DESIGN AND PROPERTIES OF NOVEL URANIUM-CONTAINING LAYERED AND FRAMEWORK MATERIALS Except where reference is made to the work of others, the work described in this dissertation is my own or was done in collaboration with my advisory committee. This dissertation does not include proprietary or classified information. ????????????????????????????????? Tatiana Yurievna Shvareva Certificate of Approval: ?????????????????????????????? ?????????????????????????????? Peter Livant Thomas E. Albrecht-Schmitt, Chair Associate Professor Associate Professor Chemistry Chemistry ?????????????????????????????? ?????????????????????????????? Holly R. Ellis Andreas J. Illies Assistant Professor Professor Chemistry Chemistry ?????????????????????????????? Joe F. Pittman Interim Dean Graduate School DESIGN AND PROPERTIES OF NOVEL URANIUM-CONTAINING LAYERED AND FRAMEWORK MATERIALS Tatiana Yurievna Shvareva A Dissertation Submitted to the Graduate Faculty of Auburn University in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy Auburn, Alabama December 15, 2006 iii DESIGN AND PROPERTIES OF NOVEL URANIUM-CONTAINING LAYERED AND FRAMEWORK MATERIALS Tatiana Yurievna Shvareva Permission is granted to Auburn University to make copies of this dissertation at its discretion, upon request of individuals or institutions and at their expense. The author reserves all publication rights. ______________________________ Signature of Author ______________________________ Date of Graduation iv VITA Tatiana Yurievna (Solonskaya) Shvareva, daughter of Yuri A. Solonsky and Galina A. (Shabalina) Solonskaya, was born on April 8, 1973, in St. Petersburg, Russia. She graduated from High School N. 419, Petergoff District, St. Petersburg in 1990, and then from St. Petersburg State University in 1995. She obtained a Bachelor of Science degree in Chemistry under the direction of Dr. Nadezhda S. Grigorova and Dr. Florentina A. Belinskaja. She entered Graduate School at Auburn University in the fall of 2003 to pursue a Doctoral degree in Chemistry under the guidance of Dr. Thomas E. Albrecht- Schmitt. She married Alexey Y. Shvarev, son of Yevgeniy A. Shvarev and Inna N. (Gulika) Shvareva on July 6 th , 1996. v DISSERTATION ABSTRACT DESIGN AND PROPERTIES OF NOVEL URANIUM-CONTAINING LAYERED AND FRAMEWORK MATERIALS Tatiana Y. Shvareva Doctor of Philosophy, December 15, 2006 (B. S., St. Petersburg State University, 1995) 217 Typed Pages Directed by Thomas E. Albrecht-Schmitt Ten uranyl-containing frameworks and one layered material have been synthesized under hydrothermal conditions. These compounds belong to eight structure types, not previously described. They are constructed from uranyl units in different coordination environments, phosphate or iodate anions, and incorporate main group or transition metal polyhedra. Six materials were found to have selective ion exchange properties. Their stabilities at elevated temperatures were examined. Thermally stable ion exchangers could be recommended for the further investigations under the simulated conditions of nuclear waste. Introduction of transition metals also leads to the appearance of non-linear optical and magnetic properties. A new synthetic method for the production of uranyl-containing framework materials is also proposed. vi ACKNOWLEDGMENT I wish to express my sincerest gratitude to all people, inside and outside of the Chemistry Department, who helped me during this work. But the very first, I would like to thank my research advisor, Professor Thomas E. Albrecht-Schmitt. He opened the world of material chemistry to me, introduced and made me feel welcome in the actinide research community. Now I leave his group with certain wish to spend my entire career in actinide material science and to make him proud. I would like to thank all former and present group members, whom I worked with: Richard, Phil, Tyler, Shehee, Jin (especially Jin), Ling, Travis, Noel and Yu for friendly and very helpful atmosphere in the group. In addition I need to thank little Angie Mustain for being very good student and very good friend. I am also very grateful to my committee members, Dr. Illies, Dr. Livant and Dr. Ellis for their advice and help along the way. I would like to express my especially warm ?thank you? to Dr. Lorraine Wolf and Dr. David Stanbury for the greatest help that I could ever imagine during my last year here at Auburn. They made my trip to Texas possible and I will never forget it. I would like to thank my parents for supporting me in any decision that I have ever made. I owe them a lot for never-ending patience and ?therapy? when I was discouraged. vii And finally I need to say that this work would never have been done without understanding, patience and day-by-day support of my love, Barbara, and Alexey, the greatest luck in my life. Thank you all very much. viii Style manual or journal used: American Chemical Society style ________________________________________________________________________ Computer software used: Microsoft Word 2000, Microsoft Excel, Atom v. 6.0, Adobe Illustrator 10 ________________________________________________________________________ ix TABLE OF CONTENTS LIST OF FIGURES ...........................................................................................................xv LIST OF TABLES......................................................................................................... xxiii CHAPTER 1. INTRODUCTION ........................................................................................1 CRYSTAL CHEMISTRY OF URANIUM (VI) UNDER HYDROTHERMAL CONDITIONS..............................................................2 HIGH-LEVEL NUCLEAR WASTE (HLNW) ............................................10 ION-EXCHANGE ........................................................................................14 ION-EXCHANGE LAYERED AND THREE-DIMENSIONAL INORGANIC MATERIALS FOR SELECTIVE Cs + AND Sr 2+ UPTAKE.......................................................................................................16 MAIN GROUP AND TRANSITION METAL PHOSPHATES .............................................................................................24 REFERENCES .............................................................................................27 CHAPTER 2. CRYSTAL CHEMISTRY AND ION-EXCHANGE PROPERTIES OF THE LAYERED POTASSIUM URANYL IODATE, K[UO 2 (IO 3 ) 3 ]...............................32 ABSTRACT..................................................................................................32 INTRODUCTION ........................................................................................33 x EXPERIMENTAL........................................................................................34 SYNTHESIS K[UO 2 (IO 3 ) 3 ].................................................................35 CRYSTALLOGRAPHIC STUDIES...................................................35 ION EXCHANGE ...............................................................................36 RESULTS AND DISCUSSION...................................................................39 SYNTHESIS........................................................................................39 STRUCTURE. K[UO 2 (IO 3 ) 3 ]..............................................................39 ION EXCHANGE STUDIES..............................................................42 CONCLUSIONS...........................................................................................45 REFERENCES .............................................................................................47 CHAPTER 3. SYNTHESES, STRUCTURES, AND ION-EXCHANGE PROPERTIES OF THE THREE-DIMENSIONAL FRAMEWORK URANYL GALLIUM PHOSPHATES, Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ]?H 2 O AND Cs [UO 2 Ga(PO 4 ) 2 ] ...................50 ABSTRACT..................................................................................................50 INTRODUCTION ........................................................................................51 EXPERIMENTAL........................................................................................53 SYNTHESIS. Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ](H 2 O) (UGaP1).................54 SYNTHESIS. Cs[UO 2 Ga(PO 4 ) 2 ] (UGaP2) .........................................54 CRYSTALLOGRAPHIC STUDIES...................................................54 ION EXCHANGE STUDY .................................................................60 THERMAL ANALYSIS .....................................................................60 RESULTS AND DISCUSSION...................................................................60 xi SYNTHESES.......................................................................................60 STRUCTURE Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ](H 2 O) (UGaP1)................61 STRUCTURE Cs[UO 2 Ga(PO 4 ) 2 ] (UGaP2).........................................67 THERMAL ANALYSIS .....................................................................70 ION EXCHANGE ...............................................................................75 CONCLUSIONS...........................................................................................78 REFERENCES .............................................................................................79 CHAPTER 4. SYNTHESIS, STRUCTURES AND PROPERTIES OF THE MULTIFUNCTIONAL URANYL VANADIUM PHOSPHATE FRAMEWORKS ...............................................................................................................83 ABSTRACT..................................................................................................83 INTRODUCTION ........................................................................................84 EXPERIMENTAL........................................................................................85 SYNTHESIS. Cs 2 [UO 2 (VO 2 ) 2 (PO 4 ) 2 ]?nH 2 O (UV V P) .........................85 SYNTHESIS. K 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (KUV IV P).........................86 SYNTHESIS. Rb 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (RbUV IV P) .....................86 CRYSTALLOGRAPHIC STUDIES...................................................86 THERMAL ANALYSIS .....................................................................87 NON-LINEAR OPTICAL PROPERTIES ..........................................87 MAGNETIC PROPERTIES................................................................92 ION EXCHANGE STUDIES..............................................................93 RESULTS AND DISCUSSION...................................................................94 xii SYNTHESES.......................................................................................94 STRUCTURE. Cs 2 [UO 2 (VO 2 ) 2 (PO 4 ) 2 ]?nH 2 O (UV V P)........................95 STRUCTURE. K 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (KUV IV P) AND Rb 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (RbUV IV P).............................................98 THERMAL ANALYSIS ...................................................................104 NON-LINEAR OPTICAL PROPERTIES ........................................108 MAGNETIC PROPERTIES..............................................................108 ION EXCHANGE STUDY ...............................................................115 CONCLUSIONS.........................................................................................116 REFERENCES ...........................................................................................117 CHAPTER 5. HYDROTHERMAL SYNTHESIS AND STRUCTURAL CHARACTERIZATION OF Cs 1.52 [(UO 2 )(TiOH 0.48 )(PO 4 ) 2 ]?nH 2 O AND Cs 1.35 [(UO 2 )(GeOH 0.65 )(PO 4 ) 2 ]?nH 2 O..............................................................................120 ABSTRACT................................................................................................120 INTRODUCTION ......................................................................................121 EXPERIMENTAL......................................................................................123 SYNTHESIS. Cs 1.52 [(UO 2 )(TiOH 0.48 )(PO 4 ) 2 ]?nH 2 O (UTiP) .............124 SYNTHESIS. Cs 1.35 [(UO 2 )(GeOH 0.65 )(PO 4 ) 2 ]?nH 2 O (UGeP)...........124 CRYSTALLOGRAPHIC STUDIES.................................................125 RESULTS AND DISCUSSION.................................................................129 SYNTHESES.....................................................................................129 STRUCTURE ....................................................................................129 xiii CONCLUSIONS.........................................................................................137 REFERENCES ...........................................................................................140 CHAPTER 6. GENERAL ROUTE TO THREE-DIMENSIONAL FRAMEWORK URANYL TRANSITION METAL PHOSPHATES WITH ATYPICAL STRUCTURAL MOTIFS: THE CASE EXAMPLES OF Cs 2 {(UO 2 ) 4 [Co(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 }, Cs 2 {(UO 2 ) 4 [Ni(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 }, AND Cs 3+x [(UO 2 ) 3 CuH 4-x (PO 4 ) 5 ]?H 2 O............................................................................144 ABSTRACT................................................................................................144 INTRODUCTION ......................................................................................145 EXPERIMENTAL......................................................................................146 SYNTHESIS. Cs 2 {(UO 2 ) 4 [Co(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 }(UCoP).........146 SYNTHESIS. Cs 2 {(UO 2 ) 4 [Ni(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 } (UNiP) .........147 SYNTHESIS. Cs 3+x [(UO 2 ) 3 CuH 4-x (PO 4 ) 5 ]?H 2 O (UCuP)...................147 X-RAY STRUCTURAL ANALYSIS...............................................147 THERMAL ANALYSIS ...................................................................148 UV-VIS-NIR DIFFUSE REFLECTANCE SPECTRA.....................156 MAGNETIC PROPERTIES..............................................................156 ION EXCHANGE STUDIES............................................................157 RESULTS AND DISCUSSION 158 SYNTHESES.....................................................................................158 STRUCTURE. Cs 2 {(UO 2 ) 4 [Co(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 }(UCoP) AND Cs 2 {(UO 2 ) 4 [Ni(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 } (UNiP) .......................159 xiv STRUCTURE . Cs 3+x [(UO 2 ) 3 CuH 4-x (PO 4 ) 5 ]?H 2 O (UCuP)................162 UV-VIS-NIR DIFFUSE REFLECTANCE STUDY.........................171 MAGNETIZATION MEASUREMENTS.........................................174 THERMAL ANALYSIS ...................................................................178 ION EXCHANGE PROPERTIES.....................................................183 CONCLUSIONS ........................................................................................184 REFERENCES ...........................................................................................186 CHAPTER 7. SUMMARY..............................................................................................189 xv LIST OF FIGURES Figure 1. 