 1 Introduction Synthesis and characterization of new materials have, to this point in time, paved the proverbial road to many new technologies that are used with little, if any thought, by the majority of people every day. Continued research in the synthesis of novel materials is essential to the continued growth of society. New materials have provided the world with mobile technologies that provide the use of cell phones, lap top computers, and tablets; in addition to rechargeable batteries. .he hardware found within each of these devices are examples of advancements in materials research. With every advancement that occurs, new problems also arise. An example of this is storage devices; computer flash drives continue to increase in memory but are essentially the same size they were when they came out in 2000. Eventually, there will be a limit to the memory these flash drives can hold. However, continued research on semiconductors provides the ability to continue this trend of increasing the memory while maintaining small size. This is just a simple example of a material that is used to make the lives of many easier. The topics that will be discussed within will focus on the use of relatively unconventional synthetic methods to synthesize materials in bulk and new materials to be used in rechargeable lithium ion batteries. Chapter 2 will discuss the instrumentation that was used in the characterization of the compounds within. It will include discussion on the background along with terminology that will be used throughout for each method as well as important features each method 1    provides to materials chemistry. Important theory will also be discussed to further understand some of the characterization methods. Chapter 3 will describe all of the synthetic methods used within. It will include background, a detailed description, the purpose of using the selected method over alternative techniques, and how each method was performed experimentally. Chapter 4 will explain the oxygen pressure regulation method used to prepare the divalent polymorphs α- and ß-NaFe2O3 that were previously only prepared as single crystals. The phase diagram of their synthesis will be described in detail along with the structural and physical properties of the two polymorphs. Chemical oxidation via sodium deintercalation to synthesize Na1-xFe2O3 to alter the average oxidation state of iron in the material along with metal doping into the iron position to synthesize NaFe2-xMxO3 will be discussed. Chapter 5 will discuss the synthesis and characterization of new tunnel polymorph LiFe1-xCoxO2 through the initial synthesis of ß-NaFe1-xCoxO2. Attempts to probe the Fe4+/3+ redox couple through electrochemical characterization of the undoped sample along with the electrochemical characterization of the Co doped sample will be described. Electrochemical characterization of the Fe3+/2+ redox couple will also be shown with discussion of the potential for two-electron transfer. Chapter 6 will describe the synthesis of the new tunnel polymorph LiFeTiO4 with the calcium ferrite structure for the purpose of a two-electron transfer rechargeable lithium ion battery. Synthesis of the precursor phase NaFeTiO4, ion exchange to form the new compound LiFeTiO4, structural and electrochemical characterization of LiFeTiO4, and 2    attempts to probe the Fe4+/3+ redox couple will also be discussed. Further characterization of LiFeTiO4 by chemical reduction of Fe3+ to Fe2+ using n-butyllithium was determined structurally to be Li2FeTiO4.  3    Chapter 2 Instrumentation 2.1 Powder X-ray Diffraction Powder X-ray Diffraction (PXD) uses X-rays to characterize powder materials for phase identification and structural characterization. In order to acquire an accurate PXD, all possible orientations of the crystalline solid must be represented equally. This is generally achieved by grinding the sample in question in a mortar and pestle or ball mill to reduce the powder size and therefore providing an increased number of representative crystals of all orientations. The detected orientational averaging of the bulk composition projects a single dimension spectrum that, in contrast to single crystal diffraction, is a three dimensional reciprocal space projection. This will be discussed later in section 2.7. Due to the resulting spectrum being the average of the orientation of the bulk composition, it is imperative that the sample represent all possible orientations equally. If it is not representative of all orientations equally, the PXD spectrum can be altered due to what is called preferred orientation. Preferred orientation is when the distribution of the powder is non-random. This can result in PXD patterns missing peaks that would otherwise have been present, therefore leading to difficulties in the data analysis such as peak indexing. Another way to prevent preferred orientation on top of homogenous grinding is sample rotation. This provides multiple different particles in different orientations while the scan is occurring. A powder X-ray diffractometer is composed of three basic parts: an X-ray tube, a sample holder, and an X-ray detector. X-rays are generated in a cathode ray tube by heating the 4    filament to produce electrons. These electrons are then accelerated toward the sample by applying a voltage. The resulting electrons that are diffracted by the material at different angles are detected producing a spectrum that corresponds to the characteristics of the sample material. X-rays produced are of a specific wavelength that is characteristic to the X-ray tube source element (Cu, Mo, Fe, Co). Herein all reported PXD patterns are the result of Cu radiation with a wavelength of 1.5418Å. X-rays with wavelengths of such magnitudes are required in PXD because they are comparable to the spacing between planes in the crystal lattice. Peaks in a PXD spectra result from constructive interference from waves, with the angle (θ) of incident radiation emerging from planes and with spacing (d) from within the crystal lattice. This is true for constructive interference when the distance is integral multiples of the wavelength λ as described by Braggs' law in Equation 2.1. nλ=2dsinθ Equation 2.1 Determining parameters such as d-spacing, cell parameters, or Miller indices of simple cubic structures can be relatively easy with an adequate PXD spectrum, but can become much more difficult with an increase in complexity of space groups. For instance, determining the d-spacing requires solving Braggs law Equation 2.1. Further calculations to determine a compound's cell parameters (a, b, and c) or Miller indices (h, k, and l) would require Equation 2.2. 1 d2hkl = h2 a2 + k2 b 2 + l2 c2 Equation 2.2 For compounds that contain a simple cubic cell space group where a=b=c, Equation 2.2 can be simplified into Equation 2.3. 5    1 d2hkl = h2 +k2 +l2 a2 Equation 2.3 The solution of Equations 2.2 or 2.3 can only be determined if the unknown variables are decreased to one. This is accomplished with the knowledge of destructive interference leading to systematic absences of reflections unique to the compound's space group. Continuing with the example of a simple cubic cell, the Miller indices, M, can be determined from Equations 2.4 and 2.5. M2 =h2 +k2 +l2 Equation 2.4 M= h2 +k2 +l2 Equation 2.5 The resulting integer value of M2 can then be compared to an appropriate table of known reflections of the compound's space group, which in turn provides three values for the sum of the h, k, and l. These indices, i.e. [220], only describe the mathematical result of M. More probable than not, this is the incorrect order of which the indexed values are for a specific peak i.e. [202] or [022]. This does though, correlate to the multiplicity, mhkl, which is indicative to the number of planes that have the same d-spacing. Due to the dspacing being inversely proportional to h2+k2+l2, as seen in Equation 2.3, all combinations of h, k, and l with the same values contribute to the observed peak intensity (Equation 2.6). Ihkl α mhkl F2hkl 1+cos2 θ sin2 θcosθ hkl Equation 2.6 Continuing this example with the hkl value of 220, because M2 is a result of the square root of h2, k2, and l2 (Equation 2.4), there are 12 possible ways to represent this value ([220], [202], [022], [-220], [-2-20]...), therefore 12 is the multiplicity. Though each 6    Miller indices contributes to each peak's intensity of which it is a subset of its multiplicity, each Miller indice can be its own peak. This becomes more complicated when the resulting M2 value is 9, therefore the resulting h, k, and l values could be 300 or 221, again most likely not in that specific order. As seen in Equation 2.6, the peak intensity, Ihkl, is proportional to the structure factor, Fhkl, which is dependent upon how the material scatters incident radiation, calculated using Equation 2.8. In order to do so, the scattering factor, f, must first be determined using the electron density of an atom, ρ(r), using Equation 2.7. f=4π ∞ sin ((4π/λ)sinθ)r 2 ρ (r) r 0 (4π/λ)sinθ)r Fhkl = ∑j fj e dr 2πi(hxj +kyj +lzj ) Equation 2.7 Equation 2.8 For simple cubic cells containing two, or perhaps three elements, all angles of the unit cell are 90° and all cell parameters are equal. These calculations can be easily performed using common software spreadsheets. Increasing the complexity of the space group along with increased variables, such as multiple atoms, different cell parameters, or different unit cell angles all add to the overall complexity of the calculations; therefore, requiring the use of specialized software such as TOPAS or GSAS for the compound's cell parameters, indexing, and crystal structure Rietveld refinement. Further analysis of PXDs and the use of specialized programs allow for determination of the crystal structure using Rietveld refinements (TOPAS or GSAS), particle size determination can be performed using a program such as PM2K, or long hand calculation using the Scherrer Equation, Equation 2.9. 