1. The UO 2 2+ cation in tetragonal bipyramidal, pentagonal bipyramidal, and hexagonal bipyramidal environments as common building units for uranyl structure formation. They can be interconnected by corner- or edge- sharing with each other or other building units in order to construct different structure types ..................................................................................5 Figure 1. 2. The structure of (C 4 N 2 H 12 )U 2 O 4 F 6 composed of only uranyl pentagonal bipyramids (green). Piperazinium cations are located in the channels of the framework .................................................................................................6 Figure 1. 3. The structure of K 2 [(UO 2 ) 2 (VO) 2 (IO 6 ) 2 O]?H 2 O is built from distorted octahedral building units only. UO 6 polyhedra are shown in green, VO 6 ?in light blue, and IO 6 ? in red color. K + cations along with water molecules occupy the tunnels. The UO 2 2+ uranyl unit exists as a nearly linear trans dioxo unit .....................................................................................7 Figure 1. 4. The structure of Li 2 (UO 2 ) 4 (WO 4 )O is an example of a three-dimensional structure with no channels. Li + cations located in the voids of uranyl tungstate layers are not seen from this projection. UO 6 polyhedra are green, WO 6 ? shown in violet color................................................................9 xvi Figure 1. 5. Chiral channels extended along c axis make the uranyl molybdate (NH 4 ) 4 [(UO 2 ) 5 (MoO 4 ) 7 ](H 2 O) 5 potentially applicable for enantio- selective separation. NH 4 + cations are not shown, UO 7 polyhedra are green, MoO 4 tetrahedra are blue .............................................................11 Figure 1. 6. A view of the crystal structure of Na 2 Ti 2 O 3 (SiO 4 )?H 2 O along the [001] direction. Yellow squares depict TiO 6 octahedra, orange - SiO 4 tetrahedra, green balls ? Na + cations. Four out of the total eight Na + cations are located in the channels, the other four (on the silicate groups) ? in small cavities. Copied from reference 50 with author?s permission.....................................................................................................18 Figure 1. 7. This fragment depicts interconnection of octahedral units in the structure of Na 2 Nb 2-x Ti x O 6-x (OH) x ?H 2 O. Blue polyhedra represent disordered NbO 6 and TiO 6 octahedra. Yellow polyhedra are Na coordination spheres, participating in the framework. Copied from reference 52 with author?s permission.....................................................................................................23 Figure 2. 1. A view of the pentagonal bipyramidal coordination environment around uranium center in KUI that is formed from UO 2 2+ linear cation connected with five iodate groups through the equatorial plane.............................................................................................41 Figure 2. 2. An illustration of the 2 ? [UO 2 (IO 3 ) 3 ] 1? sheets in KUI that extend in the [ab] plane ......................................................................................................43 Figure 2. 3. A view along the b axis of KUI showing that the monodentate iodate anion protrudes in between the layers, providing its terminal oxo xvii atoms for forming interactions with the K + cations that reside between the layers .......................................................................................................44 Figure 3. 1. A view of the one-dimensional 1 ? [Ga(OH)(PO 4 ) 2 ] 4? chains composed of corner-sharing GaO 6 octahedra and bridging PO 4 tetrahedra that extend along the c axis in Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ]?H 2 O (UGaP1). 50% probability ellipsoids are depicted........................................................63 Figure 3. 2. A depiction of the one-dimensional 1 ? [UO 2 (PO 4 ) 2 ] 4? chains in Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ]?H 2 O (UGaP1). 50% probability ellipsoids are depicted ...................................................................................................67 Figure 3. 3. An illustration of the three-dimensional 3 ? [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ] 4? anionic lattice of Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ]?H 2 O (UGaP1) that has intersecting channels running down the b and c axes. These channels house Cs + cations and water molecules. Cs is shown in magenta, UO 7 pentagonal bipyramids are shown in green, GaO 6 octahedra in blue, and PO 4 tetrahedra in yellow ...............................................................69 Figure 3. 4. A view of the one-dimensional 1 ? [Ga(PO 4 ) 2 ] 3? gallium phosphate chains in Cs[UO 2 Ga(PO 4 ) 2 ] (UGaP2) formed from fused eight- membered rings of alternating corner-shared GaO 4 and PO 4 tetrahedra. 50% probability ellipsoids are depicted........................................................70 Figure 3. 5. An illustration of the one-dimensional 1 ? [UO 2 (PO 4 ) 2 ] 4? chains in Cs[UO 2 Ga(PO 4 ) 2 ] (UGaP2). 50% probability ellipsoids are depicted ........72 xviii Figure 3. 6. A depiction of the three-dimensional 3 ? [UO 2 Ga(PO 4 ) 2 ] 1? framework in Cs[UO 2 Ga(PO 4 ) 2 ] (UGaP2) with intersecting channels running down the a and b axes. Cs is shown in magenta, UO 7 pentagonal bipyramids are shown in green, GaO 4 tetrahedra in blue, and PO 4 tetrahedra in yellow............................................................................................................73 Figure 3. 7. DSC data of UGaP1 demonstrate relative stability of the structure at temperatures below 400 ?C. Two peaks at 150 ?C and 172 ?C correspond to lost of surface and interstitial water respectively...................74 Figure 3. 8. DSC data of UGaP2.......................................................................................77 Figure 4. 1. Depiction of the channels of UV V P, extended along b and c axes, occupied by Cs + cations. Green ? uranium, blue ? vanadium, yellow ? phosphorus, purple ? cesium centers .............................................97 Figure 4. 2. The fundamental building unit of UV V P shown in 50% probability thermal ellipsoids........................................................................................100 Figure 4. 3. Thermal ellipsoid plot of the fundamental unit of K 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (KUV IV P) and Rb 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (RbUV IV P) at 50% probability....................................................................101 Figure 4. 4. Depiction of essential building blocks of the structures of KUV IV P and RbUV IV P. V 2 O 9 dimer, constructed from two vanadium octahedra, sharing two equatorial and one axial oxygen atoms (top) and vanadyl phosphate layers (bottom). Cations (purple) occupy the channels 2.8 ? ? 4.4 ? along c axis. Blue - vanadium polyhedra, yellow ? phosphate groups ........................................................................................104 xix Figure 4. 5. Depiction of the channels in the framework of KUV IV P and RbUV IV P, filled with K + or Rb + cations, extended along a axis. Blue - vanadium, yellow ? phosphorus, green ? uranium centers, purple ? K + or Rb + cations ..........................................................................106 Figure 4. 6. A view of the framework of KUV IV P and RbUV IV P along the b axis ........107 Figure 4. 7. DSC data for the structure of UV V P demonstrate thermal stability up to 510 ?C ................................................................................................108 Figure 4. 8. DSC data for the structure of KUV IV P. Several endothermic events between 160 ?C and 550 ?C reveal gradual structural transformations with increase of temperature.......................................................................110 Figure 4. 9. Magnetization of UV V P as a function of applied magnetic field at 5 K. Line is the linear fit to the data ...................................................................111 Figure 4. 10. Magnetic susceptibility of UV V P as a function of temperature with a 10000 G magnetic field ...................................................................112 Figure 4. 11. Magnetization of RbUV IV P as a function of applied magnetic field at 5 K. Line is the linear fit to the data ......................................................114 Figure 4. 12. Magnetic susceptibility of RbUV IV P as a function of temperature. Line is the fit to Curie-Weiss behavior. Inset show the temperature dependence of the inverse magnetic susceptibility.....................................115 Figure 5. 1. Thermal ellipsoid plot of Cs 1.52 [(UO 2 )(TiOH 0.48 )(PO 4 ) 2 ]?nH 2 O (UTiP) shown at 50% probability ...........................................................................132 Figure 5. 2. Depiction of one-dimensional titanium phosphate or germanium phosphate and uranyl phosphate chains in UTiP or UGeP. xx Violet - titanium or germanium polyhedra, yellow-phosphate tetrahedra, green-uranyl building units.......................................................133 Figure 5. 3. The framework of UTiP and UGeP. The channels along the c axis are filled with Cs + cations (purple) and water molecules (red). Violet octahedra represent Ti or Ge, yellow ? phosphate, and green ? uranium centers .........................................................................................................137 Figure 6. 1. The fundamental building unit of UCoP includes three crystallographically unique U centers, two in pentagonal bipyramidal, and one in hexagonal bipyramidal environment, three tetrahedral P centers, and Co octahedra. 50% probability thermal ellipsoids are represented. UNiP is isostructural with UCoP............................................162 Figure 6. 2. View down the c axis of the uranyl phosphate layers in UCoP and UNiP. One of the phosphate anions is disordered and is shown in ball-and-stick format. Uranium polyhedra are in green and phosphate are in yellow ..............................................................................163 Figure 6. 3. An illustration of the three-dimensional structure of UCoP and UNiP as viewed down the b axis. Uranium polyhedra are in green, phosphate in yellow, cobalt or nickel in orange. Purple centers represent the Cs + cations.............................................................................165 Figure 6. 4. A view of the uranyl tetragonal and pentagonal bipyramids, square planar Cu and phosphate tetrahedra in UCuP with 50% probability ellipsoids .....................................................................................................169 Figure 6. 5. Depiction of uranyl phosphate layers in UCuP. Uranyl polyhedra are xxi in green and phosphates are in yellow ........................................................171 Figure 6. 6. View of the three-dimensional structure of UCuP wherein the Cu II centers link uranyl phosphate layers together. Uranium polyhedra are in green, phosphate in yellow, and copper in blue. Some of the Cs + sites have been omitted for clarity ..............................................................172 Figure 6. 7. UV-vis-NIR diffuse reflectance spectrum of UCoP. F(R) corresponds to Kubelka-Monk function and relates reflectance data to the absorption spectra. 37 .............................................................................................................................. 174 Figure 6. 8. UV-vis-NIR diffuse reflectance spectrum of UNiP. F(R) corresponds to Kubelka-Monk function and relates reflectance data to the absorption spectra. 37 ....................................................................................175 Figure 6. 9. UV-vis-NIR diffuse reflectance spectrum for UCuP. F(R) corresponds to Kubelka-Monk function and relates reflectance data to the absorption spectra. 37 ....................................................................................177 Figure 6. 10. Magnetization of UCoP as a function of applied magnetic field at 5 K. Line is the linear fit to the data.......................................................178 Figure 6. 11. Magnetic susceptibility of UCoP as a function of temperature. Line is the fit to Curie-Weiss behavior. Inset show the temperature dependence of the inverse magnetic susceptibility with a linear fit. ..........179 Figure 6. 12. Magnetization of UCuP as a function of applied magnetic field at 5 K. Line is the linear fit to the data. ..................................................................180 Figure 6. 13. Magnetic susceptibility of UCuP as a function of temperature. Line is the fit to Curie-Weiss behavior. Inset show the temperature xxii dependence of the inverse magnetic susceptibility.....................................181 Figure 6. 14. DSC data for UCoP ...................................................................................183 Figure 6. 15. DSC data for UCuP ...................................................................................184 xxiii LIST OF TABLES Table 1. 1. The Most Significant Contributors to the Radiation Dose from HLNW........13 Table 1. 2. Crystallographic Data for Na 2 Ti 2 O 3 (SiO 4 )?H 2 O (TiSi-1), HM 3 Ti 4 O 4 (SiO 4 )?H 2 O (TiSi-2) and Na 2 Nb 2-x M x O 6-x (OH) x ?H 2 O (SOMS).........................................................................................................19 Table 1. 3. The Comparative Table of Distribution Coefficients for Sr 2+ , K DSr , of Na 4 Ti 9 O 20 ?nH 2 O, Na 2 Ti 2 O 3 (SiO 4 )?H 2 O, and HK 3 Ti 4 O 4 (SiO 4 )?H 2 O 64 in Different HLNW Simulants......................................................................21 Table 2. 1. Crystallographic Data for K[UO 2 (IO 3 ) 3 ] (KUI)..............................................37 Table 2. 2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for K[UO 2 (IO 3 ) 3 ] (KUI)................................................................................38 Table 2. 3. Selected Bond Distances (?) and Angles (?) for K[UO 2 (IO 3 ) 3 ] (KUI) ..........40 Table 3. 1. Crystallographic Data for Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ]?H 2 O (UGaP1) and Cs[UO 2 Ga(PO 4 ) 2 ] (UGaP2) ...................................................................57 Table 3. 2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ]?H 2 O (UGaP1) .............................................58 Table 3. 3. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs[UO 2 Ga(PO 4 ) 2 ] (UGaP2) ....................................................................60 Table 3. 4. Selected Bond Distances (?) and Angles ( o ) for Cs 4 [(UO 2 ) 2 (GaOH) 2 (PO 4 ) 4 ] (UGaP1) ........................................................................................................64 xxiv Table 3. 5. Selected Bond Distances (?) and Angles ( o ) for Cs[UO 2 Ga(PO 4 ) 2 ] (UGaP2) ........................................................................................................72 Table 4. 1. Crystallographic Data for Cs 2 [UO 2 (VO 2 ) 2 (PO 4 ) 2 ]?nH 2 O (UV V P), K 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (KUV IV P) and Rb 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (RbUV IV P)......................................................89 Table 4. 2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs 2 [UO 2 (VO 2 ) 2 (PO 4 ) 2 ]?nH 2 O (UV V P) ....................................................90 Table 4. 3. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for K 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (KUV IV P).................................91 Table 4. 4. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Rb 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (RbUV IV P)................................................92 Table 4. 5. Selected Bond Distances (?) and Angles (?) for Cs 2 [UO 2 (VO 2 ) 2 (PO 4 ) 2 ]?nH 2 O (UV V P)..........................................................98 Table 4. 6. Selected Bond Distances (?) and Angles (?) for K 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (KUV IV P). Data for Rb 4 (UO 2 )(VO) 4 (OH) 2 (PO 4 ) 4 (RbUV IV P) are given in parathensis.............102 Table 5. 1. Crystallographic Data for Cs 1.52 [(UO 2 )(TiOH 0.48 )(PO 4 ) 2 ]?nH 2 O (UTiP) and Cs 1.35 [(UO 2 )(GeOH 0.65 )(PO 4 ) 2 ]?nH 2 O (UGeP) ........................127 Table 5. 2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs 1.52 [(UO 2 )(TiOH 0.48 )(PO 4 ) 2 ]?nH 2 O (UTiP)........................................128 Table 5. 3. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs 1.35 [(UO 2 )(GeOH 0.65 )(PO 4 ) 2 ]?nH 2 O (UGeP)......................................129 Table 5. 4. Selected Bond Distances (?) and Angles (?) for xxv Cs 1.52 [(UO 2 )(TiOH 0.48 )(PO 4 ) 2 ]?nH 2 O (UTiP)..............................................135 Table 5. 5. Selected Bond Distances (?) and Angles (?) for Cs 1.35 [(UO 2 )(GeOH 0.65 )(PO 4 ) 2 ]?nH 2 O (UGeP) ...........................................136 Table 6. 1. Crystallographic Data for Cs 2 {(UO 2 ) 4 [Co(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 } (UCoP), Cs 2 {(UO 2 ) 4 [Ni(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 } (UNiP) and Cs 3+x [(UO 2 ) 3 CuH 4-x (PO 4 ) 5 ]?H 2 O (UCuP) ...................................................151 Table 6. 2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs 2 {(UO 2 ) 4 [Co(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 } (UCoP)...................................152 Table 6. 3. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs 2 {(UO 2 ) 4 [Ni(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 } (UNiP)....................................154 Table 6. 4. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs 3+x [(UO 2 ) 3 CuH 4-x (PO 4 ) 5 ]?H 2 O (UCuP)..............................................156 Table 6. 5. Selected Bond Distances (?) and Angles (?) for Cs 2 {(UO 2 ) 4 [Co(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 } (UCoP) ........................................166 Table 6. 6. Selected Bond Distances (?) and Angles (?) for Cs 2 {(UO 2 ) 4 [Ni(H 2 O) 2 ] 2 (HPO 4 )(PO 4 ) 4 } (UNiP)..........................................167 Table 6. 7. Selected Bond Distances (?) and Angles (?) for Cs 3+x [(UO 2 ) 3 CuH 4-x (PO 4 ) 5 ]?H 2 O (UCuP) ...................................................170 1 CHAPTER 1 INTRODUCTION The successful test of the first nuclear reactor at the University of Chicago in 19421 changed the course of history. This experiment initiated not only the ?Manhattan Project?, and later the production of nuclear power,2 but it also produced the first portion of highly radioactive nuclear waste. Since that time the search for appropriate long-term storage conditions for high-level nuclear waste (HLNW) and spent nuclear fuel (SNF), as well as investigations into possibilities for the reduction of the total nuclear waste amounts became an issue of world-wide importance.3-6 The detailed knowledge about the fundamental chemistry of actinides is an essential basis for predicting the behavior of waste under different environmental conditions. Interactions of SNF with underground water and formation of the alteration phases under storage conditions6-9, or the mobility of actinides in soils6 are just a couple of selected issues that will be addressed herein. The separation of the major fission products 137Cs and 90Sr from SNF is critical to the proper storage of radioactive waste.3,4 These two radionuclides with half-lives of 30.2 and 28.5 years, respectively, are the source of 99% of the present radioactivity in the HLNW storage tanks. They are distributed between salt cake ? the solid portion of the waste on the top, and liquid portions - hydroxide sludge and alkaline supernatant.3 The removal of Cs+ and Sr2+ will significantly reduce the total amount of nuclear waste. One of the proposed techniques for such removal, ion-exchange, also will allow the transfer of 2 Cs+ and Sr2+ from liquid to the solid phase, and will make the final disposal possible. This method requires the use of layered or framework materials, that should have a high capacity and selectivity for Cs+ and Sr2+, be resistant to elevated temperature and radiation damage, and should be easily synthesized.10 The general purpose of this research project is to expand the existing knowledge of the crystal chemistry of uranium compounds with novel structural topologies, to develop new synthetic methods, and address potential applications of these materials. The other major goal is to synthesize uranium-containing layered or three-dimensional compounds with ion-exchange properties selective for Cs+ or Sr2+. These materials in exchanged form will be able to tightly hold 137Cs or 90Sr, and provide safe and dense storage environments for U and 137Cs or 90Sr simultaneously. CRYSTAL CHEMISTRY OF URANIUM (VI) UNDER HYDROTHERMAL CONDITIONS Uranium is found as a component of approximately 200 mineral crystal structures.11,12 The majority of these minerals are present in the oxidizing zones of U deposits and should be accurately studied in both prospects as deeply stored nuclear waste and as potential products of oxidation of SNF and its interaction with water (alteration phases) in the geological repository locations. Precise investigation of the structures, as well as the stability and properties of uranium compounds synthesized under hydrothermal conditions, will help researchers to understand and predict the behavior of uranium phases in expected geological sites. 3 The term ?hydrothermal synthesis? usually refers to all reactions having water as a solvent carried out at temperatures between the boiling and supercritical points of water and at pressures above 1 bar.13 These temperature and pressure conditions were chosen for our experiments due to similar conditions for the geological formation of minerals. In addition, hydrothermal synthesis potentially allows for the synthesis of compounds with lower densities through the incorporation of water in the structures versus more dense products that are typically isolated from high temperature reactions. In the majority of both uranium-containing solids and solutions that are of an oxidizing nature, U is present in the +6 oxidation state. In order to maintain such a high positive charge, two short U=O bonds in trans arrangement with an average U=O distance of 1.79(3) ? are formed,12 yielding approximately linear [UO2]2+ cation. Because these two oxygen atoms donate their electron density to the uranium center and hence stabilize the high oxidation state of U, they seldom participate in bond formation with other units. According to bond valence sum calculations based on the uranyl- oxygen distances, these two axial oxygen atoms contribute more than half to the valence of the uranium center.14 The concept of bond valence sum can be used in solid state chemistry in order to estimate bond lengths of a metal in a given oxidation state or to calculate the valence of the atom from the known distances.15 Interactions between uranyl and other donor atoms occur in the equatorial plane where the metal center can be coordinated with four, five, or six ligands, with distances ranging from 2.2 ? to 2.6 ?. Thus, traditionally the basic building unit for uranyl- containing compounds can be drawn as tetragonal, pentagonal, or hexagonal bipyramids terminated at each apex by two uranyl oxygen atoms. These three basic uranyl polyhedra 4 can be combined through various modes of corner or edge sharing with each other (Figure 1.1), or by bridging with other building units, to yield a myriad of crystal structure topologies.11,12 Due to the presence of two terminal oxygen atoms, one direction for the connectivity in uranyl bipyramids is eliminated, yielding structures that are generally layered. In fact, more than 65% of all known structures of uranyl containing compounds are sheets.11,12 For the formation of the compounds with three-dimensional structure, where building blocks are interconnected in all three directions, usually, but not always, non-uranyl polyhedra are incorporated into the structure that give additional directions for bond formation. There are several strategies for the building of three-dimensional structures that contain uranium. Very few three-dimensional structures are built from uranyl building units only.16-18 For instance, (C4N2H12)U2O4F6 consists of only one type of building units - UO2F5 pentagonal bipyramids, which are connected by U-F-U bonds.18 This structure also can be described as an open-framework because of the presence of several large channels, extended along different directions that are filled with charge-balancing piperazinium cations (Figure 1.2). The other synthetic approach includes incorporation of octahedra of other metals such as V or W in the structure.19,20 K2[(UO2)2(VO)2(IO6)2O]?H2O (Figure 1.3) is constructed entirely from distorted octahedral building units.19 The UO6 distortion in this structure is associated with the formation of short uranyl bonds and nonequality of equatorial U-O bond lengths, VO6 ? with the presence of short vanadyl V=O bond 5 Figure 1. 1. The UO22+ cation in tetragonal bipyramidal, pentagonal bipyramidal, and hexagonal bipyramidal environments as common building units for uranyl structure formation. They can be interconnected by corner- or edge-sharing with each other or other building units in order to construct different structure types. 6 Figure 1. 2. The structure of (C4N2H12)U2O4F6 composed of only uranyl pentagonal bipyramids (green). Piperazinium cations are located in the channels of the framework. 7 Figure 1. 3. The structure of K2[(UO2)2(VO)2(IO6)2O]?H2O is built from distorted octahedral building units only. UO6 polyhedra are shown in green, VO6 ? in light blue, IO6 ? in red, and K+ cation ? in magenta . K+ cations along with water molecules occupy the tunnels. 8 (1.58(1) ?) and a long bond in the trans-position relative to V=O (2.38(1) ?). IO6 octahedra also show variations in I-O bond length from 1.84(1) ? to 1.93(1) ?. The octahedra create an open-framework structure with channels in two perpendicular directions, both occupied by K+ cations. Li2(UO2)4(WO4)O (Figure 1.4) is an example of a three-dimensional, but non- open-framework structure.20 Very small channels are present only in the [110] direction, and they are not filled with cations. Instead, Li+ cations are located within the voids in the plane of layers. Despite the absence of channels, this material was reported to have high cationic conductivity, comparable with Li -alumina and LiSiCON materials.20 Most of the uranyl-containing three-dimensional materials are built from uranyl units and tetrahedral anions, such as phosphate or arsenate,21-26 sulfate,27 silicate,28-33 vanadate 34,35 or others. These structures can be constructed in many different manners, such as the interconnection of one-dimensional chains or layers by single building blocks, intersection of several one-dimensional chains extended along different axes, or by other methods. We hypothesized that the introduction of octahedral building units containing main group (Ga, In, Ge) or transitional metal (Ti, V, Co Ni, Cu) to uranyl and tetrahedral building blocks might lead to the formation of new types of three-dimensional structures with new topologies. There are 54 framework types known for uranyl-containing compounds today.12 Due to the vast amount of stored depleted uranium with low radioactivity that has to be utilized in the near-term perspective, an extensive search for new practical applications 9 Figure 1. 4. The structure of Li2(UO2)4(WO4)O is an example of a three-dimensional structure with no channels. Li+ cations located in the voids of uranyl tungstate layers are not seen from this projection. UO6 polyhedra are green, WO6 ? shown in violet color. 10 for these structures is being conducted. For example, Na(UO2)4(VO4)3 and Li(UO2)4(VO4)3 can serve as ionic conductors due to the migration of Na+ and Li+ cations in two intersecting channels within the structure.34 Chiral uranyl molybdates [(C2H5)2NH2]2(UO2)4(MoO4)5(H2O)](H2O)36 and (NH4)4[(UO2)5(MoO4)7](H2O)537 (Figure 1.5) both have large channels extended along the [001] direction with a chiral topology as required by space group (P6522 and P61 respectively). In terms of tubular building units, the internal structure of these channels can be described as a double-helix, where both chains form spirals with axes parallel to [001]. Structures with these types of channels can potentially find applications in enantioselective separation and catalysis.37 We proposed that our new uranyl main group or transition metal phosphate materials synthesized within this project can potentially serve as ion-exchange and storage materials for immobilization of radioactive 137Cs and 90Sr from nuclear waste solutions. HIGH-LEVEL NUCLEAR WASTE (HLNW) At the present time, HLNW produced in the processing of uranium and plutonium, is stored in 273 underground carbon-steel tanks.3 This storage strategy needs to be changed due to leakage of radioactive waste into the soil, and the potential long- term impact of this to the environment. The deep geological repository in the Yucca Mountain area, Nevada, was proposed as the safest way for ultimate disposal. 