7    τ= Κλ βcosθ Equation 2.9 This equation is used, initially solving for Κ, the shape factor, by using a standard of which τ, the mean size of the ordered crystals is known. β, the line broadening FWHM, θ, the Bragg angle, and λ, the X-ray wavelength are known or determined values from the PXD spectra. LaB6 powder or Al2O3, corundum, with known crystallite size are common standards. The Κ value is then input into the Sherrer formula with the spectrum of which crystallite size is of question and τ can be solved. This formula is limited to crystallites in size up to approximately 0.1µm, which are generally much smaller than compounds synthesized at high temperatures (greater than 500 °C). Rietveld refinement method uses a least squares approach of refinement to derive a theoretical line profile calculated based on a known crystal structure model. This method is performed until the calculated line overlaps the measured reflections in a PXD spectrum. This method uses height, width, and peak positions in a spectrum to determine the structural characteristics in powder materials. Powder samples are hand ground until homogenous and fine, poured onto a zero background silicon plate, and flattened with a glass slide. The sample is placed in the powder diffractometer and run from 10 °-65 ° 2θ with a step size of a 0.02 °. The resulting diffraction pattern is matched to corresponding known diffraction patterns from the International Centre for Diffraction Data database. If the resulting pattern is determined to be unknown, further analysis using the Rietveld refinement programs TOPAS or General Structure Analysis System (GSAS) is performed [1]. 8    2.2 Inductively Coupled Plasma Atomic Emission Spectroscopy Inductively Coupled Plasma Atomic Emission Spectroscopy (ICP-AES) furthermore referred to as ICP, is an emission spectroscopy that uses a plasma torch created by argon gas and a radio frequency output coil to determine metal concentrations within a sample. The emission spectrum is the result of the excited atoms or ions emitting electromagnetic radiation with characteristic wavelengths indicative of its corresponding element. The intensity of the emission peak in the spectra corresponds to the concentration of that element within the sample. ICP can be used for the analysis of up to 70 elements at a time while being accurate to elemental analysis in the parts per billion regimes. Limitations to ICP can include inter-elemental interfaces resulting in inaccurate data, and due to the nature of the analysis, loss of sample. The final sample must be dissolved in a 2-5 % acidic solution, preferably a HNO3 (Nitric Acid) solution. Therefore, depending on the sample, direct addition into concentrated HNO3 will result in the dissolution of the metal compound. Due to the stability of some metal oxide compounds, concentrated HNO3 will not dissolve them. Therefore, a concentrated solution of aqua regia is required. This is a mixture of concentrated HCl (hydrochloric acid) and concentrated HNO3 (all within solutions of aqua regia are composed of a 1:1 mixture of concentrated HNO3:HCl). The resulting solution of dissolved metal oxides is then diluted to the 2-5 % acidic solution for ICP analysis. 9    Sample calibration must be performed on every element to be analyzed. This is achieved by the appropriated dilution into 2 % HNO3 of a purchased 1000 parts per million (ppm) standards, using a micropipette to construct a calibration curve of which the analyzed sample concentrations will fall within. Qualitative ICP is performed by the analysis of the molar ratio of multiple metals within the sample, while quantitative ICP requires the accurate mass of the analyzed sample and accurate volume the dissolved sample is diluted to. Qualitative ICP is useful for analysis when determining only the metal-tometal ratios within a sample. Quantitative ICP provides increased information due to the careful preparation of the metal oxide compound. This allows for the assumption that the remaining unobservable concentration in ppm, via ICP, are the non-metals within the compound; therefore, providing oxygen content of the material without the use of titration. 2.3 Mössbauer Spectroscopy Mössbauer spectroscopy is used to probe the oxidation states and local environments of metals. It is most commonly used for iron in mineralogy or geophysics. It is significant in solid state chemistry for investigating the oxidation state of iron (Fe0, Fe2+, Fe3+ or Fe4+) and exploring the local environment of iron's coordination polyhedron (trigonal, tetrahedral, or octahedral). Mössbauer spectroscopy is an extremely useful characterization tool for explaining observed properties of different Mössbauer active compounds. 10    Mössbauer spectroscopy uses 57Fe, the decay product of 57Co, and an unstable isotope of iron, which in turn emits different types of energy including γ-rays. These energies can result in a number of different possible processes when they collide with the sample (the absorber), such as the emission of core shell electrons, which are not of interest in Mössbauer spectroscopy. If the emitted γ-rays were to collide with the nucleus of the absorber iron, and have the correct energy (14.4 keV), it would move the nucleus of iron. The nucleus must then recoil with equal and opposite energy, resulting in an emitted γray with the as mentioned 14.4 keV recoil energy. This can be seen in Figure 2.1. 11    Figure 2.1 Mössbauer schematic illustrating the decomposition of 57 Co into 57 Fe resulting in the emission of energy into an absorber, in this case an iron source, at which point the absorber emits non-resonant, resonant, conversion, Auger electrons, and x-rays with other various energies. The units of energy for Mössbauer spectroscopy are in mm/sec due to the set up of the spectrometer. The source for conventional spectrometers is mounted on a drive that oscillates the source toward and away from the absorber, therefore increasing the energy of the emitted photons. For sources such as a synchrotron, the absorber is oscillated with respect to the source. 12    In order to understand the resulting spectrum of Mössbauer spectroscopy, three important parameters must be discussed: the isomer shift, quadrupole splitting, and the magnetic hyperfine splitting. The combination of these three parameters is used to determine the oxidation state along with the local iron polyhedron environment. Figure 2.2 illustrates the observed results of isomer shift, quadrupole splitting, and magnetic hyperfine splitting in a Mössbauer spectrum. Isomer shift is the comparison of two different nuclear isomeric states in two different physical or chemical environments due to the combined effect of the recoil-free transition between the two states. This difference is observed on the Mössbauer spectrum by the shift from the 0 mm/sec point. The general trend when considering two iron atoms within the same local environment is: as the isomer shift increases in positive value the oxidation state of iron is less. Quadrupole splitting is the result of the nuclei having a non-radially-symmetric shape and is in the presences of an external electric field gradient. This results in the splitting of what would normally be a singlet peak in the Mössbauer spectrum into a doublet. The separation between the doublet peaks can be measured to provide further information on the chemical environment of the nuclei. Magnetic hyperfine splitting (also referred to as electronic field gradient or EFG), is the result of the energy of the nuclear magnetic dipole moment of the magnetic field 13    generated by the electrons. This is observed in a Mössbauer spectrum by the further splitting of a singlet or doublet into a sextet from the resulting magnetic field. Figure 2.2 Schematic of Mössbauer isomer shift, quadrupole splitting, and hyperfine splitting illustrating the effect each has upon the Mössbauer spectrum [2,3]. The following procedure takes place inside an inert, dry nitrogen glove box to prevent any reaction that may occur between the sample and oxygen or water from air. The prepared powder sample of interest is spread in a homogenous fashion on the sticky side of a piece of Kapton tape, with a concentration of at least 10 mg of iron per square centimeter of tape. A second piece of Kapton tape is then placed sticky side to sample to "sandwich" the sample between the two pieces of tape. The now sandwiched tape sample 14    is cut into a square with sides of approximate length of one centimeter. Next, all sides of the square are taped to seal the edges from ambient conditions when placed in the Mössbauer spectrometer. The sample is placed inside a small vial and parafilm wrapped, then placed in a larger vial and parafilm wrapped. The large vial is finally placed inside a zip lock bag for transportation to the measuring facility. 