11 Figure 1. 5. Chiral channels extended along the c axis make the uranyl molybdate (NH4)4[(UO2)5(MoO4)7](H2O)5 potentially applicable for enantio-selective separation. NH4+ cations are not shown, UO7 polyhedra are green, MoO4 tetrahedra are blue. 12 It is apparent that the application of geological disposal methods requires pretreatment of highly inhomogeneous waste in order to convert it to a solid form and to separate long-living isotopes pertinent for geological deposition from short-living isotopes and low-activity portions. The main input to the present radioactivity in the tanks is from 137Cs and 90Sr. These isotopes are major dose contributors only during the first 100 years of deposition. (Table 1.1) Their removal from the tanks will significantly reduce the total amount of waste prepared for the geological repository and will make this process more reasonable. Solvent extraction is the method currently used for 137Cs and 90Sr separation.3 90Sr is extracted by 4,4?(5?)-di-(tert-butylcyclohexano)-18-crown-6 in an organic solution of 1.5 M TBP and hydrocarbon diluent called Isopar L (SREX process).3,4 Tests in actual waste containers demonstrated that SREX reduces the activity of waste in the tanks after 90Sr extraction to 0.04 Ci/L, which can be considered as low-level waste activity.3 137Cs can be separated by various crown ethers or cobalt dicarbollide dissolved in a very polar solvent such as nitrobenzene.3,4 This process removes over 99.998% of the 137Cs. In order to co-extract both radionuclides simultaneously, polyethylene glycol can be added to the polar solvent-cobalt dicarbollide system.3 Other techniques such as in-tank precipitation of 137Cs with sodium tetraphenylborate or membrane separation were proposed for removal of 137Cs and 90Sr, but several issues, such as pretreatment of the solutions, capacity of the separation, stability of precipitating agents and membranes in actual waste, and volume dependence still need to be optimized.3 13 Table 1. 1. The most significant contributors to the radiation dose from HLNW.6 Activity (Ci/canister), Time out of Reactor Radionuclide Half-life 10 years 100 years 1000 years 10000 years 90Sr 29 years 1.4?105 1.5?104 3.5?10-6 0 90Y 64 years 1.4?105 1.5?104 3.5?10-6 0 99Tc 2.13?105 years 3.2?101 3.2?101 3.2?101 3.2?101 106Ru 369 days 1.3?103 0 0 0 106Rh 2.18 hours 1.3?103 0 0 0 125Sb 2.73 years 1.6?103 0 0 0 129I 1.59?105 years 1.8?10-4 1.8?10-4 1.8?10-4 1.8?10-4 134Cs 2.06 years 1.9?104 0 0 0 137Cs 30.1 years 2.0?105 2.4?104 2.3?10-5 0 144Ce 284.4 days 3.5?102 0 0 0 147Pm 2.64 years 1.8?104 8.1?10-7 0 0 154Eu 8.6 years 1.0?104 2.1?102 0 0 226Ra 1600 years 2.5?10-7 2.6?10-6 1.6?10-4 6.5?10-3 229Th 7340 years 9.6?10-8 1.7?10-6 1.6?10-4 1.3?10-2 230Th 7.7?104 years 4.9?10-5 7.4?10-5 8.8?10-4 8.1?10-3 213Pa 3.25?104 years 5.7?10-5 5.7?10-5 6.0?10-5 8.8?10-5 233U 1.58?105 years 3.4?10-5 3.4?10-5 3.5?10-3 3.6?10-2 237Np 2.14?106 years 8.1?10-1 8.1?10-1 8.8?10-1 8.8?10-1 238Pu 87.8 years 2.4?102 1.2?102 2.8?10-1 0 239Pu 2.439?104 years 3.8 3.8 4.7 9.6 240Pu 6540 years 1.0?101 2.0?101 1.8?101 7.3 241Pu 15 years 7.4?101 1.1?101 7.2?10-1 3.4?10-1 241Am 433 years 4.1?102 3.8?102 8.1?101 3.4?10-1 243Am 7370 years 4.0?101 4.0?101 3.7?101 1.7?101 242Cm 163 days 1.6?101 1.1?101 1.8?10-1 0 244Cm 17.9 years 4.0?103 1.3?102 0 0 14 Ion-exchange is also considered as a prospective method for the separation of short-lived 137Cs and 90Sr radionuclides. ION-EXCHANGE Ion-exchange is a stoichiometric equilibrium process between the ions of the exchanger and ions of the media in which the exchanger is immersed. In the case of an insoluble anionic matrix, containing exchangeable cations, and a solution that contains cations, available for exchange, the reaction can be written as: ++-++- +fi+ ABMBAM (1) where -M is the negatively charged insoluble matrix, A+ and B+ are exchangeable cations, and barred symbols represent the solid phase.38 The properties of any ion-exchange material can be described in capacity and selectivity terms.39-40 The measurements of uptake values c or distribution coefficients KD can characterize the capacity of ion-exchange material that illustrates the amount of ions that can be replaced in the unit mass of solid. Uptake c may be expressed as: zm VCCc finin )( -= (2) where Cin and Cfin are concentrations of B+ solution before and after ion-exchange, m is mass of the dry sample, z ? charge of the cation and V is aliquot volume of the solution. KD might be calculated as: 15 B B D C CK = (3) where CB is the concentration of B+ in solid phase (barred) and solution, or m V C CCK fin finin D )( -= (4) The selectivity constant describes the affinity of material for exchange of the cation A+ over the B+: B B A A BA C C C CK = / (5) where CA and CB are concentrations of the cations A+ and B+, respectively (barred symbols represent the solids). For organic resins, capacity and selectivity are controlled by synthetic and pretreatment conditions. In general, the resin is based on the hydrocarbon matrix with introduced functional groups, whose nature and strength define the functionality of the material. The water content or swelling of the resin is inversely related to the degree of crosslinking in the hydrocarbon matrix and determines the capacity values.38 Due to the highly reproducible ion-exchange capacities (~2 meq/g of dry sample), and predictable properties, organic resins are one of the best-known methods today for conventional water purification as well as the special ultrapure water pretreatment in pharmaceutics, electronics, and other highly technological industries. It is also widely used in extraction metallurgy, medicine, and agriculture.38 There is no straightforward theory that clarifies the origin of selectivity of inorganic framework ion-exchange materials. Size, charge, and magnitude of hydration 16 energy of exchanged cations, charge distribution of the matrix, structural features, degree of solution loading are just several major issues that should be considered and combined in order to explain ion-exchange properties for every particular material. Arrangements of all factors are able to favor the formation of the inorganic exchange materials with up to 9 meq/g uptake values (Na+ exchange in titanium silicates10) and high selectivity to chosen cations. High resistance to temperature and radiation damage, and low cost of synthesis make inorganic ion-exchange materials highly competitive for the removal of 137Cs and 90Sr isotopes. ION-EXCHANGE LAYERED AND THREE-DIMENSIONAL INORGANIC MATERIALS FOR SELECTIVE Cs+ AND Sr2+ UPTAKE Many layered and framework compounds with remarkable selectivity for Cs+ or Sr2+ were tested under conditions relevant to nuclear waste. Ammonium molybdophosphate (AMP) possesses a high capacity for Cs+, but its application in the highly basic media is limited.41 Potassium cobalt hexacyanoferrate, a three-dimensional compound with a cubic face-centered structure, belonging to the Prussian Blue family, was extensively studied due to the high uptake of Cs+ in the presence of interfering cations over the pH range of 1 to 13.42,43 The framework is built out of iron atoms, that reside in corners and face centers of the unit cell, cobalt atoms, located at the edges, and exchangeable potassium cations occupying the body center. Low thermal stability resulting in decomposition of the sorbent at temperatures above 240 ?C 44 has restricted the application of the material, and has driven the search for alternatives. 17 The structure of sodium nonatitanate Na4Ti9O20 nH2O is still not well-resolved.45 This semicrystalline compound with promising ion-exchange properties has a layered structure with an interlayer distance that depends on the water content, and can be as large as 10.0 ?.46 Batch and column experiments in two waste simulants, that differ in the content of citric acid, Na4EDTA, iminodiacetic acid, sodium gluconate, and other complex-forming agents, demonstrated that sodium nonatitanate is applicable for ion- exchange with an extremely high Kd for Sr2+ of 2.35 ? 105 mL/g at the volume to mass ratio of 200 and 3.96 ? 104 mL/g at volume to mass ratio of 1000 at waste solutions with no formation of complex occurring.47,48 The family of titanium silicates has been considered as the most promising sorbents for Cs+ and Sr2+ removal from nuclear waste for many years. The structure of sodium titanium silicate Na2Ti2O3(SiO4) H2O (commercial product name IONSIV? IE- 911) belongs to the hexagonal space group P42/mcm.3,49,50 (Figure 1.6) The crystallographic data are given in the Table 1.2. It is a framework material constructed from Ti four-member group clusters bridged by silicate groups in the a and b directions, with channels formed along the c axis. In every cluster titanium octahedra are pairwise edge-shared, with pairs rotated by 90?. In addition to tunnels, where four Na+ cations are located, there are small framework cavities located perpendicular to the channels, also occupied by Na+ cations. This material is selective to Cs+ with Kd values ranging from 581 mL/g to 870 mL/g in different simulants.63,64 At a pH>12, measured uptake value for the protonated form is ~1.9 meq/g, which corresponds only to 25% of the total calculated exchange 18 Figure 1.6. A view of the crystal structure of Na2Ti2O3(SiO4) H2O along the [001] direction. Yellow squares depict TiO6 octahedra, orange - SiO4 tetrahedra, green balls ? Na+ cations. Four out of the total eight Na+ cations are located in the channels, the other four (on the silicate groups) ? in small cavities. Copied from reference 50 with author?s permission. 19 Table 1.2. Crystallographic data for Na2Ti2O3(SiO4) H2O (TiSi-1), HM3Ti4O4(SiO4) H2O (TiSi-2) and Na2Nb2-x MxO6-x(OH)x?H2O (SOMS) Compound TiSi-1 TiSi-2 SOMS Crystal System tetragonal cubic monoclinic Space group P42/mcm (No. 132) P 43m (No. 215) C2/c (No.15) a (?) 7.8082(2) 7.7644(3) 16.940(3) b (?) 7.8082(2) 7.7644(3) 5.033(5) c (?) 11.9735(4) 7.7644(3) 16.466(3) a (?) 90 90 90  (?) 90 90 114.00 g (?) 90 90 90 Z 4 1 8 R(F) 0.055 0.055 0.0629 Rw 0.142 0.114 0.1606 20 capacity of the H2Ti2O3(SiO4) H2O. This phenomenon, as well as preferred selectivity to Cs+ relative to other alkaline cations, can be easily explained using coordination distances and ionic radius data for Cs+: only two out of four exchanged cations can occupy the tunnel simultaneously, but bond distances of coordinated Cs+ are very close to the sum of Cs+ and O2- ionic radii. At the same time Na+ can compete for ion-exchange selectivity in the mixed solutions, occupying framework cavity sites.48 For Sr2+ cation exchange, sodium titanium silicate possesses the highest observed in this family of the compounds KD values of ~3.0 ? 105 mL/g. This high selectivity might be ascribed to the presence of Sr2+ as Sr(OH)+ in the strongly basic conditions of nuclear waste and possibility of Sr(OH)+ occupying both tunnel and cavity exchange sites. The other member of the titanium silicate family, a compound with the structure of mineral pharmacosiderite type HM3Ti4O4(SiO4) H2O (M=K, Na), has a framework built similarly to that described above.10,50 All differences between these two related structures, and their ion-exchange properties can be ascribed to the cubic nature of pharmacosiderite and its space group. (Table 1.2) Titanium clusters are connected now in all three directions equivalently, which produces a system of perpendicular channels, parallel to all three axes. Three potential ion-exchange sites are placed in the cube face centers and one in the cubic body center of the unit cell. After Cs+ exchange, three face sites can be selectively filled with Cs+ cations and the body site remains in the protonated form (or occupied by smaller cations) due to the restricted space requirements. KDCs in alkaline solutions is 15360 mL/g, which can be compared with KDRb - 6190 mL/g, KDK - 2580 mL/g, and KDLi ? 630 mL/g. The distribution coefficients are summarized in the 21 Table 1.3. Distribution coefficients for Sr2+, KDSr, of Na4Ti9O20 nH2O, Na2Ti2O3(SiO4) H2O, and HK3Ti4O4(SiO4) H2O64 in different HLNW simulants. KDSr, mL/g NCAW simulant * 101SY-Cs5 simulant ** Compound V:m = 200 V:m = 1000 V:m = 200 Na4Ti9O20 nH2O 2.35 ? 105 3.96 ? 104 295 Na2Ti2O3(SiO4) H2O 2.70 ? 105 2.26 ? 105 231 HK3Ti4O4(SiO4) H2O 2.02 ? 105 5.00 ? 103 31 * The composition of NCAW is restricted only to presence of 0.43 M Al3+, 5 ? 10-4 M Cs+, 0.12 M K+, 4.99 M Na+, 5 ? 10-5 M Rb+, 2.7 ? 10-7 Sr2+, 0.23 M carbonate, 0.09 M fluoride, 3.4 M hydroxide, 1.67 M nitrate, 0.43 M nitrite, 0.15 M sulfate, and 0.025 M phosphate. ** In addition in 101SY-Cs5 present 4.2 ? 10-3 M Ca2+, 1.96 ? 10-4 M Fe3+, 4.2 ? 10-4 M Mo4+, 2.5 ? 10-4 M Ni2+, 5.0 ? 10-4 M Zn2+, 5.0 ? 10-3 M citric acid, 5.0 ? 10-3 M Na4EDTA, 0.031 M iminodiacetic acid, 2.5?10-4 M Na3 nitrilotriacetate, 0.013 M sodium gluconate. 22 partially explain the preference of the structure for Cs+ in terms of its hydration energy, which is not high enough to deny the introduction of cations into the relatively small channels.10,50 Reported Sr2+ selectivity for pharmacosiderite structure in HLNW simulants with different pH values and Cs+ content is lower than for other titanium silicates and titanates.47,48 The ion-exchange data were also measured in reference to the degree of the crystallinity of the material. Results revealed participation of both absorption and ion-exchange mechanisms in the separation of Sr2+. Recently Sandia National Laboratory invented a new series of compounds named Sandia Octahedral Molecular Sieves (SOMS).51-54 They are sodium niobate-based framework materials with the general formula Na2Nb2-xMxO6-x(OH)x?H2O, where M could be Ti or Zr with x < 0.4. (Table 1.2) Substitution of pentavalent Nb coordination octahedra with tetravalent Ti or Zr coordination octahedra plays the same role as a substitution of Si with Al in zeolite structures. It creates an anionic charge within the matrix that must be balanced. The overall structure consists of double chains of edge- shared octahedra of NbO6 and TiO6 (ZrO6) extending along the b axis. These chains are intersected with layers of edge-shared NaO6 octahedra and this structural interaction forms a negatively charged framework. The fragment of the structure is depicted in the Figure 1.7. There are two different sorts of Na+ cations in this compound: framework Na octahedra that participate in the matrix formation and Na+ counter cations located in the channels that balance negative charge of the matrix. These latter cations are exchangeable. 23 Figure 1.7. This fragment depicts interconnection of octahedral units in the structure of Na2Nb2-xTixO6-x(OH)x?H2O. Blue polyhedra represent disordered NbO6 and TiO6 octahedra. Yellow polyhedra are Na coordination spheres, participating in the framework. Copied from reference 52 with author?s permission. 24 SOMS materials demonstrate exceptional selectivity toward all alkaline-earth cations relative to alkali metal cations For instance, the KD value of SOMS-1 (Nb:Ti ratio is 4:1) in Na form for Ba2+, Sr2+, and Mg2+ are all above 99800 mL/g , but the KD values for Li+, K+, and Cs+ do not exceed 150 mL/g. Upon heating they have one- step phase transformation to perovskite type sodium niobate NaNbO3 without any intermediate stages like amorphization or decomposition. Measured enthalpies of formation of hydrated, dehydrated, and dense forms of materials proved that dense perovskite structure is more thermodynamically favorable and may serve as a permanent ceramic host phase for the storage of 90Sr.51-54 MAIN GROUP AND TRANSITION METAL PHOSPHATES There have been countless remarkable catalysts,55-58 ion-exchangers,59,60 conductors60 and many other materials with other applications in the area of various main group and transition metal phosphates.61 Along with useful properties, they display extraordinary structural diversity. For instance, among aluminum phosphates alone there are more than 40 different framework types. One eminent example is aluminophosphate VPI-5, which contains 18-membered ring channels with a diameter of 13 ?.62 Vanadium phosphates, including (VO)2P2O7, NaVOPO4, CaV2(P2O7),63 VOPO4? 2H2O, VOHPO4?0.5H2O, VO(H2PO4)256 and others, represent, probably, the largest group of catalysts within one class of compounds. Vanadyl pyrophosphate (VO)2P2O7, for example, is currently used in industry for the oxidation of butane to maleic anhydride. Its structure consist of the double chains of edge-sharing VO6 octahedra, interconnected by pyrophosphate groups into the framework.64 These pairs of VO6 octahedra are associated 25 with active sites, where dehydrogenation of butane occurs.65 Other authors66 address the high oxidation activity of (VO)2P2O7 and other vanadium phosphate catalysts to the presence of V=O double bonds at the active VO6 sites and the ability of oxygen from V=O to attack the reactants leading to the oxidative hydrogenation reaction. In contrast, the group of nickel phosphates illustrates the extraordinary flexibility of their properties in this small group of compounds. Two 24-membered ring framework compounds Ni18(HPO4)14(OH)3F9(H3O+,NH4+)4?12H2O (VSB-1)57 and Ni20[(OH)12(H2O)6][(HPO4)8(PO4)4]?12H2O (VSB-5)58 demonstrate a variety of properties. VSB-1 exhibits ion-exchange properties, antiferromagnetic ordering at 10 K and thermal stability up to temperatures above 550 ?C, as well as shape-selective catalytic activity in the hydroxylation of phenol.67 VSB-5 also possesses good temperature resistance and antiferromagnetic ordering.58 It also might serve as a hydrogenation catalyst upon calcination due to the availability of Ni sites, and as hydrogen storage material.68 The ion-exchange properties of layered transition metal phosphates Ti(HPO4)2?H2O, Zr(HPO4)2?H2O, Hf(HPO4)2?H2O, Ge(HPO4)2?H2O, Sn(HPO4)2?H2O, and Pb(HPO4)2?H2O are well characterized.59 The interlayer spacing, ranging between 7.56 ? and 7.8 ?, is confined by the oxygen atoms of the phosphate groups, which could be protonated. As a consequence, the ion-exchange properties of these compounds depend strongly on the pH, and reveal the maximal ion-exchange capacities between 4.17 meq/g for Hf(HPO4)2?H2O and 7.76 meq/g for Ti(HPO4)2?H2O in strongly basic conditions. The other titanium phosphate, potassium titanyl phosphate KTiOPO4 (KTP) is one of the best known nonlinear optical materials.69 The cause of frequency doubling 26 properties lies in the overall polarity of the structure, consisting of chains of highly distorted corner-sharing TiO6 octahedra, connected by phosphate tetrahedra. As a summary of this short review and many other main group or transition metal phosphates not mentioned above, one can conclude that the presence of metal centers in these compounds are the source of their diverse properties. We anticipate that the addition of the main group or transition metal centers in uranyl phosphate structures could also enrich their properties and could make uranyl-containing materials potentially applicable. 27 REFERENCES 1. Fermi, E. Proceedings of the American Philosophy Society, 1946, 90, 20. 2. Rhodes, R. The Making the Atomic Bomb; Simon and Schuster, 1986. 3. Schulz, W. W.; Lombardo, N. J. Science and Technology for Disposal of Radioactive Tank Wastes, Proceedings of the ACS Symposium on Science and Technology for Disposal of Radioactive Tank Wastes, September, 7-11, Las Vegas, Nevada, 1997. 4. Disposal of Radioactive Waste, Proceedings of the Information Meeting NEA, April, 12-14, Paris, 1972. 5. Alternative High-Level Waste Treatments at the Idaho National Engineering and Environmental Laboratory, National Academy Press, 1999. 6. Brookins, D. G. Geochemical Aspects of Radioactive Waste Disposal, Springer- Verlag: New-York, 1984. 7. Finch, R. J.; Buck, E. C.; Finn P. A.; Bates, J. K. in Scientific Basis for Nuclear Waste Management XXII, Materials Research Society Symposium Proceedings, 1999, 556, 431. 8. Finn, P. A.; Hoh, J. C.; Wolf, S. F.; Slater, S. A.; Bates, J. K. Radiochim. Acta, 1996, 74, 65. 9. Wronkiewicz, D. J.; Bates, J. K.; Gerding, T. J.; Veleckis, E.; Tani, B. S. J. Nucl. Mater. 1992, 190, 107. 10. Clearfield, A. ?Inorganic Ion Exchangers, Past, Present and Future? in Solvent Extraction and Ion Exchange, 2000, 18, 655. 11. Burns, P. C.; Miller, M. L.; Ewing, R. C. Can. Mineral. 1996, 34, 845. 28 12. Burns, P. C. Can. Mineral. 2005, 43, 1839. 13. Rabenau, A. Angew. Chem. 1985, 97, 1017. 14. Burns, P. C.; Ewing, R. C.; Hawthorne, F. C. Can. Mineral. 1997, 35, 1551. 15. Brese, N. E.; O?Keeffe, M. Acta Cryst. 1991, B47, 192. 16. Li, Y.; Cahill, C. L.; Burns, P. Chem. Mater. 2001, 13, 4026. 17. Li, Y.; Burns, P. Can. Mineral. 2000, 38, 1433. 18. Halasyamani, P. S.; Walker, S. M.; O'Hare, D. J. Am. Chem. Soc. 1999, 121, 7415. 19. Sykora, R. E.; Albrecht-Schmitt, T. E. Inorg. Chem. 2003, 42, 2179. 20. Obbade, S.; Yagoubi, S.; Dion, C.; Saadi, M.; Abraham, F. J. Solid State Chem. 2004, 177, 1681. 21. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2002, 163, 275. 22. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2002, 167, 226. 23. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2003, 175, 372. 24. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2003, 176, 18. 25. Burns, P. C.; Alexopoulos, C. M.; Hotchkiss, P. J.; Locock, A. J. Inorg. Chem. 2004, 43, 1816. 26. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2004, 177, 2675. 27. Doran, M.; Norquist, A. J.; O'Hare, D. Chem. Commun. (Cambridge) 2002, 24, 2946. 28. Li, Y.; Burns, P. C. J. Nucl. Mater. 2001, 299, 219. 29. Wang, X.; Huang, J.; Liu, L.; Jacobson, A. J. J. Mater. Chem. 2002, 12, 406. 30. Wang, X.; Huang, J.; Jacobson, A. J. J. Amer. Chem. Soc. 2002, 124, 15190. 31. Huang, J.; Wang, X.; Jacobson, A. J. J. Mater. Chem. 2003, 13, 191. 29 32. Chen, C.-S.; Kao, H.-M.; Lii, K.-H. Inorg. Chem. 2005, 44, 935. 33. Chen, C.-S.; Chiang, R.-K.; Kao, H.-M.; Lii, K.-H. Inorg. Chem. 2005, 44, 3914. 34. Obbade, S.; Dion, C.; Rivenet, M.; Saadi, M.; Abraham, F. J. Solid State Chem. 2004, 177, 2058. 35. Obbade, S.; Dion, C.; Saadi, M.; Yagoubi, S.; Abraham, F. J. Solid State Chem. 2004, 177, 3909. 36. Krivovichev, S. V.; Cahill, C. L.; Nazarchuk, E. V.; Burns, P. C.; Armbruster, Th.; Depmeier, W. Micropor. Mesopor. Mater. 2005, 78, 209. 37. Krivovichev, S. V.; Cahill, C. L.; Burns, P. C. Inorg. Chem. 2003, 42, 2459. 38. Harland, C. E. Ion Exchange: Theory and Practice, 2nd ed.; Royal Society of Chemistry, 1994. 39. Helfferich, F. C. Ion-Exchange, Dover Science, 1995. 40. Clearfield, A. Inorganic Ion Exchange Materials, CRC Press, 1982. 41. Todd, T. A.; Mann, N. R.; Tranter, T. J.; Sebesta, F.; John, J.; Motl, A. J. Radioanal. Nucl. Chem. 2002, 254, 47. 42. Harjula, R.; Lehto, J.; Wallace, J. Waste Management, Tucson, Arizona, 1987, 93. 43. Lehto, J.; Paajanen, A.; Harjula, R. J. Radioanal. Nucl. Chem. 1992, 164, 39. 44. Lehto, J.; Pettersson, M.; Hinkula, J.; Raesaenen, M.; Elomaa, M. Thermochim. Acta 1995, 265, 25. 45. Lehto, J.; Clearfield, A. J. Radioanal. Nucl. Chem. Lett. 1987, 118, 1. 46. Clearfield A.; Lehto, J. J. Solid State Chem. 1988, 73, 98. 47. Behrens, E. A.; Sylvester, P.; Clearfield, A. Env. Sci. Technol. 1998, 32, 101. 30 48. Sylvester, P.; Behrens, E. A.; Graziano, G. M.; Clearfield, A. Sep. Sci. Technol. 1999, 34, 1981. 49. Poojary, D. M.; Cahill, R. A.; Clearfield, A. Chem. Mater. 1994, 6, 2364. 50. Clearfield, A. Solid State Sci. 2001, 3, 103. 51. Nyman, M.; Tripathi, A.; Parise, J. B.; Maxwell, R. S.; Harrison, W. T. A.; Nenoff, T. M. J. Am. Chem. Soc. 2001, 123, 1529. 52. Nyman, M.; Tripathi, A.; Parise, J. B.; Maxwell, R. S.; Nenoff, T. M. J. Am. Chem. Soc. 2002, 124, 1704. 53. Xu, H.; Nyman, M.; Nenoff, T. M.; Navrotsky, A. Chem. Mater. 2004, 16, 2034. 54. Nenoff, T. M.; Pless, J. D.; Michaels, E.; Phillips, M. L. F. Chem. Mater. 2005, 17, 950. 55. Bordes, E.; Courtine, P. J. Catal. 1979, 57, 236. 56. Hutchings, G. J. J. Mater. Chem. 2004, 14, 3385. 57. Guillou, N.; Gao, Q.; Nogu?s, M.; Morris, R. E.; Hervieu, M.; F?rey, G.; Cheetham, A. K. C. R. Acad. Sci. Paris, 1999, t.2, Serie II, 387. 58. Guillou, N.; Gao, Q.; Forster, P. M.; Chang, J.- S.; Nogu?s, M.; Park, S.- E.; F?rey, G.; Cheetham, A. K. Angew. Chem., Int. Ed. Engl. 2001, 40, 15. 59. Clearfield, A. Ann. Rev. Mater. Sci. 1984, 14, 205 60. Clearfield, A. Chem. Rev. 1988, 88, 125. 61. Cheetham, A. K.; F?rey, G.; Loiseau, T. Angew. Chem., Int. Ed. Engl. 1999, 38, 3268. 62. Davis, M. E.; Saldarriaga, C.; Montes, C.; Garces, J. M.; Crowder, C. Nature 1988, 331, 698. 31 63. Boudin, S.; Guesdon, A.; Leclaire, A.; Borel, M.-M. Int. J. Inorg. Mater. 2000, 2, 561. 64. Torardi C. C.; Calabrese, J. C. Inorg. Chem. 1984, 23, 1308. 65. Bordes, E.; Courtine, P. J. Chem. Soc., Chem. Commun. 1985, 5, 294. 66. Ai, M. Catalysis Today 2003, 85, 193. 67. Chang, J.- S.; Hwang, J.- S.; Jhung, S. H.; Park, S. - E.; F?rey, G.; Cheetham, A. K. Angew. Chem., Int. Ed. Engl. 2004, 43, 2819. 68. Forster, P. M.; Juergen, E.; Chang, J.- S.; Park, S. - E.; F?rey, G.; Cheetham, A. K. J. Am. Chem. Soc. 2003, 125, 1309. 69. Wang, J.; Wei, J.; Liu, Y.; Yin, X.; Hu, X.; Shao, Z.; Jiang, M. Progr. Cryst. Grow. Character. Mater. 2000, 40, 3. 32 CHAPTER 2 CRYSTAL CHEMISTRY AND ION-EXCHANGE PROPERTIES OF THE LAYERED POTASSIUM URANYL IODATE, K[UO 2 (IO 3 ) 3 ] ABSTRACT Single crystals of potassium uranyl iodate, K[UO 2 (IO 3 ) 3 ] (KUI), were grown under mild hydrothermal conditions. The structure of KUI contains two-dimensional 2 ? [UO 2 (IO 3 ) 3 ] 1? sheets extending in the [ab] plane that consist of approximately linear UO 2 2+ cations bound by iodate anions to yield UO 7 pentagonal bipyramids. There are three crystallographically unique iodate anions, two of which bridge between uranyl cations to create sheets, and one that is monodentate and protrudes in between the layers in cavities. K + cations have long ionic contacts with oxygen atoms from the layers forming an eight-coordinate distorted dodecahedral geometry. These cations join the 2 ? [UO 2 (IO 3 ) 3 ] 1? sheets together. Ion-exchange reactions were carried out that indicate the selective uptake of Cs + over Na + or K + by KUI. Crystallographic data (193 K, MoK?, ? = 0.71073 ?): KUI, orthorhombic, Pbca, a = 11.495(1) ?, b = 7.2293(7) ?, c = 25.394(2) ?, Z = 8, R(F) = 1.95% for 146 parameters with 2619 reflections with I > 2?(I). 33 INTRODUCTION Uranyl iodates have been known since at least the early 1900s with the correct composition of UO 2 (IO 3 ) 2 (H 2 O) being reported in 1913. 1 The precipitation of uranyl iodate as its hydrate has been used as a method of quantifying uranium for almost fifty years. 2,3 The speciation of uranyl iodates in solution and subsequent precipitation has also been implicated in the fate of 129 I (t 1/2 = 1.7 x 10 7 y), a key fission product in spent nuclear fuel. 4 UO 2 (IO 3 ) 2 (H 2 O) was also used as an early probe of the effects of radiation damage on uranium-bearing materials. 5 Despite the importance of iodate in uranyl chemistry, the solid-state structure of a uranyl iodate was not known until Weigel and Engelhardt reported that UO 2 (IO 3 ) 2 (H 2 O) was isostructural with UO 2 (BrO 3 ) 2 (H 2 O), the latter structure being identified by X-ray powder diffraction methods. 6 In recent years uranyl iodates have been a rapidly expanding group of compounds prepared via hydrothermal conditions, whose structures are dominated by the formation of one-dimensional topologies. These one-dimensional structures are known for relatively simple compounds such as UO 2 (IO 3 ) 2 , 7 as well as for materials that incorporate a diverse set of ligands e.g. Cs 2 [(UO 2 ) 3 Cl 2 (IO 3 )(OH)O 2 ]?2H 2 O, 8 Rb[(UO 2 )(CrO 4 )(IO 3 )(H 2 O)], 9 and A 2 [(UO 2 )(MO 4 )(IO 3 ) 2 ] (A = K, Rb, Cs; M = Cr, Mo). 9,10 Unidimensional features occur in additional uranyl iodates that incorporate ? 3 - oxo atoms such as A 2 [(UO 2 ) 3 (IO 3 ) 4 O 2 ] (A = K, Rb, Tl) and AE[(UO 2 ) 2 (IO 3 ) 2 O 2 ] (AE = Sr, Ba, Pb), 11,12 and in complex low-symmetry compounds like K 3 [(UO 2 ) 2 (IO 3 ) 6 ](IO 3 )?H 2 O. 13 Two-dimensional structures are less common, but are known for UO 2 (IO 3 ) 2 (H 2 O) 6,7 and Ag 4 (UO 2 ) 4 (IO 3 ) 2 (IO 4 ) 2 O 2 . 14 One of the key features of the uranyl iodate system is the presence of a stereochemically active lone-pair of 34 electrons on the I(V) center. The nonbonding nature of these electron pairs can cause the loss of a second dimension of connectivity, thereby leading to the formation of one- dimensional topologies that are in some cases seemingly excised from previously known two-dimensional compounds. 