2.4 Electrochemistry Electrochemical characterization of potential cathode materials were performed in an argon filled glove box due to the reactivity of the anode (lithium metal) with nitrogen to form Li3N. The cells were assembled in a split test cell. The cells used a lithium disc as an anode and counter electrode. A 1 M solution of LiPF6 in ethylene carbonate (EC), dimethyl carbonate (DMC), and diethyl carbonate (DEC) (EC/DMC/DEC = 1:1:1 by volume) was used as the electrolyte. The anode/cathode separator was a 25 pm Microporous Monolayer Membrane. The cathode was a 19 mm disc of slurry prepared by ball milling 75% by weight active material, 10 % by weight PTFE, and 15 % by weight Super C65 carbon in 1-methyl-2-pyrrolidone. A 30 micrometers thick slurry layer was deposited on aluminum foil using a doctor-blade method. The layer was then vacuum dried and pressed for increased particle-particle connectivity. Cyclic voltammetry were initially run from 1.5 V to 5.0 V to determine the oxidation and reduction potentials of the active material. Further analysis of the active material was then performed with a narrower window and slower scan rate. 15    2.5 Thermogravimetric Analysis and Differential Scanning Calorimetry Thermogravimetric analysis (TGA) is the analysis of materials in which the physical properties are measures as a function of increasing temperature at a constant rate in different gases. Differential scanning calorimetry (DSC) is the analysis for which the difference for heat required to increase the temperature of a sample function of increasing temperature at a constant rate. When coupled, TGA-DSC provides information in different gasses about mass loss or gain in oxidizing, reducing, or inert atmospheres. This information can explain exothermic or endothermic reactions, the temperature at which point a phase transition occurs, or basic information such as the temperature required to dehydrate a sample without decomposition. 2.6 Scanning Electron Microscopy Scanning electron microscopy (SEM) uses high-speed electrons to study surface morphology and particle size in the sub micrometer regime as low as about 10 nm into the micro-regime. Morphology of the micro sized particle can provide insight into characteristics such as crystal growth along a particular axis. This may provide insight into possible observations such as preferred orientation within the PXD or intercalation or deintercalation of the electrochemical results. 16    Elemental analysis within both SEM and TEM is performed using energy measurements of X-rays produced by elements in the sample from the electron beam of the instrument called energy-dispersive spectroscopy (EDS). This is accomplished by evaluating the critical excitation energy difference between different elements. The differences in the critical excitation energy within the analytical spatial resolution of the whole sample are measured by the loss of energy over the distance the X-rays travel into the sample. Comparing two elements of different critical excitation energies, the greater distance the incident electron travels, the more energy is lost due to scattering. Therefore, an element with a low excitation energy would produce X-rays from a greater distance within the sample as compared to an element with a greater excitation energy. The X-rays from within the sample lead to the production of characteristic X-rays of the specific elements the sample is composed of. Each element has a fingerprint per say, a unique energy created by the filling of an inner orbital shell electron by an electron from a shell of higher energy. This specific energy difference is emitted as detectable X-rays for most elements. An element often produces many types of X-rays because multiple beam electrons are striking the sample. Each electron may differ in the interaction with the sample leading to the emission of different X-ray wavelengths produced from the relaxation of different sub-shell electrons. These different electron relaxations are named based on which shell the vacancy was created in (K, L, M, N), and the number of shells by which the electron jumped to fill the vacancy (α, ß, γ ...). A downfall of EDS is the inability to detect lighter elements with lower critical excitation energies, due to the beryllium window absorbing the excitation energies of about 1 keV of energy. Peak 17    overlap in the spectrum is another disadvantage of EDS; therefore, analysis of compounds with multiple elements may cause peak overlap resulting in difficulties for quantitative elemental analysis. EDS has several advantages over other elemental determining methods for SEM, such as wavelength-dispersive spectroscopy (WDS). WDS does not exhibit peak overlap as does EDS, but EDS is generally the preferred elemental analysis spectroscopy in SEM and TEM due to its ability to analyze more than two elements at a time unlike WDS. Analysis software can provide relatively accurate quantitative analysis even in the instance of peak overlap in an EDS spectrum. Backscattering electron imagining (BSE) uses a different source of excitation radiation. This radiation is created by an elastic interaction of beam electrons with the nuclei of the atom. Heavy atoms have a larger atomic number, therefore, scatter stronger than lighter elements with smaller atomic numbers. Due to this fact, backscattering images can provide an accurate elemental composition and an image illustrating the distribution of elements within the sample. This is increasingly important when determining the metal ratio and composition of new compounds or determining phase composition of impurities or element deficient compounds [4–7]. Powder samples are spread on a sticky carbon tab on an aluminum stub with the excess powder removed using compressed air. This is performed to prevent any damage or contamination by loose particles ejecting from the stub into the detectors, vacuum system, or any other parts of the SEM. Depending on the electrical conductivity of the sample, it can then be left alone or coated with carbon, gold, or osmium with appropriate 18    sputter coaters to increase the sample's electrical conductivity for better resolution. The sample is then inserted into the instrument for analysis. 2.7 Transmission Electron Microscopy Transmission electron microscopy (TEM) uses high-speed electrons to study particle sizes or layer thickness of nano-materials or thin films. TEM was employed on prepared nano-crystalline samples to determine morphology, particle size, and potential core shell structures. TEM was used rather than SEM due to the magnification ability of over 500,000 times providing resolution of particles with sizes as low as 0.2 nm. TEM is useful to determine particle size and morphology of nano-crystals, however, high resolution TEM (HRTEM) provides the ability to probe further into the crystalline nanoparticles. HRTEM permits an increased accuracy in particle size measurements and greater detailed images of the sample. HRTEM also grants the ability to observe the lattice fringes of the nano-crystalline structure. The lattice fringe is the observed plane within the nano-crystal that correlates to the Miller indice observed in the PXD. Therefore measuring the distance between two lines of the observed lattice fringe provides the ability to determine which plane the nano-crystal is orientated. Furthermore, single crystal diffraction of nano-crystals is performed using Fourier transforms of HRTEM resulting in electron diffraction spots. 19    In order to discuss single crystal diffraction and better understand some aspects of PXD (as mentioned in 2.1), the topic of reciprocal lattices must be discussed. First, a few mathematical definitions must be defined and worked through. Initially considering planewave front vectors in terms of k, where the initial wavefront, kI, at a specific wavelength, λ, is scattered by a plane of atoms resulting in a diffracted wavefront, kD, that scatters at an angle of 2θ. With this, we can determine the difference in kI and kD as seen in Equation 2.9, where K is the change in the vector due to diffraction as illustrated in Figure 2.3. Figure 2.3 The relationship between kI, kD, K, θ, and λ. The initial wavefront (green), the top of the triangular shape is the point at which the wavefront is diffracted by the plane of atoms, resulting in the diffracted wavefront (blue), and the change in the vector due to diffraction (black) is K. K=kD -kI 20    Equation 2.10 If the energy of the electron is unchanged during the diffraction, therefore is elastic so we can relate these wavefront vectors as seen in Equation 2.11. 1 |kI |=|kD |= =|K| λ Equation 2.11 Furthermore using trigonometry to write an expression in terms of θ, sin θ= |K|/2 Equation 2.12 |kI | then algebraically rearranging, Equation 2.13 is observed. |K|= 2 sin θ Equation 2.13 λ In relating Equation 2.13 to Braggs' law, Equation 2.1, these conditions must be met: the angle θ must equal the Bragg angle and the integer value of n is one. The equations can be combined into Equation 2.14, where the vector K at the Bragg angle is KB. |KB |= 1 d Equation 2.14 Under the conditions where vector K is at the Bragg angle, the magnitude of vector K has a special value, therefore is defined by g. Equation 2.15 is also very important because it represents the Laue conditions for constructive interference. The Laue diffraction will later be discussed. KB =g Equation 2.15 Now that it has been established that K is reciprocally related to d, and K at the Bragg angle is an important value, of which further discussion will come later, it is important that real space lattice and reciprocal space lattice vectors be discussed. Here a lattice in real space, rn, will be defined as follows; rn =n1 a+n2 b+n3 c 21    Equation 2.16 where the vectors a, b, and c are unit-cell translations in real space and n1, n2, and n3 are integers. Similarly the reciprocal-lattice vector, r*, can be defined as: rn =m1 a* +m2 b* +m3 c* Equation 2.17 where in this case the vectors a*, b*, and c* are unit-cell translations in reciprocal space and m1, m2, and m3, are integers. The direction of the vectors can be seen using the dot product of the real space vectors with the reciprocal space vectors as seen in Equation 2.16, where a* is normal to both vectors b and c. a* ·b=a* ·c=b* ·c=b* ·a=c* ·a=c* ·b=0 Equation 2.18 The length of these vectors can be related as follows: a* ·a=1;b* ·b=1;c* ·c=1 Equation 2.19 Here it is seen that the dot product of the reciprocal unit-cell length on the real space unitcell length is defined to be one, consequently, providing the scale or dimension of the reciprocal lattice. Furthermore, this relationship provides insight into the dimensionality of the reciprocal lattice, in that if a, b, and c are