11,12 These lone-pairs also have the ability to align in the solid-state to create polar structures, albeit this is rare in uranyl compounds because the approximately linear uranyl cations often reside on, or are related by, centers of inversion or higher symmetry. In fact, all uranyl iodates known to date are centrosymmetric. If the discussion is expanded to include compounds containing NpO 2 + , NpO 2 2+ , and PuO 2 2+ , then three actinyl iodates are known to form polar structures, NpO 2 (IO 3 ), 18 NpO 2 (IO 3 ) 2 ?H 2 O, 19 and PuO 2 (IO 3 ) 2 ?H 2 O, 20 all of which crystallize in the polar space group Pna2 1 . If the composition is allowed to diverge further by considering all uranyl compounds with anions containing nonbonding electrons, then a few other acentric structures can be identified for PbUO 2 (SeO 3 ) 2 21 and Na 8 (UO 2 ) 6 (TeO 3 ) 10 . 22 Members of the uranyl family of compounds with oxoanions of Sb(III) and Bi(III) are thus far found to be centrosymmetric. 23,24 Channels can also be created to house the lone-pair of electrons in these types of compounds, which occurs in ?-AgNpO 2 (SeO 3 ). 18 Herein we report the preparation, crystal structure, and ion-exchange properties of a new layered uranyl iodate, K[UO 2 (IO 3 ) 3 ] (KUI). EXPERIMENTAL Syntheses. UO 3 (99.8%, Alfa-Aesar), H 5 IO 6 (99%, Alfa-Aesar), NaOH (97%, Aldrich), NaCl (99.99%, Aldrich), KOH (85.8%, Fisher), KCl (99%, Aldrich), K 2 CO 3 (99%, Aldrich), CsCl (99.9%, Aldrich), CsOH (99.9%, Aldrich) and V 2 O 5 (99.95%, Alfa- 35 Aesar) were used as received. Distilled and Millipore filtered water with a resistance of 18.2 M??cm was used in all reactions. Reactions were run in Parr 4749 23-mL autoclaves with PTFE liners. Accumet Basic AB 15 pH-meter was used for potentiometric measurements. SEM/electron dispersive analysis by X-ray (EDX) analyses were performed using a JEOL 840/Link Isis instrument. K, U, and I percentages were calibrated against standards. Typical results were surprisingly good and are within 1% of ratios determined from single crystal X-ray diffraction. Warning: While the UO 3 contains depleted U, standard precautions for handling radioactive materials should be followed. K[UO 2 (IO 3 ) 3 ]. UO 3 (0.269 g, 0.940 mmol), K 2 CO 3 (0.130 g, 0.940 mmol), V 2 O 5 (0.171 g, 0.940 mmol), H 5 IO 6 (0.429 g, 1.88 mmol), and 1.5 mL of deionized water were loaded into the 23 mL autoclave. The pH of the initial mixture was 1.1. The autoclave was heated to 180 ?C in a box furnace for 72 h, and cooled at a rate of 9 ?C/h to room temperature. The product mixture consisted of a colorless solution with a pH of 2.9 over a bright yellow precipitate containing tablets of KUI. The product was washed with methanol and allowed to dry. Crystals were separated from the mixture for further study. Yield 0.26 g (33% yield based on the U). EDX analysis provided K:U:I ratio of 1:1:3 (21%:19%:60%). IR (KBr, cm ?1 ): 902 (? 3, UO 2 2+ ), 857 (? 1 , UO 2 2+ ), 832 (?, IO 3 ? ), 817 (?, IO 3 ? ), 808 (?, IO 3 ? ), 786 (?, IO 3 ? ), 763 (?, IO 3 ? ), 733 (?, IO 3 ? ), 525 (?, IO 3 ? ). Crystallographic Studies. A tablet of KUI with dimensions of 0.133 x 0.046 x 0.013 mm was mounted on a thin glass fiber with epoxy, secured on a goniometer head, cooled to ?80 ?C with an Oxford Cryostat, and optically aligned on a Bruker SMART APEX CCD X-ray diffractometer using a digital camera. Intensity measurements were 36 performed using graphite monochromated Mo-K? radiation from a sealed tube with a monocapillary collimator. SMART was used for preliminary determination of the cell parameters and data collection control. The intensities of reflections of a sphere were collected by a combination of 3 sets of exposures (frames). Each set had a different ? angle for the crystal and each exposure covered a range of 0.3? in ?. A total of 1800 frames were collected with an exposure time per frame of 30 s. Determination of integrated intensities and global cell refinement were performed with the Bruker SAINT (v 6.02) software package using a narrow-frame integration algorithm. An analytical absorption correction 25 was applied followed by a semi- empirical absorption correction using SADABS. 26 The program suite SHELXTL (v 6.12) was used for space group determination (XPREP), structure solution (XS), and refinement (XL). 25 The final refinement included anisotropic displacement parameters for all atoms and a secondary extinction parameter. Additional crystallographic details are listed in Table 2.1. Atomic coordinates and equivalent isotropic displacement parameters are given in Table 2.2. Ion-exchange studies. 0.01 M NaCl, 0.01 M KCl, and 0.01 M CsCl solutions were used for the study of the ion-exchange properties of KUI for Na + , K + , and Cs + ions, respectively. All samples were equilibrated in the solutions for three days The concentrations of the solutions before and after uptake were measured with ion-selective electrodes. 27 The mass to volume ratio was 1:1000 (0.01 g for 10 ml of solution). 37 Table 2.1. Crystallographic data for K[UO 2 (IO 3 ) 3 ] (KUI). Compound KUI Formula mass 833.83 Color and habit yellow tablet Crystal system orthorhombic Space group Pbca (No. 61) a (?) 11.495(1) b (?) 7.2293(7) c (?) 25.394(2) ? ( o ) 90 ? ( o ) 90 ? ( o ) 90 V (? 3 ) 2110.3(3) Z 8 T (?C) 193 K ? (?) 0.71073 2? max 56.64 ? calcd (g cm ?3 ) 5.249 ?(Mo K?) (cm ?1 ) 246.03 R(F) for F o 2 > 2?(F o 2 ) a 0.0195 R w (F o 2 ) b 0.0458 a () RF F F F=??? oc o . b () ( ) RF wF F wF wo 2 o 2 c 2 2 o 4 12 =? ? ? ? ? ? ? ?? ? ? ? ? ? ? . 38 Table 2.2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for K[UO 2 (IO 3 ) 3 ] (KUI). Atom Symmetry a x y z U eq (? 2 ) a Occup. U(1) 8 c 1 0.0255(1) 0.0351(1) 0.6519(1) 0.009(1) 1 I(1) 8 c 1 0.0654(1) 0.4563(1) 0.5630(1) 0.013(1) 1 I(2) 8 c 1 -0.2992(1) 0.1366(1) 0.6557(1) 0.011(1) 1 I(3) 8 c 1 0.3409(1) 0.0593(1) 0.7031(1) 0.010(1) 1 K(1) 8 c 1 0.1642(1) -0.9480(2) 0.4781(1) 0.026(1) 1 O(1) 8 c 1 0.0605(3) 0.3375(4) 0.6280(1) 0.012(1) 1 O(2) 8 c 1 -0.0357(3) 0.3232(5) 0.5246(1) 0.018(1) 1 O(3) 8 c 1 0.1980(3) 0.3599(5) 0.5391(1) 0.021(1) 1 O(4) 8 c 1 -0.1523(3) 0.2142(5) 0.6423(1) 0.016(1) 1 O(5) 8 c 1 -0.3464(3) 0.1100(5) 0.5888(1) 0.019(1) 1 O(6) 8 c 1 -0.3497(3) 0.3740(4) 0.6669(2) 0.017(1) 1 O(7) 8 c 1 0.2277(3) 0.0367(5) 0.6539(1) 0.014(1) 1 O(8) 8 c 1 0.2723(3) 0.2224(5) 0.7463(1) 0.016(1) 1 O(9) 8 c 1 0.4362(3) 0.2138(4) 0.6671(1) 0.014(1) 1 O(10) 8 c 1 0.0261(3) 0.0927(5) 0.7202(1) 0.015(1) 1 O(11) 8 c 1 0.0272(3) -0.0278(4) 0.5842(1) 0.014(1) 1 a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b U eq is defined as one-third of the trace of the orthogonalized U ij tensor. 39 RESULTS AND DISCUSSION Syntheses. The reaction of UO 3 with KIO 4 and V 2 O 5 under mild basic hydrothermal conditions has been previously shown to result in the formation of K 2 [(UO 2 ) 2 (VO) 2 (IO 6 ) 2 O]?H 2 O in high yield. 28 In an effort to explore both the compositional space of this reaction and the effects of pH on product formation we have continued to investigate this reaction. In acidic conditions K[UO 2 (IO 3 ) 3 ] (KUI) is formed in moderate yield from similar reactants (K 2 CO 3 and H 5 IO 6 instead of KIO 4 ) that yielded K 2 [(UO 2 ) 2 (VO) 2 (IO 6 ) 2 O]?H 2 O. Structure. K[UO 2 (IO 3 ) 3 ]. The structure of KUI consists of an approximately linear UO 2 2+ cation with U=O distances of 1.777(3) and 1.785(3) ? and an O=U=O angle of 178.39(15)?. (Table 2.3) This moiety is bound by five iodate anions to form a pentagonal bipyramidal coordination environment around the uranium center as is depicted in Figure 2.1. A similar building unit also occurs in one-dimensional K 3 [(UO 2 ) 2 (IO 3 ) 6 ](IO 3 )?H 2 O, 13 although the overall connectivity and dimensionality is different from KUI. The U?O bond distances in the equatorial plane range from 2.304(3) to 2.432(3) ?. These oxygen atoms are approximately planar and deviate from planarity by only approximately 0.08 ?. There are three crystallographically unique iodate anions in KUI. Two of these anions bridge between two uranyl cations and have one terminal oxygen atom. The bridging versus terminal I?O bond distances can be differentiated with bridging distances occurring from 1.811(3) to 1.834(3) ?, whereas the terminal I?O distances are both shorter at 1.793(3) ?. The third iodate anion is monodentate and has a bridging I?O distance of 1.862(3) ?, and two terminal distances of 1.782(3) and 1.796(3) ?. 40 Table 2.3. Selected Bond Distances (?) and Angles (?) for K[UO 2 (IO 3 ) 3 ] (KUI). Bond Lengths (?) U(1)?O(1) 2.304(3) I(1)?O(2) 1.796(3) U(1)?O(4) 2.432(3) I(1)?O(3) 1.782(3) U(1)?O(6) 2.363(3) I(2)?O(4) 1.812(3) U(1)?O(7) 2.325(3) I(2)?O(5) 1.793(3) U(1)?O(9) 2.396(3) I(2)?O(6) 1.834(3) U(1)?O(10) 1.785(3) I(3)?O(7) 1.811(3) U(1)?O(11) 1.777(3) I(3)?O(8) 1.793(3) I(1)?O(1) 1.862(3) I(3)?O(9) 1.812(3) Angles (?) O(10)?U(1)?O(11) 178.39(15) O(5)?I(2)?O(6) 98.72(16) O(3)?I(1)?O(11) 99.18(17) O(4)?I(2)?O(6) 91.97(15) O(3)?U(1)?O(1) 98.46(15) O(8)?I(3)?O(7) 99.50(15) O(2)?I(1)?O(1) 102.40(15) O(8)?I(3)?O(9) 99.77(15) O(5)?I(2)?O(4) 97.84(16) O(7)?I(3)?O(9) 98.19(15) 41 Figure 2.1. A view of the pentagonal bipyramidal coordination environment around uranium center in KUI that is formed from a UO 2 2+ linear cation connected with five iodate groups through the equatorial plane. 42 The bridging of the uranyl cations by the iodate anions creates two-dimensional sheets in the [ab] plane as shown in Figure 2.2. From this perspective it appears that the monodentate iodate anions reside in cavities within the layers. In fact, a view along the b axis (Figure 1.3) shows that these iodate anions protrudes in between the layers and provides their terminal oxo atoms for forming interactions with the K + cations. The interactions between the K + cations and iodate anions stitch the structure together. The K + cations form long interactions with the terminal oxo atoms from both the monodentate iodate group as well as with the terminal oxo atoms from one of the bridging iodate anions. These cations also interact with the oxo atoms from the uranyl cations forming a surprisingly short contact of 2.715(3) ?. 34-39 A total of eight K + ???O contacts are formed, with distances ranging from 2.615(4) to 3.175(4) ?, creating a distorted dodecahedral environment around the potassium centers. Ion-exchange Studies. The exchange of interlayer cations has been previously achieved in a number of layered uranyl compounds, most notably in the autunite and meta-autunite group. 40-42 In addition, certain uranyl compounds such as hydrogen uranyl phosphate (HUP) 43-46 and Cs 4 [(UO 2 ) 2 (V 2 O 7 )O 2 ] 47 have been shown to exhibit high cationic conductivities. In order to evaluate the potential for ion-exchange reactions of alkali metal cations with KUI, samples were equilibrated with 0.01 M solutions of NaCl, KCl, or CsCl. The potential ion exchange of K + with Na + or Cs + was measured potentiometrically using ion-selective electrodes. In order to distinguish between absorption and actual ion exchange, the release of K + was also quantified. No measurable exchange was noted with Na + . Similarly, there was also no change noted in 43 Figure 2.2. An illustration of the 2 ? [UO 2 (IO 3 ) 3 ] 1? sheets in KUI that extend in the [ab] plane. 44 Figure 2.3. A view along the b axis of KUI showing that the monodentate iodate anion protrudes in between the layers, providing its terminal oxo atoms for forming interactions with the K + cations that reside between the layers. 45 the concentration of K + in solution in K + self-exchange reactions, which does not discount the possibility that equilibrium is established and exchange takes places. However, substantial Cs + ion exchange was noted with high uptake values of up to 3.7(1.2) meq/g. The Cs + ionophore is 2.53 orders of magnitude more selective for Cs + than K + . Therefore, the Cs + uptake value probably overestimates Cs + exchange because some K + is simultaneously being detected. In order to address this issue we also measured the release of K + back into solution and found a large value of 1.7 meq/g. This determination demonstrates that ion exchange is actually taking place. The resultant solids were examined using EDX, and these measurements demonstrate complete exchange of K + for Cs + with no measurable amount of K + being observed in the ion- exchanged solid. Given that the reaction takes place at room temperature in basic media, and that uranyl iodates are extremely insoluble in water even at low pH, we suggest that dissolution and recrystallization do not occur concomitantly with the exchange process. In order to confirm this, powder X-ray diffraction data were collected before and after ion exchange. The diffraction patterns were superimposable, indicating that there is virtually no structural change that occurs concomitantly with ion exchange. CONCLUSIONS The uranyl iodate system continues to provide remarkable new structure types unparalleled by other types of oxoanions. The formation of K[UO 2 (IO 3 ) 3 ] under less acidic conditions (final pH = 2.9 vs. ca. 1) than have been typically employed in the past to prepare uranyl iodates may point to a new family of compounds that form at higher pH. Studies to address this issue are underway. We note that attempts to prepare 46 Cs[UO 2 (IO 3 ) 3 ] directly under conditions equivalent to those used to prepare KUI failed. Instead a new polycrystalline phase forms with a Cs/U/I ratio of 1:3:1. We have been unable as yet to determine the structure of this new cesium uranyl iodate. This result suggests that the K + cations play a structure-directing role in the formation of KUI. 47 REFERENCES 1. Artmann, P. Z. Anorg. Chem. 1913, 79, 327. 2. Venugopalan, M. Z. Anal. Chem. 1956, 153, 187. 3. Ciavatta, L.; De Tommaso, G.; Iuliano, M. Ann. Chim. 2003, 93, 269. 4. Karbowiak, M.; Fourest, B.; Hubert, S.; Moulin, C. Radiochim. Acta 2003, 91, 505. 5. Hall, D. J. Inorg. Nucl. Chem. 1958, 6, 3. 6. Weigel, F.; Engelhardt, L. W. H. J. Less-Common Met. 1983, 91, 339. 7. Bean, A. C.; Peper, S. M.; Albrecht-Schmitt, T. E. Chem. Mater. 2001, 13, 1266. 8. Bean, A. C.; Xu, Y. ; Danis, J. A.; Albrecht-Schmitt, T. E.; Runde, W. Inorg. Chem. 2002, 41, 6775. 9. Sykora, R. E.; McDaniel, S. M.; Wells, D. M.; Albrecht-Schmitt, T. E. Inorg. Chem. 2002, 41, 5126. 10. Sykora, R. E.; Wells, D. M.; Albrecht-Schmitt, T. E. Inorg. Chem. 2002, 41, 2304. 11. Bean, A. C.; Albrecht-Schmitt, T. E. J. Solid State Chem. 2001, 161, 416. 12. Bean, A. C.; Ruf, M.; Albrecht-Schmitt, T. E. Inorg. Chem. 2001, 40, 3959. 13. Sykora, R. E.; Bean, A. C.; Scott, B. L. ; Runde, W. Albrecht-Schmitt, T. E. J. Solid State Chem. 2004, 177, 725. 14. Bean, A. C.; Campana, C. F.; Kwon, O.; Albrecht-Schmitt, T. E. J. Am. Chem. Soc. 2001, 123, 8806. 15. Burns, P. C.; Ewing, R. C.; Hawthorne, F. C. Can. Mineral. 1997, 35, 1551. 