all large the corresponding reciprocal lattice vectors will be small. Continuing discussion on the concept of g, the K vector at the Bragg angle, it can be established that any vector in reciprocal space can be defined as seen in Equation 2.20. K=ξa* +ηb* +ζc* Here , , and Equation 2.20 are three numbers, which do not have to be integers. The ghkl vector, similarly written to the K vector, is an important vector in reciprocal lattice vectors because it provides the definition of h, k, and l, which are integer values that together define a plane (hkl). Due to h, k, and l being in reciprocal space, the definition of the 22    plane (hkl) is that it cuts the a, b, and c axes at 1/h, 1/k, and 1/l respectively. With this notation the vector BC, can also be written as c/l-b/k. ghkl =ha* +kb* +lc* Equation 2.21 Acquiring the dot product of g and rn by using Equations 2.16 and 2.20, an integer value, N, is always achieved, therefore Equation 2.21 describes similarly, that in order to have Laue (or Bragg) diffraction, certain conditions must be satisfied. In other words, from Equation 2.20, , , and must be integer values and the integer values are h, k, and l in order for Laue diffraction to occur. K·rn =N Equation 2.22 Setting rn to three unit vectors, establishes the following relationships, which are the core principle of Laue diffraction: K·a=h Equation 2.23 K·b=k Equation 2.24 K·c=l Equation 2.25 Knowledge of reciprocal space is required to understand Ewald's sphere, which is observed in single crystal diffraction in TEM. When the Bragg (or Laue) diffraction condition is satisfied, the scattering vector is equal to the reciprocal lattice vector, as previously described. Therefore, diffraction will occur only for the reciprocal lattice points that lie on the surface of the Ewald sphere. If the diffraction pattern is aligned perfectly, at which the Bragg condition is satisfied, where no spots will be observed, tilting of the sample will result in the appearance of diffraction spots. Continued tilting of 23    the sample down different planes of the sample will result in the appearance and disappearance of diffraction spots that cut through different Laue zones. [4–7] Powder samples are dispersed in a solvent of choice in a test tube. The solvents must be completely evaporated prior to analysis in the TEM, therefore methanol or ethanol is chosen due to their volatilities. The dispersed sample is then sonicated for 15 minutes to reduce agglomerations. A drop of the resulting solution is then placed on a TEM grid and allowed to dry and be inserted into the instrument for analysis. 24    REFERENCES 25    REFERENCES [1] V.K. Pecharsky, P.Y. Zavalij, Fundamentals of Powder Diffraction and Structural Characterization of Materials, Second Edi, Springer US, Boston, MA, 2009. [2] K. J. Kuhn, K. Th. uller, N.J. A. Ruckenstein, U. F. Steiner, G. J. Tumper, K. P. Wölfle, Nuclear Condensed Matter Physics with Synchrotron Radiation Basic Principles, Methodology and Application, Spring Berlin Heidelberg, Berlin, Heidelberg, 2004. [3] P. Gütlich, E. Bill, A.X. Trautwein, Mössbauer Spectroscopy and Transition Metal Chemistry, Springer Berlin Heidelberg, Berlin, Heidelberg, 2011. [4] D.B. Williams, C.B. Carter, Transmission Electron Microscopy A Textbook for Materials Carter, Springer US, New York, n.d. [5] S. Amelinckx, D. Van Dyck, J. Van Landuyt, G. Van Tendeloo, Electron Microscopy Principles and Fundamentals, VCH Verlagsgesellschaft mbH, Weinheim, 2008. [6] P. Echlin, Handbook of Sample Preparation for Scanning Electron Microscopy and X-Ray Microanalysis, Springer US, Boston, MA, 2009. [7] L. Reimer, H. Kohl, Transmission Electron Microscopy, Fifth Edit, Springer US, New York, 2008. 26    Chapter 3 Synthetic Methods 3.1 Pechini (Sol-Gel) The Pechini method, also referred to as the polymerisable complex method liquid mix technique, or sol-gel, was first proposed in the late 1960's as a technique to deposit different elements in the production of capacitors. It was later determined useful as a synthetic technique for the synthesis of multicomponent metal containing materials. The method relies on the mixing of ions in solution then controlled temperature increase to transform the solution into a gelatin-like polymer. Increasing the temperature furthermore removes the matrix resulting in homogenously mixed metal oxide precursor. The sol-gel process involves the addition of stoichiometric ratios of metal salts or metal oxides with citric acid, ethylene glycol, nitric acid, and water. The citric acid is added to form citric complexes that upon heating to above 100 °C form the polymer citrate gel with the ethylene glycol. The addition of nitric acid is to increase the dissolution of the metal oxides within the solution. Continued heating above 400 °C removes the matrix leading to the formation of a uniform metal oxide and/or metal carbonate mixture to be further annealed under appropriate conditions to acquire the desired material. This method has been widely used for the synthesis of materials with many different properties such as superconductivity, dielectrics, catalyst, etc. Nanocrystalline powders have also been synthesized using the Pechini method for different purposes such as dyesensitized solar cells (DSSCs) and water splitting reactions for hydrogen production. [1– 6] 27    3.2 Solid-state (Ceramic) Solid-state reaction, also referred to as ceramic method, is generally the standard technique of synthesis in material sciences. Limitations are due to slow diffusion in solids as compared to liquid/liquid chemistry. High temperatures are required to increase the rate of diffusion within the reaction. The ceramic method is generally the easiest way to synthesize stable, solid-state materials directly from the metals, metal oxides, or metal oxoanion precursors. To increase the rate of diffusion, it is preferable the starting materials are homogenously mixed prior to sintering. This provides a decreased path length of atomic diffusion for nucleation sites of the desired material to form. This type of method can result in incomplete reactions if the starting reagents are inhomogeneously mixed. After initial sintering, the material is generally reground and reannealed. This allows the nucleation sites to react with unreacted starting materials resulting in increased purity of the final product. This method has been given the nickname "the heat and beat method" because of the firing and grinding repetition that is sometimes needed to acquire a pure product. High temperature synthesis can also lead to the loss of volatile metals such as lithium or sodium; therefore, stoichiometric compensation is taken into consideration in the initial reactant addition. [7] 3.3 Reduction Reduction of complex metal oxides, if carefully performed, can result in the formation of meta-stable complex metal oxides that may contain properties otherwise not seen in the 28    non-reduced phase. There are multiple reduction types in the material science synthesis like solvothermal reduction, which involves the use of metal hydrides such as sodium hydride (NaH) and a solvent inside a high-pressure autoclave with the temperature of the reaction elevated to a desired temperature. An approach resembling a solid-state synthesis reaction involves the mixing and grinding of metal hydrides with the complex metal oxide. The mixture is then heated to evolve hydrogen gas that can be involved in the reduction, or in the formation of the hydrogen gas. The resulting metal acts as a reducing agent toward the complex metal oxide. A third reduction type includes the flow of hydrogen gas at an elevated temperature to form water and the reduced complex metal oxide. All reduction of complex metal oxides discussed within will include only the third reduction process. Reduction using 100% hydrogen gas was performed with extensive care due to the high flammability of hydrogen and the fact that hydrogen gas is the only gas that, upon expansion, is exothermic. For these reasons, all reactions with hydrogen gas were performed in a tube furnace using a fused silica tube rather than the typical alumina tube, due to the porosity of the alumna tube and the elevated temperature at which the reaction takes place. All reactions using hydrogen gas also took place inside a fume hood rather than out in the open for extra safety. The hydrogen gas was flowed at a rate of approximately 1 ml per second through the reaction tube into an oil bubbler. Final compounds of reduction are moisture sensitive; therefore, the hydrogen gas was also flowed over glass wool coated with P2O5 to ensure the gas was dry prior to entering the reaction tube. P2O5 was also placed at the end of the furnace to collect water that was 29    evolved from the reaction to ensure diffusion of water back into the reaction tube did not occur. The sample was placed in the tube as a non-pelleted powder and spread in an alumina boat in a thin layer to increase surface exposure to the reactive gas. Intermittent grindings inside a glove box were performed for homogenous sample exposure to the reactive gas for optimal purity of the reduced phase. 3.4 Partial Oxygen Pressure Regulation The ceramic method provides the ability to control the atmosphere within reaction by using a reactive gas that provides different conditions to acquire desired products. Oxidizing atmospheres such as oxygen, inert atmospheres such as nitrogen or argon gas, or reducing atmospheres such as hydrogen gas, are all common gasses used in solid-state reactions providing versatility in the synthesis process of new compounds. In the synthesis of compounds, generally only one valence electronic state is available within the transition metal of the complex compound. Therefore, synthesis of bulk compounds with multiple oxidation states for the transition metals can only be achieved by increasing the temperature of the reaction; thus, decreasing the oxygen reactivity and forming oxygen deficient phases. In these circumstances, though common, the transition metal is only partially divalent due to the oxygen deficiencies within the compound only resulting in a small percentage while still maintaining the overall structure. Single crystals on the other hand, can be synthesized regularly with multiple oxidation states of the transition metal. This usually occurs within vacuum-sealed ampoules composed of different transition metals such as iron, copper, or platinum, or the ampoule can be non-metallic 30    such as glass or quartz. These different types of ampoules, though under vacuum, can provide different types of oxidizing environments, like the metal's reduction potentials in electrochemistry. If under the correct heating and cooling conditions, this can result in single crystal formation of compounds that can have metals with multiple oxidation states. As mentioned, synthesizing bulk materials with transition metals in multiple oxidation states of significant magnitude cannot be performed using normal reactive gasses such as oxygen or hydrogen. Therefore, the use of an oxygen buffer system to control the oxygen partial pressure within a reaction vessel to synthesize complex compounds, to this point, is the only known reaction system to be successful. Controlled oxygen partial pressure reactions were performed using a two zone furnace; one zone to control the temperature of a metal/metal oxide buffer such as Cu/Cu2O (copper/copper(I)oxide) or Ni/NiO (nickel/nickel(II) oxide) and the other zone to control the temperature of the sample. Nitrogen gas was flowed over glass wool covered with P2O5 then flowed through the reaction tube with P2O5 at both ends to collect any water that may have been released due to the increase in temperature during the reaction and absorbed by the powders. The temperature of the sample for all reactions is kept constant, while the temperature of the metal/metal oxide buffer is changed to appropriate temperatures for the desired partial oxygen pressure characteristic of the specific buffer. The reaction vessel is purged for at least one hour with nitrogen gas, the flow is then stopped, and valve 1 as seen in Figure 4.2, is closed. The two zones of the furnace are then initiated simultaneously with an identical ramp time, therefore, each zone reaches the desired temperatures at the same time. At the time the temperatures reach their 31    desired values, valve 2 is closed (Figure 4.2). The reaction is allowed to proceed for a minimum of 48 hours, but can go longer due to the equilibrium state at which the system is in. The reaction vessel is quenched to room temperature to prevent the buffer from decreasing the partial oxygen pressure in the reaction vessel for a prolonged period at an increased temperature therefore altering the final state the product. The sample is then placed in a vial and taken into an inert glove box for further analysis. 3.5 Moisture and Oxygen Sensitive Techniques Synthesis of metal-stable, oxygen sensitive, or moisture sensitive compounds requires an increased measure of care to prevent decomposition, uncontrolled, or unknown reactions from occurring that alter the properties of the final product. Properties of moisture sensitive samples can be altered. This can be seen in the example of Na0.35CoO2·yH2O, where in the case of y=1.3, the product is superconducting whereas the anhydrous product or the further water-intercalated products do not exhibit such properties. The discovery of the superconducting material was somewhat serendipitous due to the high relative humidity in the area where the magnetic properties were measured, which resulted in the intercalation of water molecules into the material. The original material did not contain the intercalated water; thus, the exposure to the humid atmosphere resulted in the final product's properties. Further structure refinement provided insight on the water intercalation. Without the further analysis that was performed on the material, incorrect data could have easily been presented, if the researcher assumed that, in fact, the material they sent out for measurements maintained its original composition. To 32    prevent any unwanted water intercalation, oxidation, or decomposition, samples can be prepared in inert nitrogen or argon glove boxes. All samples that are prepared are initially assumed to be moisture sensitive until further analysis determines otherwise. This assumption is made for all sodium and lithium containing compounds due to the general hygroscopic nature of most compounds that contain these elements. This assumption is also made for meta-stable compounds synthesized through solid-state reduction. To determine whether the sample is in fact moisture sensitive, the sample can be washed in water then analyzed using PXD and compared with previous PXD of the non-water washed phase. Precautions for sample exposure to ambient conditions are minimized as much as possible. In the process of solid-state reduction of Perovskite structure to infinite-layer structure, the fused silica tube was made to fit within the antechamber of the nitrogen glove box, therefore, the entire reaction vessel can be inserted into the glove box without any exposure. In the case of α- and ß- NaFe2O3, due to the increased size of the reaction vessel to fit in a two-zone furnace, the transfer of the entire vessel cannot occur. The sample is removed from the reaction vessel, placed in a vial, immediately placed inside the glove box antechamber, and evacuated. Removing the sample from the reaction vessel is performed under a nitrogen gas flow creating an inert "blanket" over the sample when possible. Materials that are synthesized from moisture sensitive precursors, such as NaFe2O3 with the moisture sensitive precursor α-NaFeO2, or the preparation of α-NaFeO2 which is synthesized from the hygroscopic compound NaOH and non-hygroscopic compound 33    Fe2O3, are all prepared in a glove box. The compounds are ground, mixed, and pelleted, if required, inside the glove box then transferred to an appropriate reaction vessel. 3.6 Chemical oxidation and reduction Studying the properties of new materials can sometimes lead to further questions, such as the effects of lithium intercalation and deintercalation on the overall structural integrity of the new material. To answer such a question, a few different processes can occur; either in situ PXD and electrochemical characterization, or oxidation via chemical lithium intercalation and reduction via chemical lithium deintercalation. In situ PXD and electrochemical characterization provides insight on the structural properties directly as the cyclic voltammogram is produced. The downfall is the expensive set up required to do such measurements. Generally, these measurements are performed at national laboratories. Oxidation and reduction via chemical lithium intercalation and deintercalation can provide the compound with deficient or inserted lithium from or into the structure respectively; therefore allowing PXD to be performed on the samples providing information on the structure and the effect of the inserted or removed lithium. Chemical lithium intercalation was performed in a nitrogen glove box using 1 M nbutyllithium in hexanes. The ground powder sample was dispersed by stirring in pentane and the n-butyllithium solution was added drop wise until the correct volume for the required molar ratio was achieved. The mixture was stirred for at least 48 hours then 34    washed with pentane and filtered. The resulting powder was then stored in an inert glove box for further analysis. Chemical lithium deintercalation was performed in a nitrogen glove bag due to the oxidizing nature of the compounds used and the metal surroundings inside a glove box. Bromine (Br2), iodine (I2), nitrosyl tetrafluorborate (NOBF4), or nitrosonium tetrafluoroborate (NO2BF4) are all oxidizing reagents with different reduction potentials providing the ability to deintercalate different amounts of lithium from a compound. The sample is dispersed and stirred in acetonitrile (CH3CN) if the reaction was with NOBF4 or NO2BF4, or pentane if the reaction was with bromine or iodine. The slow addition of the dilute oxidizing agent was then performed until the required volume to reach the desired molar ratio was achieved. The mixture was then stirred for at least 48 hours or until the reaction had come to completion by visual indication. Oxidation using bromine and iodine, upon completion, would result in the color change from purple to clear indicating the bromine or iodine had reacted to produce the bromide or iodide salt. The color change provided information that the reaction has finished, which became useful for some reactions that only require as little as 10 hours to become complete. NOBF4 and NO2BF4 are both colorless; therefore the reaction cannot be visibly determined to be complete, so controlled reaction times were performed. [8–12] 35    REFERENCES 36    REFERENCES [1] M.P. Pechini, US Patent, 3,3306,97, 1967. [2] L.-W. Tai, P. a. Lessing, J. Mater. Res. 7 (1992) 502. [3] L.-W. Tai, P. a. Lessing, J. Mater. Res. 7 (1992) 511. [4] J.-Y. Bae, D. Lim, H.-G. Yun, M. Kim, J. Jin, B.-S. Bae, RSC Adv. 2 (2012) 5524. [5] M. Kakihana, M. Yoshimura, Chem. Soc. Japan 72 (1999) 1427. [6] T. Sreethawong, Y. Suzuki, S. Yoshikawa, J. Solid-state Chem. 178 (2005) 329. [7] B.D. Fahlman, Materials Chemistry, Springer Netherlands, Dordrecht, 2011. [8] A.R. Wizansky, P.E. Rauch, F.J. DiSalvo, J. Solid-state Chem. 81 (1989) 203. [9] R. Tripathi, T.N. Ramesh, B.L. Ellis, L.F. Nazar, Angew. Chem. Int. Ed. Engl. 49 (2010) 8738. [10] B.L. Ellis, W.R.M. Makahnouk, Y. Makimura, K. Toghill, L.F. Nazar, (2007) 749. [11] L.A. de Picciotto, M.M. Thackeray, Mater. Res. Bull. 20 (1985) 1409. [12] K. Takada, H. Sakurai, E. Takayama-muromachi, Nature 422 (2003) 53.   37      Chapter 4 Synthesis of NaFe2O3 Using Partial Oxygen Pressure Regulation Method 4.1 Introduction Crystal structures of a large class of transition metal oxides AMO2 (A = alkali metal, M = 3d metal) can be described as an ordered rock-salt lattice with cubic close-packed oxygen layers where A and M occupy the octahedral sites of alternating layers. These structures contain MO2 layers composed of face-shared MO6 octahedra (CdI2-type layers) separated by A+ ions. -NaFeO2 is the prototype compound, which lends its name to the structural type. Compounds with the -NaFeO2 structure are important from both fundamental and technological points of view. NaMO2 phases demonstrate a large variety of physical properties: frustrated magnetism in NaTiO2 [1], two successive orbital ordering transitions in NaVO2 [2], broad fluctuating cross-over regimes in NaCrO2 [3, 4], strongly coupled antiferromagnetic chains in NaMnO2 [5], two antiferromagnetic structures in NaFeO2 [6], and a nonmagnetic insulator state in NaCoO2 [7, 8]. From the technological point of view, LiCoO2 is an important cathode material for Li-ion batteries [9, 10] and NaxCoO2 is a candidate for thermoelectric applications [11-13]. Knowledge on AM2O3 rock-salt related compounds with layered A and M ordering is very limited. The only known example is ß-NaFe2O3; its crystal structure consists of double rock-salt layers of octahedrally coordinated iron (Fe2O3 blocks) separated by Na ions (Figure 4.1) [14-17]. Generally speaking, 3d metal compounds with ß-NaFe2O3 structures are expected to demonstrate different properties than their -NaFeO2 type 38      counterparts due to the difference in M oxidation states as well as in the dimensionality of the 3d metal - oxygen blocks. At the same time, synthesis of the mixed valent AM 22.5 O3 compounds is expected to require delicate control of the preparation conditions. In the case of ß-NaFe2O3, the situation is complicated by the existence of an -NaFe2O3 polymorph. Similarly to the ß phase, the crystal structure of -NaFe2O3 contains Fe2O3 blocks; however, these blocks are separated by two layers of octahedrally coordinated cations with (1/3 Fe + 2/3 Na) in this shared crystallographic position (Figure 4.1). Figure 4.1 Structure models of - and ß-NaFe2O3. The FeO6 octahedra are highlighted. For - NaFe2O3, atoms in the shared Na/Fe position are shown as large blue spheres and small red balls represent O ions. For ß-NaFe2O3, large green spheres correspond to Na ions. 39      Synthesis methods and crystal structures of - and ß-NaFe2O3 were reported in the 1970's by El Balkhi et al. [14, 15]. For synthesis of the NaFe2O3 polymorphs, mixtures of NaFeO2, Fe2O3, and Fe were sealed in ampoules and annealed at 1000 ºC for 48 hours. The  and ß phases were prepared in sealed steel and Cu ampoules respectively. Fe powder was also present in a separate tube inside the ampoule for the preparation of NaFe2O3. It needs to be emphasized that the experimental technique used relies on kinetic factors rather than on equilibrium conditions. At high temperatures, metal oxides generate some oxygen pressure. Materials of the ampoules as well as Fe powder reacted with O2, decreasing the partial O2 pressure (pO2), thus stabilizing NaFe2O3. At the same time, Fe metal from the ampoule walls cannot be in equilibrium with a Fe3+ containing phase, i.e. with NaFe2O3; therefore, reactions for extended times would result in further ampoule wall oxidation to FeO as well as NaFe2O3 decomposition to Na doped wüstite (FeO). This consideration is in agreement with a study of the equilibrium oxygen pressure in a broad composition range in the Fe–Na–O system by measurements in electrochemical cells with an oxygen-ion-conducting solid electrolyte [18]. A synthetic approach relying on equilibrium conditions is needed for the systematic search for AM2O3 compounds with different 3d metals. 40      4.2 Synthesis of NaFe2O3 via controlled oxygen pressure. The syntheses of α- and ß-NaFe2O3 were carried out in closed silica tubes under controlled partial oxygen pressure. Stoichiometric mixtures of α-NaFeO2, α-Fe2O3 (99.5%, Alfa Aesar) and Fe metal (≥99.9%, Sigma Aldrich) were mixed, pressed into pellets, placed in alumina boats, and closed in quartz tubes together with a second boat containing an oxygen buffer (oxygen getter) mixture. A split-open two-zone furnace was used to control the temperature of both the sample and getter mixtures independently. All samples were annealed at 850 °C for 48–72 hours. The temperature of the getter mixtures was varied between 350 °C and 900 °C in 29 different experiments, allowing for the precise control of pO2 in the system. After annealing, all quartz tubes were quenched to ambient temperature in air. The syntheses of all samples and subsequent X-ray diffraction measurements were successfully reproduced. It was reported that NaFe2O3 slowly hydrolyzes in air [14]; therefore, all operations with initial mixtures and final compounds were performed under nitrogen. α-NaFeO2 was synthesized by the reaction of α-Fe2O3 (99.5%, Alfa Aesar) with a 20% stoichiometric excess of NaOH (98.8%, Mallinckrodt Chemicals) under O2 flow at 500 °C for 48 hours. Excess NaOH was removed by washing with anhydrous methanol in a nitrogen-filled glove box. The samples were characterized by Powder X-ray diffraction (PXD) on a Bruker D8 Advanced with DAVINCI design diffractometer using Cu Kα radiation. The patterns 41      were recorded at room temperature with a step size of 0.02° (2θ) on a zero background sample holder. High resolution synchrotron powder X-ray diffraction data were collected using beamline 11-BM at the Advanced Photon Source, Argonne National Laboratory with an average wavelength λ=0.37382 Å. Discrete detectors covering an angular range from −6 to 16° 2θ were scanned over a 34° 2θ range, with data points collected every 0.001° 2θ and a scan speed of 0.01°/s. A three-axis translation stage holds the sample mounting and allows it to be spun, typically at ∼5400 RPM. Rietveld refinement [19] of SPXD data was performed with the GSAS [20] program and EXPGUI [21] interface. The profile function 4, which contains pseudo-Voigt, asymmetry, and microstrain broadening components, was employed for all phases. The DRAWxtl program was employed to prepare the difference Fourier map, (Figure 4.9) [22]. Elemental composition was determined by inductively coupled plasma (ICP) spectroscopy using a Vista-MPX CCD Simultaneous ICP-OES instrument (Varian Inc.). Oxygen stoichiometry was determined using triplicate iodometric titrations. KI (ACS Grade, Jade Scientific), K2S2O3 (≥95%, Sigma Aldrich), KIO3 (ACS Grade, Columbus Chemical), and Soluble Starch (ACS Grade, Columbus Chemical) were used in these titrations [23]. The initial attempt at the synthesis of NaFe2O3 at 850 °C under N2 flow resulted in a mixture of NaFeO2 and Fe3O4. In both compounds, the average iron oxidation state was above that in AM22.5+O3. It appears that the presence of about 2 ppm O2 in the N2 used 42      was sufficient to stabilize the Fe3+ compound NaFeO2, leading to the conclusion that the pO2 of the system was too high. Synthesis in a sealed quartz ampoule resulted in the mixture of α- and ß-NaFe2O3 polymorphs. At very low pO2, Na2O and Na doped wüstite (FeO) can be expected as equilibrium phases [18]. From this data, it becomes obvious that the synthesis of NaFe2O3 requires very careful control over oxygen stoichiometry. Since both the α- and ß-NaFe2O3 polymorphs were prepared earlier at distinct, but unknown pO2 [14], the dependence of phase composition vs. pO2 was studied in order to prepare the pure compounds. To control the pO2 during experiments, mixtures of Cu/Cu2O or Ni/NiO were employed as oxygen buffers. Under equilibrium conditions for mixtures of metal and metal oxides, the pO2 depends on both the chemical nature of the metal as well as temperature. Equilibrium pO2 temperature dependences were determined experimentally for many systems [24] and [25]. In the case of Ni/NiO, the oxygen fugacity follows the Equation: logfO2 atm =9.36- 24930 T 519 K16° 2θ) is enlarged by a factor of 20 for clarity. 54      Table 4.1 Crystallographic Data for the Rietveld Refinement of ß-NaFe2O3 and ßNa0.78Fe2O3 from SXPD data. (Goodness of Fit Parameters: 2 = 5.87, Rwp = 10.6%). Crystallographic Data for ß-Na1Fe2O3.a Atom Wyckoff position x y z 102 Uiso (Å2) Occ. Na 1b 0 0 0.5 1.06(3) 1 Fe 2d 1/3 2/3 0.1675(1) 0.64(1) 1 O1 1a 0 0 0 2.26(7) 1 O2 2d 1/3 2/3 0.7040(2) 1.03(3) 1  a Space group P 3m1 (No. 164). a = 3.05795(1) Å, c = 7.79807(2) Å; weight fraction 73%. Crystallographic Data for ß-Na0.78Fe2O3.b Atom b Wyckoff position x y z 102 Uiso (Å2) Occ. Na 1b 0 0 0.5 1.06(2) 0.78(1) Fe 2d 1/3 2/3 0.1705(2) 0.77(2) 1 O1 1a 0 0 0 2.28(9) 1 O2 2d 1/3 2/3 0.6416(6) 2.28(9) 1  Space group P 3m1 (No. 164). a = 3.05106(4) Å, c = 7.7898(2) Å; weight fraction 21%. 55      Table 4.2 Phase composition of nominal “NaFe2O3” stoichiometry at different partial O2 pressures at 850 ºC. Phase composition Equilibrium partial O2 pressure, atm [19, 20] Oxygen buffer (getter) mixture Temperature of the oxygen buffer mixture, ºC -NaFeO2, Fe3O4 ~2  10-6 - (N2 flow) - ß-NaFe2O3, NaFeO2, Fe3O4 2.7  10-10 Cu/Cu2O 750 3.5  10-11 Cu/Cu2O 700 2+ 3+ Na 0.88(2) Fe 0.93 Fe1.07 O 2.98(2) 1.8  10-14 Cu/Cu2O 550 2+ 3+ Na 0.80(2) Fe 0.89 Fe1.11 O 2.96(2) 7.5  10-16 Cu/Cu2O 500 2+ 3+ Na 0.88(2) Fe 0.9 Fe1.1 O 2.99(2) 5.5  10-17 Ni/NiO 700 2+ 3+ Na 0.82(2) Fe 0.89 Fe1.11 O 2.97(2) 3.0  10-19 Cu/Cu2O 400 2+ 3+ Na 0.76(2) Fe1.00 Fe1.00 O 2.88(2) 6.4  10-20 Ni/NiO 600 1.2  10-21 Ni/NiO 550 2+ Na 0.92(2) Fe1.09 Fe3+ 0.91O 2.91(2) 1.3  10-23 Ni/NiO 500 2+ 3+ Na 0.74(2) Fe 0.98 Fe1.02 O 2.88(2) 2.3  10-31 Ni/NiO 350 ß-NaFe2O3 -NaFe2O3 56    Chemical analysis results   ICP analysis confirmed Na loss via volatilization, giving an average composition of ßNa0.88(2)Fe2.00O3 for the sample. Knowing the sample composition and fraction of each phase present, one can calculate Na content (x) for ß-NaxFe2O3. ß-NaFe2O3, ß-NaxFe2O3, and Fe3O4 weight fractions were calculated during the Rietveld refinement and were equal to 0.73, 0.21, and 0.06, respectively, giving a derived composition of ßNa0.76Fe2O3. The Na content in ß-NaxFe2O3 refined from SPXD data was equal to 0.78, in good agreement with the chemical analysis results. According to the SPXD data, α-NaFe2O3 was essentially a pure phase. According to ICP analysis, a compositional formula of Na0.98(2)Fe2.00O3 was obtained. Peaks related to the ß-NaxFe2O3 polymorph were absent in the SPXD pattern while the strongest peak of the Fe3O4 admixture was at ∼0.1 % of relative intensity level. Attempts to perform a Rietveld refinement using the literature structural model [14] were unsuccessful (χ2≈22). While peak positions and peak profile could be fitted adequately, there were substantial differences in peak intensities, which did not change systematically with peak indices, indicating a problem with the structural model rather than preferred orientation. Attempts to include preferred orientation in the refinement were unsuccessful. Difference Fourier maps were calculated to determine the cause of said discrepancies. The difference Fourier map indicated regions of significant electron density around the shared Na/Fe and adjacent O3 positions (Figure 4.9). The extra electron density has a spherical shape near the O3 position, suggesting splitting of the position. The donut-like shape of the extra electron density around the shared Na/Fe position can be related to a Na/Fe shift from the average position. Attempts to split O3 and the shared Na/Fe positions to account for the 57      observed differences did not result in an improvement to the refinement; therefore, the crystal structure published by El Balkhi et al. [14] describes the average structure well, while the real structure is more complex with substantial disorder in the shared Na/Fe and O3 positions. Pair distribution function (PDF) analysis is needed to study the structure of α-NaFe2O3 in further detail. 58      Figure 4.9 Difference Fourier map for α-NaFe2O3 indicating large residual electron density (green) around the Na/Fe and O3 positions. Atomic positions are denoted for this layered structure and Fe–O octahedra are highlighted. Blue and red spheres represent atoms in the Na/Fe and O positions, respectively. 59      A nominal NaFe2O3 composition of both α- and ß-NaFe2O3 polymorphs implies the same formal Fe oxidation state in both compounds. It is not clear why α-NaFe2O3 begins to form at a pO2 eight orders of magnitude lower than ß-NaFe2O3. To clarify the origin of such a drastic difference in the stability conditions, stoichiometry of representative samples were determined by a combination of ICP analysis and iodometric titration. In spite of slight differences in Na content, oxygen stoichiometry of ß-NaFe2O3 samples prepared in a wide range of pO2 (3.5×10−11−5.5×10−17) were very close to ideal with a negligible amount of oxygen vacancies (Table 4.1). The Fe2+/Fe3+ ratio calculated from the sample's stoichiometry and charge balance requirements resulted in a 0.9:1.1 ratio (within error) for all ß-NaFe2O3 samples. Contrary to our expectations, oxygen nonstoichiometry in α-NaFe2O3 was minimal, with 0.1 oxygen vacancies per formula unit (Table 4.1). The Fe2+/Fe3+ ratio was between 1.0/1.0 and 1.1/0.9 for α-NaFe2O3 samples prepared under different experimental conditions. It has been shown that order in layered structures is favored by a large difference in the ionic radii (R) of A and M. An ionic radii ratio RM/RA<0.86 is typical for ordered phases [30], which coincides well with the empirical rule requiring less than a 15% ionic radii difference between two elements for substitution to occur. Theoretical investigation identified that the elastic energy associated with ions of different size stabilizes the ordering of M and A into alternate layers [31]. In high spin configurations with coordination number 6, the Shannon ionic radii values for Fe2+ and Fe3+ ions are equal to 0.920 and 0.785 Å, respectively [32]. The formal Fe ionic radii for α- and ß-NaFe2O3 are very close: 0.859 and 0.846 Å, respectively. Taking the large ionic radius of Na+ (1.16 Å) 60      into consideration, one can see that RFe/RNa≈0.74, which is well below the 0.86 value required for ordering. According to previous reports in the literature, structural strain has been shown to facilitate structural transformations [33]. The global instability index (GII) is often utilized as a metric of structural stress due to deviation of the bond valence sums of ions from ideal values [34]. The GII calculated for ß-Na1Fe2O3 using structural data from Table 4.2 was 0.14. While the GII indicates presence of structural strain, values above 0.2 generally correspond to unstable structures; thus, the simple radii ratio picture cannot unveil a reason for the ß- to α-NaFe2O3 transformation at low pO2. A correlation between Na content and α- or ß-NaFe2O3 polymorph formation (Table 4.2) was not established. Differences in Na content seem to be related to minute details of individual sample preparation, which indicates that further optimization of the synthetic conditions, is needed. Rietveld refinement of SPXD data for ß-NaFe2O3 points to the presence of two phases with different Na compositions. It is probable that multiple α- or ß-NaFe2O3 type line phases with well defined Na stoichiometry exist, as shown previously in NaxCoO2 [35]. Importantly, the existence of ß-NaxFe2O3 indicates that partial deintercalation of Na should be possible by soft chemistry methods at room temperature, allowing for the fine tuning of the properties of the compounds by controlling the oxidation state of Fe. In the case of NaxCoO2, such fine tuning results in a complex Na-content vs. physical properties diagram [36]. It would be intriguing to explore the properties of ß-NaxFe2O3 with different Na compositions. 61      Oxygen stoichiometry of α- and ß-NaFe2O3 differs by ∼0.1 per formula unit. Values very close to full oxygen stoichiometry were consistently observed for ß-NaFe2O3 while the αNaFe2O3 phase compositions were systematically close to the 2.9 oxygen stoichiometry (Table 4.1). Based on the above consideration, it seems reasonable to assume that a higher stability of α-NaFe2O3 at low pO2 is related to its ability to accommodate a greater amount of oxygen vacancies than that for ß-NaFe2O3. The sodium volatility during the reaction led to further investigation of controlled sodium deintercalation using bromine. The bromine oxidation potential to deintercalate sodium against iron in the 3+ oxidation states is minimal; therefore, the oxidation of Fe2+ to Fe3+ in NaFe2O3 will occur and form the stable salt NaBr, as seen in Figure 4.10. 62      Figure 4.10 PXD of unwashed sodium deintercalated ß-NaxFe2O3 (red ticks) with the byproduct NaBr (green ticks). Similar to the results seen in Table 4.1, the result of the controlled removal of sodium from the initial compound will lead to the formation of an increased amount of Fe3+. Chemical sodium deintercalation was performed in increasing increments of five mole percent of bromine starting at 0% and ending at 50%. This would result in a phase diagram of the sodium deficient NaxFe2O3 from the initial sodium deficient NaFe2O3 from the initial synthetic reaction, to approximately 0.5 parts per formula unit of sodium removed, or approximately Na0.5Fe2O3. The samples were prepared starting with the 63      initial compound ß-Na0.96Fe2O3, determined from ICP, and reacted as previously described. The resulting powders were washed and analyzed for purity using a powder xray diffractometer. After the samples were deemed pure, further analysis using ICP to determine the actual stoichiometry of the sodium deficient phase was performed. Le Bail fit, to acquire cell parameters, was then performed on the pure phases and the cell parameters were plotted against the sodium content as seen in Figures 4.11-4.13. Figure 4.11 Cell parameter a as a function of parts per formula unit of sodium in NaxFe2O3 Le Bail fit from PXD. 64      According to Vegard's law, the cell parameters are expected to continue linearly as a function of sodium content in a crystal structure. This trend is seen in both the a and c parameters along with the volume, which is expected due the relationship of volume to the a and c parameters of a trigonal space group. Figure 4.12 Cell parameter c as a function of parts per formula unit of sodium in NaxFe2O3 Le Bail fit from PXD. 65      As it can be seen in the Figures 4.11-4.13, there is an absence of the stable phases with the approximate formulas of Na0.75Fe2O3 and Na0.65Fe2O3. Multiple attempts were made to synthesize these two compounds to no avail. Therefore, it can be determined that these two sodium deficient phases are not kinetically stable. Coupled with the Rietveld refined compound in Table 4.1, Na0.78Fe2O3, it can be then said that the stability of the structure within this range only varies by a few mole percent of sodium. Figure 4.13 Volume as a function of parts per formula unit of sodium in NaxFe2O3 Le Bail fit from PXD. 66      Figures 4.11-4.13 are summarized in table 4.3 with the determined Le Bail fit values from laboratory PXD. Table 4.3 NaxFe2O3 Cell parameters a, c and volume determined from Le Bail fit from PXD. (x) in NaxFe2O3 Cell Parameter a Cell Parameter c Cell Volume 0.96 3.0792 7.7765 63.865 0.91 3.0746 7.7746 63.646 0.87 3.0693 7.7705 63.394 0.81 3.0640 7.7658 63.140 0.76 3.0576 7.7625 62.848 0.67 3.0461 7.7576 62.336 0.56 3.0356 7.7513 61.858 0.51 3.0317 7.7476 61.669 0.46 3.0232 7.7448 61.303 Synthesis of transition metal doped NaFe2-XMxO3 was performed using manganese, cobalt, nickel, and copper. These transition metals were chosen due to their ability to form stable 2+ oxidation states along with having relatively similar ionic radii to iron. These two requirements are essential for the doping to be successful. Initial attempts were performed at a pO2 of about 10-23 atm. This pO2 is in the stability zone of α-NaFe2O3, but due to the substituted transition metals' oxidation potentials different from that of iron, it can be expected that the reactivity of the doped NaFe2-xMxO3 will differ from the 67      undoped NaFe2O3. With that said, the α-NaFe2O3 zone of stability was chosen to be on the cautious side. In the case that the reactivity of the doped transition metals required less oxidizing conditions to form, the formation of the ß- structure would, at the very least, appear in a mixture of phases in a PXD. This was observed with the cobalt-doped compounds NaFe1.75Co0.25O3 and NaFe1.85Co0.15O3. As seen in Figure 4.14 at a pO2 of about 10-23 atm, with NaFe0.5Co0.5O3, the α-NaFe2O3 structure formed. Figure 4.14 NaFe1.5Co0.5O3 synthesized at 850 °C at a pO2 of about 10-23 atm resulting in the formation of the α-NaFe2O3 structure type (red ticks). 68      Similarly, at a pO2 of about 10-23 atm, NaFe1.75Co0.25O3 (Figure 4.15) and NaFe1.85Co0.15 (Figure 4.16) were synthesized resulting in mixtures of ß-NaFe2O3 structure and NaFeO2 for NaFe1.75Co0.25O3, and ß-NaFe2O3 mixed with NaFeO2 and Fe3O4 for NaFe1.85Co0.15. This illustrates nicely the effect of the different doping amounts on reactivity at the same pO2. Figure 4.15 NaFe1.75Co0.25O3 synthesized at 850 °C at a pO2 of about 10-23 atm resulting in the formation of the ß-NaFe2O3 structure type (red) and γ-NaFeO2 (green). 69      Attempts to synthesize NaFe1.25Co0.75O3 were deemed unsuccessful with the appearance of cobalt oxide. With this information, it can be said that doping of much more than 0.5 parts per formula unit of a transition metal will lead to the appearance of the respective transmission metal oxide. Figure 4.16 PXD of NaFe1.85Co0.15O3 synthesized at 850 °C at a pO2 of about 10-23 atm resulting in the formation of the ß-NaFe2O3 structure type (red), γ-NaFeO2 (green), and Fe3O4 (blue). 70      Doping NaFe2O3 with manganese was performed and the initial attempt was deemed successful. The attempt to synthesize NaFe1.5Mn0.5O3 was performed at 850 °C at a pO2 of about 10-23 atm. The resulting PXD, as seen in Figure 4.17, shows the formation of the α-NaFe2O3 structure, γ-NaFeO2, and Fe3O4. These results with both cobalt and manganese doping were deemed successful due to the formation of either the α- or ßphase, indicating that further attempts with different buffer conditions to form the pure phases of both α- and ß-NaFe2-xMxO3 will form the pure phases.  Figure 4.17 NaFe1.5Mn0.5O3 synthesized at 850 °C at a pO2 of about 10-23 atm resulting in the formation of the α-NaFe2O3 (red), structure type γ-NaFeO2 (blue), and Fe3O4 (green). 71      Attempts to dope copper and nickel into NaFe2O3 were deemed unsuccessful. The resulting PXD of both reactions still possessed CuO and NiO indicating no reaction occurred. This result is not surprising due to the relatively high stability of the two oxides. 4.6 Conclusion A phase composition vs. partial O2 pressure diagram at 850 °C for Na/Fe=1/2 was determined which allowed for the reproducible synthesis of the divalent sodium iron oxides, α- and ß-NaFe2O3 polymorphs. Sodium extraction from ß-NaFe2O3 at room temperature was successfully investigated thus modifying the average Fe oxidation state. Doping cobalt and manganese into the iron position was also successfully performed and illustrates the ability to form dopant series of the compound NaFe2-xMxO3 (03,000 LiMn2O4[9] Spinel 144 110 3.2 >1000 Due to the success of LiCoO2 with the layered α-NaFeO2 structure, as previously stated similar layered compounds were investigated, including LiFeO2. LiFeO2 has nine known polymorphs including four ordered and disordered cubic rocksalt (NaCl) structure polymorphs. They differ depending on the order or disorder of the cations and are referred to as α-, ß-, ß'-, and γ-LiFeO2 [17], the layered α-NaFeO2 structure [18], the corrugated-layer [18], Goethite-type [2], Hollandite-type [19], and the tetrahedral LiFeO2 (T-LiFeO2) [20]. This compound is of particular interest due to iron’s cost, abundance, and benign toxicity. The nine polymorphs can be synthesized though a variety of different methods including solid state reaction, low-temperature molten salt synthesis, hydrothermal, solvothermal, ion-exchange, and heat treatment. Table 5.2 illustrates a few synthetic procedures, the structure, and the best electrochemical performance for the different LiFeO2 polymorphs. 82      Table 5.2 Polymorphs of LiFeO2, structure, preparation, preparation procedure, and best electrochemical performance. Type Structure Preparation Preparation Procedure Best Electrochemical Performance (mAh/g) α-LiFeO2[17,21] Fm3m (cubic) Cations disordered Solid State Reaction Li2CO3 and Fe2O3 annealed above 700 °C 150 -LiFeO2[17,21] Fm3m (cubic) Cations disordered Solid State Reaction Li2CO3 and Fe2O3 annealed at 400 °C * Heat Treatment α-LiFeO2 heated at 400 °C * '-LiFeO2[17,21] Fm3m (cubic) Cations disordered Fm3m (cubic) Cations ordered R-3m (rhombohedral) Pbnm (orthorhombic) Li2CO3 and Fe2O3 annealed above 600 °C α-NaFeO2 and LiNO3 Layered-LiFeO2[18,21] Ion-exchange eutectic melt α-FeOOH and C2H5OLi in Geothite- LiFeO2[2,21] Ion-exchange ethanol at 170 °C ß-FeOOH and LiOH HollanditeI4/3 (tetragonal) Ion-exchange eutectic melt LiFeO2[19,21] Corrugated layerPmnm γ-FeOOH and C2H5OLi in Ion-exchange LiFeO2[2,21,22] (orthorhombic) ethanol at 170 °C Pna21 ß-NaFeO2 and LiNO3 T-LiFeO2[20] Ion-exchange eutectic melt (orthorhombic) * Do not have electrochemical performance due to the blocked diffusion pathway by the ordering of the cations. γ-LiFeO2[17,21] Solid State Reaction 83    * 115 70 150 150 120   As seen in Table 5.2, the layered α-NaFeO2 structure type LiFeO2 cannot be prepared by the conventional high-temperature solid-state route, rather the precursor α-NaFeO2 must first be prepared followed by ion exchange with a lithium salt. The direct solid-state route leads to the synthesis of different crystallographic forms of LiFeO2 with the rock-salt structure. These ordered and disordered rock-salt structures do not show favorable intercalation and deintercalation due to their structure type, with the exception of αLiFeO2, of which the capacity eventually fades to less than 100 mAh/g. A nanoporous αLiFeO2 carbon composite was shown to have an initial capacity of 255 mAh/g, about 90 % its theoretical capacity of 282 mAh/g. This further illustrates the advancements nanoparticle preparation and carbon coating have on materials [23]. Particle size of LiFeO2 with the layered α-NaFeO2 structure was investigated and it was similarly determined that smaller particle size increased the discharge capacity. It was established that samples with an average particle size of 400 nm exhibited a capacity of 80 mAh/g while samples with an average particle size of 40 nm showed a higher capacity of 115 mAh/g. Each study was performed with a cycling voltage between 2.0 and 4.5 V [18]. T-LiFeO2 is of particular interest due to its tunnel-like structure similar to that seen in LiFePO4 [20]. This structure is also the only LiFeO2 polymorph that contains iron-oxygen in tetrahedral coordination, whereas the other polymorphs are octahedral. T-LiFeO2 is synthesized through ion exchange from the precursor ß-NaFeO2, which has a similar structure. The precursor was ball milled prior to ion exchange to decrease the particle size of the final compound. Electrochemical characterization of the T-LiFeO2 polymorph 84      eventually stabilized with a capacity of about 100 mAh/g, but further analysis of the compounds showed a reaction with the electrolyte resulting in the formation of hydrogen, carbon monoxide, and carbon dioxide evolution. This reaction also led to the decomposition of T-LiFeO2 to the spinel structure LiFe5O8. The major flaw of LiFeO2 in any of its polymorphs is its relatively low electrical conductivity; therefore, carbon coating or nano-particles, as previously stated, is a route commonly taken. Doping other transition metals into a structure is not only a route taken for reducing the amount of some elements as seen in LiCo1/3Mn1/3Ni1/3O2, but can also be performed to increase desired properties such as electrical conductivity. An example of this is cobalt doping into the α-NaFeO2 layered structure LiFeO2 to synthesize LiFe0.9Co0.1O2 [24]. LiFe0.9Co0.1O2 was synthesized simultaneously with an undoped LiFeO2, both samples with the α-NaFeO2 layered structure. It was reported that LiFeO2 had an initial capacity of 205 mAh/g that eventually faded and stabilized at 140 mAh/g. Similarly, the LiFe0.9Co0.1O2 had an initial capacity of 205 mAh/g, which faded and stabilized at 190 mAh/g. This finding illustrates that doping cobalt into layered LiFeO2 produces a higher stabilized capacity than that of layered LiCoO2 or LiFeO2. Further investigation into the T-LiFeO2 polymorph with a tunnel-like structure will be discussed herein. The effects of chemical lithium intercalation and chemical lithium deintercalation on the structure were analyzed along with Mössbauer spectroscopy. Doping T-LiFeO2 with cobalt was also investigated to determine if the electrochemical performance and structural stability could be enhanced and maintained respectively without transformation into LiFe5O8 and decomposition of the electrolyte. 85      5.2 Synthesis of ß-NaFeO2 and NaFe1-xCoxO2 ß-NaFeO2 was synthesized through the directed ceramic method with stoichiometric amounts of Fe2O3 (99.5%, Alfa Aesar) and 5 % excess Na2CO3 (99%, Mallinckrodt Chemicals), to compensate for sodium volatility. ß-NaFe1-xCoxO2 (0