16. Burns, P. C. In Uranium: Mineralogy, Geochemistry and the Environment; Burns, P. C.; Finch, R., Eds. Ch. 1, Mineralogical Society of America: Washington, DC, 1999. 17. Burns, P. C.; Miller, M. L.; Ewing, R. C. Can. Mineral. 1996, 34, 845. 48 18. Albrecht-Schmitt, T. E.; Almond, P. M.; Sykora, R. E. Inorg. Chem. 2003, 42, 3788. 19. Bean, A. C.; Scott, B. L.; Albrecht-Schmitt, T. E. ; Runde, W. Inorg. Chem. 2003, 42, 5632. 20. Runde, W.; Bean, A. C.; Albrecht-Schmitt, T. E.; Scott, B. L. Chem. Commun. (Cambridge) 2003, 4, 478. 21. Almond, P. M.; Albrecht-Schmitt, T. E. Inorg. Chem. 2002, 41, 1177. 22. Almond, P. M.; McKee, M. L.; Albrecht-Schmitt, T. E. Angew. Chem. 2002, 114, 3576. 23. Sykora, R. E.; King, J. E.; Illies, A. J.; Albrecht-Schmitt, T. E. J. Solid State Chem. 2004, 177, 1717 24. Hughes, K.-A.; Burns, P. C.; Kolitsch, U. Can. Mineral. 2003, 41, 677. 25. G. M. Sheldrick, SHELXTL PC, Version 6.12, An Integrated System for Solving, Refining, and Displaying Crystal Structures from Diffraction Data; Siemens Analytical X-Ray Instruments, Inc.: Madison, WI 2001. 26. SADABS. Program for absorption correction using SMART CCD based on the method of Blessing: Blessing, R. H. Acta Crystallogr. 1995, A51, 33. 27. Buehlmann, P.; Pretsch, E.; Bakker, E. Chem. Rev. 1998, 98, 1593. 28. Sykora R. E.; Albrecht-Schmitt, T. E. Inorg. Chem. 2003, 42, 2179. 29. Abrahams, S. C.; Bernstein, J. L.; Elemans, J. B. A. A.; Verschoor, G. C. J. Chem. Phys. 1973, 59, 2007. 30. Burns, P. C.; Hawthorne, F. C. Can. Mineral. 1993, 31, 313. 31. Cooper, M. A.; Hawthorne, F. C.; Roberts, A. C.; Grice, J. D.; Stirling, J. A. R. E.; Moffatt, A. Am. Mineral. 1998, 83, 390. 49 32. Brown, I. D.; Altermatt, D. Acta Crystallogr. 1985, B41, 244. 33. Brese, N. E.; O?Keeffe, M. Acta Crystallogr. 1991, B47, 192. 34. Zachariasen, W. H. Acta Crystallogr. 1948, 1, 281. 35. Zachariasen, W. H. Acta Crystallogr. 1954, 7, 788. 36. Zachariasen, W. H. Acta Crystallogr. 1954, 7, 795. 37. Thu?ry, P.; Masci, B. Polyhedron 2004, 23, 649. 38. Thu?ry, P.; Masci, B. J. Chem. Soc., Dalton Trans. 2003, 12, 2411. 39. Sullens, T. A.; Jensen, R. A.; Shvareva, T. Y.; Albrecht-Schmitt, T. E. J. Am. Chem. Soc. 2004, 126, 2676. 40. Dieckmann, G. H.; Ellis, A. B. Solid State Ionics 1989, 32/33, 50. 41. Vochten, R. Am. Mineral. 1990, 75, 221. 42. Benavente, J.; Ramos Barrado, J. R.; Cabeza, A.; Bruque, S.; Martinez, M. Colloids Surf. 1995, A 97, 13. 43. Grohol, D.; Blinn, E. L. Inorg. Chem. 1997, 36, 3422. 44. Johnson, C. H.; Shilton, M. G.; Howe, A. T. J. Solid State Chem. 1981, 37, 37. 45. Moreno-Real, L.; Pozas-Tormo, R.; Martinez-Lara, M.; Bruque-Gamez, S. Mater. Res. Bull. 1987, 22, 29. 46. Pozas-Tormo, R.; Moreno-Real, L.; Martinez-Lara, M.; Rodriguez-Castellon, E. Can. J. Chem. 1986, 64, 35. 47. Obbade, S.; Dion, C.; Saadi, M.; Abraham, F. J. Solid State Chem. 2004, 177, 1567. 50 CHAPTER 3 SYNTHESES, STRUCTURES, AND ION-EXCHANGE PROPERTIES OF THE THREE-DIMENSIONAL FRAMEWORK URANYL GALLIUM PHOSPHATES, Cs4[(UO2)2(GaOH)2(PO4)4]?H2O AND Cs[UO2Ga (PO4)2] ABSTRACT The reaction of UO2(NO3)2?6H2O with Cs2CO3 or CsCl, H3PO4, and Ga2O3 under mild hydrothermal conditions results in the formation of Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) or Cs[UO2Ga(PO4)2] (UGaP2). The structure of UGaP1 was solved from a twinned crystal revealing a three-dimensional framework structure consisting of one- dimensional 1? [Ga(OH)(PO4)2]4? chains composed of corner-sharing GaO6 octahedra and bridging PO4 tetrahedra that extend along the c axis. The phosphate anions bind the UO22+ cations to form UO7 pentagonal bipyramids. The UO7 moieties edge-share to create dimers that link the gallium phosphate substructure into a three-dimensional 3 ? [(UO2)2(GaOH)2(PO4)4] 4? anionic lattice that has intersecting channels running down the b and c axes. Cs+ cations and water molecules occupy these channels. The structure of UGaP2 is also three-dimensional and contains one-dimensional 1? [Ga(PO4)2]3? gallium phosphate chains that extend down the a axis. These chains are formed from fused eight- membered rings of corner-sharing GaO4 and PO4 tetrahedra. The chains are in turn linked together into a three-dimensional 3? [UO2Ga(PO4)2]1? framework by edge-sharing 51 UO7 dimers as occurs in UGaP1. There are channels that run down the a and b axes through the framework. These channels contain the Cs+ cations. Ion-exchange studies indicate that the Cs+ cations in UGaP1 and UGaP2 can be exchanged for Ca2+ and Ba2+. Crystallographic data: UGaP1, monoclinic, space group P21/c, a = 18.872(1), b = 9.5105(7), c = 14.007(1) ?,  = 109.65(3), Z = 4 (T = 295 K); UGaP2, triclinic, space group P1, a = 7.7765(6), b = 8.5043(7), c = 8.9115(7) ?,  = 66.642(1),  = 70.563(1),  = 84.003(2), Z = 2 (T = 193 K). INTRODUCTION The syntheses, structures, and physicochemical properties of gallium phosphates have been the subjects of intense interest for two decades owing to their enormously rich structural chemistry that is particularly well noted for forming microporous open- framework topologies that in some cases are similar to the AlPO4 family of compounds.1- 4 The diversity in this class of compounds finds its origins in the fundamental coordination chemistry of Ga(III), which occurs as GaO4 tetrahedra (e.g. in [H3N(CH2)3NH3][GaH(PO4)2]5 and [Ga(HPO4)(PO4)(OH)]?[(C2H7N)3N]?H2O6), GaO5 trigonal bipyramids (e.g. in [NH4][Ga(OH)(PO4)]7 and Ga3(PO4)3?C3H9N?H2O8), and GaO6 octahedra (e.g. in [NH4]3[Ga2(PO4)3]7 and Rb(GaPO4)2(OH)(H2O)?H2O9). In fact, there are examples of compounds where Ga exhibits multiple geometries in the same compound as has been found for Na[Ga2(OH)(PO4)2]10 and Na3Ga5(PO4)4O2(OH)2?2H2O,11 which contain GaO5 trigonal bipyramids and GaO6 octahedra. These fundamental building units can combine with phosphate tetrahedra in a vast number of permutations to create zero-dimensional structures in 52 [Ga(HPO4)(PO4)(OH)]?[(C2H7N)3N]?H2O6 and [C6H5NH]2[H8Ga4P4O21],12 one-dimensional chains as found in the organically templated gallium phosphates [NH3(CH2)4NH3][Ga(PO4)(PO3OH)],13 [(1R,2R)-C6H10(NH3)2][Ga(OH)(HPO4)2]?H2O,14 and [NH3(CH2)2(NH3)][Ga(H2PO4)(P2O7)],15 two-dimensional sheets in [C5H6N][Ga(HPO4)2(H2O)2]16 and [NH4]Ga(OH)(PO4)],7 and three-dimensional networks in NaGa2(OH)(PO4)2,10 Na3Ga5(PO4)4O2(OH)2?2H2O,11 and [NH3(CH2)3(NH3)][Ga3(PO4)3(H2O)].17 The structural complexity and properties of gallium phosphates can be expanded through the inclusion of transition metals into the anionic lattice. The incorporation of transition metals into gallium phosphates can take place in several ways. First, partial substitution at Ga sites can take place to yield disordered compounds, which is illustrated by [C5H5NH][CoGa2P3O12], where CoO4/GaO4 disorder exists.18 Anomalous X-ray scattering is a powerful tool in this regard, and has been used to establish the distribution of Zn2+ and Ga3+ in a series of zinc-substituted gallium phosphates.19 Second, the transition metal centers can force the adoptions of new structure types to accommodate additional building units as occurs in Rb[(VO)(H2O)Ga(PO4)2]20 and Mn3(H2O)6Ga4(PO4)6,21 the latter of which forms gallium phosphate channels to house Mn3(H2O)6O8 clusters. This compound also exhibits superexchange interactions between Mn(II) centers, illustrating one of the many properties that can be added to gallium phosphates by incorporating transition metals.21 In order to prepare uranyl compounds with three-dimensional open-framework structures, we have focused on the use of octahedral anions, such as periodate, IO65?, as building units.22,23 The formation of three-dimensional network structures incorporating the approximately linear uranyl, UO22+, cation are not common because uranyl- 53 containing polyhedra typically condense with uranyl units parallel to one another, and because the uranyl oxo atoms are typically terminal, layered structures most often result.24 In spite of this structural propensity, a number of different uranyl systems have been shown to adopt three-dimensional framework structures including phosphates such as [C4H12N2](UO2)[(UO2)(PO4)]4?2H2O25 and [(UO2)3(PO4)O(OH)(H2O)2](H2O),26 the fluorophosphate [C6H14N2]2[(UO2)6(H2O)2F2(PO4)2(HPO4)4]?4H2O,27 arsenates, e.g. [NH(C2H5)3][(UO2)2(AsO4)(AsO3OH)],25 vanadates, e.g. A(UO2)4(VO4)3 (A = Li, Na),28 and oxides, e.g. (NH4)3(H2O)2{[(UO2)10O10(OH)][(UO4)(H2O)2]}.29 In this chapter we report the preparation and characterization of the first examples of actinide-containing gallium phosphates, Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) and Cs[UO2Ga(PO4)2] (UGaP2), both of which possess three-dimensional framework structures. EXPERIMENTAL Syntheses. UO2(NO3)2?6H2O (98%, Alfa-Aesar), Ga2O3 (99.99%, Alfa-Aesar), Cs2CO3 (99.99%, Alfa-Aesar), H3PO4 (98%, Aldrich), KCl (99%, Aldrich), NaCl (99%, Aldrich), CsCl (99.9%, Cerac), CaCl2 (99%, Aldrich), and BaCl2 (99%, Aldrich) were used as received. Reactions were performed in the PTFE-lined Parr 4749 autoclaves. Distilled and Millipore filtered water with a resistance of 18.2 M?cm was used in all reactions. Standard precautions were performed for handling radioactive materials during work with UO2(NO3)2?6H2O and the products of the reactions. Semi-quantitative SEM/EDX analyses were performed using a JEOL 840/Link Isis instrument. Cs, U, Ga, and P percentages were calibrated against standards. 54 Cs4[(UO2)2(GaOH)2(PO4)4](H2O) (UGaP1). UO2(NO3)2?6H2O (0.238 g, 0.474 mmol), Ga2O3 (0.112 g, 0.948 mmol), H3PO4 (0.186 g, 1.88 mmol), Cs2CO3 (0.464 g, 1.41 mmol), and 1 mL water were loaded into a 23 mL autoclave. The autoclave was then heated to 180 ?C in a box furnace for 12 d and slow cooled at 3 ?C/h to room temperature. The product mixture consisted of a colorless solution and a bright yellow precipitate. The solid was washed with water and methanol and allowed to dry. Rectangular shape tablets of UGaP1 were then separated, with a yield of 130 mg (55% yield based on the U), from the mixture for further study. The Cs:U:Ga:P ratio determined from EDX analysis was approximately 2:1:1:2. Cs[UO2Ga(PO4)2] (UGaP2). UO2(NO3)2?6H2O (0.307 g, 0.611 mmol), Ga2O3 (0.144 g, 1.22 mmol), H3PO4 (0.239 g, 2.44 mmol), and CsCl (0.309 g, 1.83 mmol) were loaded into the 23 mL autoclave followed by the addition of 2 mL of water. The autoclave was heated for 4 d at 180 ?C in a box furnace and then cooled to room temperature at 9 ?C/h. The solid product consisted of a mixture of white and green- yellow crystals. After decanting the mother liquor, the product was washed with water and methanol and allowed to dry. Rectangular shape green-yellow tablets of UGaP-2 were then separated from the mixture of yellow and white amorphous precipitates. The isolated yield was 46 mg (15% yield based on the U). EDX analysis provided Cs:U:Ga:P ratio of 1:1:1:2. Crystallographic Studies. Crystals of Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) and Cs[UO2Ga(PO4)2] (UGaP2) were mounted on glass fibers and aligned on a Bruker SMART APEX CCD X-ray diffractometer. Intensity measurements were performed using graphite monochromated Mo Ka radiation from a sealed tube and 55 monocapillary collimator. SMART (v 5.624) was used for preliminary determination of the cell constants and data collection control. The intensities of reflections of a sphere were collected by a combination of 3 sets of exposures (frames). Each set had a different f angle for the crystal and each exposure covered a range of 0.3? in w. A total of 1800 frames were collected with an exposure time per frame of 30 s for UGaP1 and 120 s for UGaP2. Crystals of UGaP1 suffered from two problems. First, the crystal cracked when cooled to 193 K, and the data was therefore collected at 295 K. Second, all of the crystals examined proved to be twinned. GEMINI was used to separated the individual components of the twin, and the structure was solved as follows. For UGaP1 and UGaP2 determination of integrated intensities and global refinement were performed with the Bruker SAINT (v 6.02) software package using a narrow-frame integration algorithm. A face-indexed analytical absorption correction was initially applied using XPREP, where individual shells of unmerged data were corrected analytically.30 These files were subsequently treated with a semiempirical absorption correction by SADABS.31 The program suite SHELXTL (v 6.12) was used for space group determination (XPREP), direct methods structure solution (XS), and least-squares refinement (XL).30 The final refinements included anisotropic displacement parameters for all atoms. Secondary extinction was not noted for either crystal. Some crystallographic details are given in Table 3.1. Atomic coordinates, symmetry, displacement parameters and occupancy for UGaP1 and UGaP2 can be found in Tables 3.2 and 3.3 respectively. 56 Table 3.1. Crystallographic Data for Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) and Cs[UO2Ga(PO4)2] (UGaP2). Formula Cs4[(UO2)2(GaOH)2(PO4)4]?H2O Cs[UO2Ga(PO4)2] Formula Mass 1641.04 662.60 Color and habit yellow tablet yellow-green tablet Space group P21/c (No. 14) P1 (No. 2) a (?) 18.872(1) 7.7765(6) b (?) 9.5105(7) 8.5043(7) c (?) 14.007(1) 8.9115(7)  (?) 90 66.642(1)  (?) 109.65(3) 70.563(1)  (?) 90 84.003(2) V (?3) 2367.5(3) 509.96(7) Z 4 2 T (K) 295 193 l (?) 0.71073 0.71073 Maximum 2q (deg.) 56.64 56.60 rcalcd (g cm?3) 4.598 4.315 m(Mo Ka) (cm?1) 223.20 223.65 R(F) for Fo2 > 2s(Fo2) 0.0424 0.0283 Rw(Fo2) b 0.0973 0.0626 a ( )R F F F F= -  o c o . b ( ) ( )R F w F F wF w o 2 o 2 c 2 2 o 4 1 2 = -    . 57 Table 3.2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) Atom Symmetrya x y z Ueq (?2) b Occup. U(1) 4 e 1 0.53831(2) -0.70597(4) 0.51964(3) 0.01196(9) 1 U(2) 4 e 1 0.96212(2) -0.71325(4) 0.98701(3) 0.01205(9) 1 Ga(1) 4 e 1 1.25499(6) -0.77386(9) 1.23126(8) 0.0119(2) 1 Ga(2) 4 e 1 1.25230(6) -0.74603(9) 0.98122(8) 0.0116(2) 1 P(1) 4 e 1 0.65190(15) -0.4630(3) 0.6033(2) 0.0125(5) 1 P(2) 4 e 1 1.37577(14) -0.9350(3) 1.41044(19) 0.0114(5) 1 P(3) 4 e 1 1.13089(14) -0.9361(3) 1.05030(19) 0.0119(5) 1 P(4) 4 e 1 0.84790(14) -0.4578(3) 0.92909(19) 0.0121(5) 1 O(1) 4 e 1 0.6951(4) -0.4196(8) 0.5334(6) 0.0191(16) 1 O(2) 4 e 1 0.6759(4) -0.3792(8) 0.7011(5) 0.0214(17) 1 O(3) 4 e 1 0.5646(4) -0.4485(8) 0.5480(6) 0.0201(16) 1 O(4) 4 e 1 0.6597(4) -0.6228(8) 0.6239(6) 0.0194(16) 1 O(5) 4 e 1 1.4520(4) -0.8644(8) 1.4320(6) 0.0221(17) 1 O(6) 4 e 1 1.3367(4) -0.8782(8) 1.4808(5) 0.0189(16) 1 O(7) 4 e 1 1.3898(4) -1.0935(7) 1.4311(6) 0.0191(16) 1 O(8) 4 e 1 1.3282(4) -0.9177(7) 1.2987(5) 0.0171(16) 1 O(9) 4 e 1 1.1138(4) -1.0931(8) 1.0239(5) 0.0167(15) 1 O(10) 4 e 1 1.1620(4) -0.8683(7) 0.9734(5) 0.0157(15) 1 O(11) 4 e 1 1.0566(4) -0.8644(8) 1.0469(6) 0.0207(16) 1 O(12) 4 e 1 1.1873(4) -0.9266(7) 1.1595(5) 0.0132(14) 1 O(13) 4 e 1 0.8396(4) -0.6171(7) 0.9056(5) 0.0160(15) 1 O(14) 4 e 1 0.9351(4) -0.4443(8) 0.9846(6) 0.0193(16) 1 O(15) 4 e 1 0.8051(4) -0.4207(7) 1.0013(6) 0.0186(16) 1 58 O(16) 4 e 1 0.8234(4) -0.3690(7) 0.8332(5) 0.0158(15) 1 O(17) 4 e 1 0.5129(4) -0.7187(7) 0.6326(5) 0.0203(16) 1 O(18) 4 e 1 0.5646(4) -0.6883(8) 0.4085(5) 0.0204(16) 1 O(19) 4 e 1 0.9609(4) -0.6969(7) 1.1141(6) 0.0205(16) 1 O(20) 4 e 1 0.9633(4) -0.7357(8) 0.8610(5) 0.0183(16) 1 O(21) 4 e 1 1.2074(4) -0.7272(7) 0.8361(5) 0.0143(15) 1 O(22) 4 e 1 1.2998(4) -0.7656(7) 1.1254(5) 0.0139(14) 1 O(23) 4 e 1 0.7509(7) -0.6656(13) 1.0535(10) 0.068(3) 1 Cs(1) 4 e 1 0.70707(4) -0.81967(8) 0.83263(7) 0.0337(2) 1 Cs(2) 4 e 1 0.55975(5) -0.47401(9) 0.80426(6) 0.03157(19) 1 Cs(3) 4 e 1 0.83707(4) -0.60810(8) 0.67258(6) 0.02768(17) 1 Cs(4) 4 e 1 0.93184(4) -0.45410(8) 1.23164(6) 0.02884(18) 1 a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 59 Table 3.3. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs[UO2Ga(PO4)2] (UGaP2). Atom Symmetrya x y z Ueq (?2) b Occup. U(1) 2 i 1 0.21442(3) 0.92139(3) 0.33450(3) 0.00955(7) 1 Ga(1) 2 i 1 0.24602(10) 0.57980(9) 0.98833(9) 0.01023(14) 1 P(1) 2 i 1 -0.0413(2) 0.3270(2) 1.2841(2) 0.0103(3) 1 P(2) 2 i 1 0.5302(2) 0.7365(2) 1.0581(2) 0.0099(3) 1 O(1) 2 i 1 0.0560(6) 0.8460(6) 0.2720(6) 0.0159(10) 1 O(2) 2 i 1 0.3671(7) 0.9913(6) 0.4080(6) 0.0145(9) 1 O(3) 2 i 1 -0.2013(6) 0.3571(6) 0.0150(10) 0.0195(18) 1 O(4) 2 i 1 0.0248(7) 0.1492(6) 1.3771(6) 0.0140(9) 1 O(5) 2 i 1 -0.0888(6) 0.3241(6) 1.1313(6) 0.0137(9) 1 O(6) 2 i 1 0.1153(7) 0.4610(6) 1.2133(6) 0.0158(10) 1 O(7) 2 i 1 0.3828(7) 0.7476(6) 0.9725(6) 0.0153(10) 1 O(8) 2 i 1 0.5978(6) 0.5503(6) 1.1144(6) 0.0131(9) 1 O(9) 2 i 1 0.6848(7) 0.8614(6) 0.9304(6) 0.0156(10) 1 O(10) 2 i 1 0.4475(6) 0.7700(6) 1.2225(6) 0.0141(9) 1 Cs(1) 2 i 1 0.66307(6) 0.68745(6) 0.49084(6) 0.02003(11) 1 a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 60 Ion-exchange Studies. An Accumet Basic AB 15 pH meter was used for potentiometric measurements. The ion-exchange properties of UGaP1 and UGaP2 were evaluated to determine the uptake values of Na+, K+, Ca2+, and Ba2+ ions from NaCl, KCl, CaCl2, and BaCl2 solutions respectively in the range of the concentrations from 0.0001 M to 0.1 M. The release values of Cs+ ions were also measured. The concentrations of the solutions before and after uptake and release were determined using ion-selective electrodes.32 The volume to mass ratio was 200:1 (2 mL of the solution for 0.01 g of sample). Uptake of Sr2+ and duplicated measurements of Ca2+ concentrations were performed on a Varian 240 AA spectrometer with Hollow cathode Ca and Sr lamps. The volume to mass ratio in these experiments was 200:1 (4 mL of the solution for 0.02 g of sample). Thermal analysis. DSC data were collected using differential scanning calorimeter TA Instrument, model. The samples of known mass encapsulated in aluminum container, were heated under nitrogen atmosphere from 30 ?C to 600 ?C with heating rate 10?C /minute. RESULTS AND DISCUSSION Synthesis. While the syntheses of Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) and Cs[UO2Ga(PO4)2] (UGaP2) appear superficially similar, the reaction stoichiometries, durations, and perhaps most importantly pH are dramatically different. For the preparation of both compounds, we noted a reduction in pH between the starting reaction mixtures and that of the final mother liquor. The pH decreased from 9.1 to 7.1 during the 61 crystallization of UGaP1 and from 2.4 to 1.9 during the formation of UGaP2. The substitution of Cs2CO3 by CsCl in the preparation of UGaP2 plays a significant role in controlling the pH of these reactions. The synthesis of UGaP1 and UGaP2 also proved to be highly sensitive to temperature, and changes by as little as 10 ?C had dramatic effects on yield, product crystallinity, and crystal size. Structure. Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1). The structure of UGaP1 was solved to reveal a three-dimensional framework of interconnected one-dimensional gallium phosphate and uranyl phosphate chains that run approximately perpendicular to one another. The gallium phosphate chain (Figure 3.1) is composed of corner-sharing GaO6 octahedra and bridging PO4 tetrahedra that extend along the c axis, and can be formulated as 1? [Ga(OH)(PO4)2]4?. We propose based on charge balance considerations and bond-valence sums that the Ga?O?Ga backbone that runs down the gallium phosphate chains is protonated. The bond-valence sums for the O(21) and O(22) atoms in this chain are 1.12 and 1.16, which are values expected for hydroxyl sites.33,34 Not surprisingly, given the presence of cesium and uranium in this structure, the hydrogen atoms could not be reliably located. This type of chain is similar to that found in the aluminophosphate mineral tancoite, LiNa2H[Al(PO4)2(OH)],35,36 and has been recognized in a large number of aluminum, gallium, and transition metal phosphates.14,37-43 The Ga? O bond distances range from 1.940(7) to 2.036(7) ?. The P?O bond distances show little variation and occur from 1.515(7) to 1.574(7) ?. Selected bond distances and angles for UGaP1 are given in Table 3.4. 62 Figure 3.1. A view of the one-dimensional 1? [Ga(OH)(PO4)2]4? chains composed of corner-sharing GaO6 octahedra and bridging PO4 tetrahedra that extend along the c axis in Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1). 50% probability ellipsoids are depicted. 63 Table 3.4 Selected Bond Distances (?) and Angles (o) for Cs4[(UO2)2(GaOH)2(PO4)4] (UGaP1). Distances (?) U(1)?O(3) 2.364(7) Ga(2)?O(6) 1.984(7) U(1)?O(3)' 2.503(7) Ga(2)?O(10) 2.036(7) U(1)?O(4) 2.400(7) Ga(2)?O(15) 1.982(7) U(1)?O(5) 2.255(8) Ga(2)?O(21) 1.930(7) U(1)?O(7) 2.307(7) Ga(2)?O(22) 1.924(7) U(1)?O(17) 1.805(7) P(1)?O(1) 1.526(7) U(1)?O(18) 1.794(7) P(1)?O(2) 1.516(8) U(2)?O(9) 2.306(7) P(1)?O(3) 1.574(7) U(2)?O(11) 2.223(7) P(1)?O(4) 1.544(8) U(2)?O(13) 2.390(7) P(2)?O(5) 1.523(8) U(2)?O(14) 2.376(7) P(2)?O(6) 1.515(7) U(2)?O(14) 2.606(7) P(2)?O(7) 1.542(7) U(2)?O(19) 1.795(7) P(2)?O(8) 1.529(7) U(2)?O(20) 1.786(7) P(3)?O(9) 1.546(8) Ga(1)?O(2) 1.971(7) P(3)?O(10) 1.531(7) Ga(1)?O(8) 1.948(7) P(3)?O(11) 1.546(8) Ga(1)?O(12) 1.971(7) P(3)?O(12) 1.545(7) Ga(1)?O(16) 1.988(7) P(4)?O(13) 1.547(7) Ga(1)?O(21) 1.962(7) P(4)?O(14) 1.571(8) Ga(1)?O(22) 1.940(7) P(4)?O(15) 1.532(7) Ga(2)?O(1) 1.973(7) P(4)?O(16) 1.521(7) 64 Angles (?) O(18)?U(1)?O(17) 178.3(3) O(20)?U(2)?O(19) 178.1(3) 65 There are two crystallographically unique uranium centers in UGaP1. Both of these U atoms are found as part of uranyl, UO22+, cations with short U=O bond distances of 1.794(7) and 1.805(7) ? to U(1), and 1.786(7) and 1.795(7) ? to U(2). In both cases the uranyl cations are nearly linear with O(18)?U(1)?O(17) and O(20)?U(2)?O(19) bond angles of 178.3(3)? and 178.1(3)?, respectively. The phosphate anions bind the UO22+ cations to form UO7 pentagonal bipyramids with U(1)?O and U(2)?O distances being found from 2.255(8) to 2.503(7) ? and 2.223(7) to 2.606(7) ?, respectively. These bond distances can be used to calculate bond-valence sums for the U(1) and U(2) centers of 5.94 and 5.96, which are consistent with hexavalent uranium.33,34,44 The bond angles in the pentagonal plane show substantial variation because the phosphate anions are found in both chelating and bridging modes. Therefore, there is an acute angle of 58.9(3)? in the equatorial plane ascribed to the chelation of the uranyl cation containing U(1) by a phosphate anion. The remaining angles in this plane range from 63.1(3)? to 81.7(3)?. The same situation occurs for U(2) with an acute angle of 57.6(2)? and the other angles ranging from 61.9(3)? to 84.8(3)?. The UO7 moieties edge-share to create dimers that are bridged by phosphate anions to create one-dimensional chains, shown in the Figure 3.2, that can be formulated as 1? [UO2(PO4)2]4?. There are two different uranyl phosphate chains, one chain exclusively contains U(1) and the other U(2). The one-dimensional topology of these chains is similar to that found in Sr[UO2(SeO3)2]?2H2O where one of the P?O bonds on each phosphate anion has been replaced by a stereochemically active lone-pair on Se(IV).45 These chains are distinct from those found in the one-dimensional uranyl phosphate, parsonite, Pb2[UO2(PO4)2].46 66 Figure 3.2. A depiction of the one-dimensional 1? [UO2(PO4)2]4? chains in Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1). 50% probability ellipsoids are depicted. 67 The interconnection of the 1? [Ga(OH)(PO4)2]4? and 1? [UO2(PO4)2]4? substructures creates a three-dimensional 3? [(UO2)2(GaOH)2(PO4)4]4? anionic lattice that has intersecting channels running down the b and c axes. Small tunnels also extends down [111]. A view showing the largest channels, which measure approximately 3.5 x 7.3 ?, extending along b is shown in Figure 3.3. There are four crystallographically unique Cs+ cations and a water molecule in UGaP1 that occupy these channels. The Cs+ cations form between eight and ten long contacts with oxygen atoms in the structure with average distances of 3.254(8), 3.374(8), 3.171(8), and 3.243(8) ? to Cs(1), Cs(2), Cs(3), and Cs(4), respectively. There is a short interaction of 2.869(13) ? between the water molecule and Cs(3). In addition, the water molecule forms hydrogen-bonding interactions of 2.740(8) and 2.821(8) ? with phosphate oxygen atoms in the gallium phosphate chain. Cs[UO2Ga(PO4)2] (UGaP2). The structure of UGaP2 is three-dimensional, being constructed from one-dimensional 1? [Ga(PO4)2]3? gallium phosphate chains and one-dimensional uranyl phosphate chains, the former of which extend down the a axis. These gallium phosphate chains are formed from fused eight-membered rings of alternating corner-shared GaO4 and PO4 tetrahedra. (Figure 3.4) These types of rings are important building units in gallium phosphate chemistry and have been observed before in compounds such as [Ga(HPO4)(PO4)(OH)]?[(C2H7N)3N]?H2O6 and [NH3(CH2)4NH3][Ga(PO4)(HPO4)].47 The Ga atoms are the sites of ring fusion for these chains. The Ga and P sites are easily distinguished from one another in these rings with average Ga?O and P?O bond distances of 1.814(5) and 1.532(5) ?, respectively. The 68 Figure 3.3. An illustration of the three-dimensional 3? [(UO2)2(GaOH)2(PO4)4]4? anionic lattice of Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) that has intersecting channels running down the b and c axes. These channels house Cs+ cations and water molecules. Cs is shown in magenta, UO7 pentagonal bipyramids are shown in green, GaO6 octahedra in blue, and PO4 tetrahedra in yellow. 69 Figure 3.4. A view of the one-dimensional 1? [Ga(PO4)2]3? gallium phosphate chains in Cs[UO2Ga(PO4)2] (UGaP2) formed from fused eight-membered rings of alternating corner-shared GaO4 and PO4 tetrahedra. 50% probability ellipsoids are depicted. 70 two crystallographically unique PO43? anions each use two oxygen atoms to bridge between Ga centers. The oxygen atoms that are not used in this fashion instead bind the uranium centers in the structure. Selected bond distances and angles for UGaP2 can be found in Table 3. 5. Unlike UGaP1 there is only one crystallographically unique U center in UGaP2 that is found in the form of a UO7 pentagonal bipyramid. There are two short U=O bonds of 1.774(5) and 1.779(5) ? that make up the approximately linear uranyl cation with an O?U?O angle of 177.2(2)?. The equatorial U?O distances range from 2.283(5) to 2.501(5) ? and are within normal limits. Bond-valence sum calculations on U yield a value of 5.97, which is consistent with U(VI).33,34,44 The UO7 units edge-share to create dimers with a central inversion center. These dimers are bridged by phosphate anions to create one-dimensional chains whose formula can be described as 1? [UO2(PO4)2]4?, and are illustrated in Figure 3.5 As can be seen from this figure, these chains are essentially the same as those found in UGaP1. The gallium phosphate and uranyl phosphate chains run approximately perpendicular to one another to create a three-dimensional 3? [UO2Ga(PO4)2]1? framework with intersecting channels running down the a and b axes. (Figure 3.6) These channels at their largest are 3.1 x 6.7 ?, and are of the size expected to house Cs+ cations in the absence of occluded water molecules. The Cs+ cations form long ionic contacts with the surrounding anionic lattice with eight contacts ranging from 3.115(5) ? to 3.440(5) ?. Thermal analysis. DSC data for UGaP1 demonstrate that the structure of the compound remains relatively stable at temperatures below 400 ?C. (Figure 3.7) The loss 71 Table 3.5. Selected Bond Distances (?) and Angles (o) for Cs[UO2Ga(PO4)2] (UGaP2). Distances (?) U(1)?O(1) 1.774(5) P(1)?O(3) 1.512(5) U(1)?O(2) 1.779(5) P(1)?O(4) 1.538(5) U(1)?O(3) 2.482(5) P(1)?O(5) 1.534(5) U(1)?O(4) 2.385(5) P(1)?O(6) 1.546(5) U(1)?O(4)' 2.501(5) P(2)?O(7) 1.552(5) U(1)?O(9) 2.283(5) P(2)?O(8) 1.552(5) U(1)?O(10) 2.295(5) P(2)?O(9) 1.505(5) Ga(1)?O(5) 1.807(5) P(2)?O(10) 1.519(5) Ga(1)?O(6) 1.824(5) Ga(1)?O(7) 1.801(5) Ga(1)?O(8) 1.822(5) Angles (?) O(1)?U(1)?O(2) 177.2(2) 72 Figure 3.5 An illustration of the one-dimensional 1? [UO2(PO4)2]4? chains in Cs[UO2Ga(PO4)2] (UGaP2). 50% probability ellipsoids are depicted. 73 Figure 3.6. A depiction of the three-dimensional 3? [UO2Ga(PO4)2]1? framework in Cs[UO2Ga(PO4)2] (UGaP2) with intersecting channels running down the a and b axes. Cs is shown in magenta, UO7 pentagonal bipyramids are shown in green, GaO4 tetrahedra in blue, and PO4 tetrahedra in yellow. 74 -14 -12 -10 -8 -6 -4 -2 0 0 100 200 300 400 500 600 Temperature, 0C H e a t f l o w , m W . . 150oC 172oC 536oC 546oC Figure 3.7. DSC data of UGaP1 demonstrate relative stability of the structure at temperatures below 400 ?C. Two peaks at 150 ?C and 172 ?C correspond to loss of surface and interstitial water respectively. 75 of surface and interstitial water occurs at 150 ?C and 172 ?C, respectively. Loss of hydroxyl groups and overall sluggish decomposition of the structure take place at temperatures above 400 ?C. The peaks cannot be assigned to defined events due to approach of the instrumentation limits. The data for UGaP2 surprisingly exhibit a significant exothermic peak at 87 ?C, followed by a series of three successive endothermic peaks in the temperature range of 127 ?C ? 155 ?C and single endothermic events at 260 ?C, 365 ?C, 508 ?C and 528 ?C. (Figure 3.8) The data clearly show that UGaP2 undergoes different structural rearrangements and phase transformations upon increases in temperature. Ion-Exchange Studies. The exchange and mobility of interlayer cations in two- dimensional uranyl compounds have been studied in great detail, particularly for layered uranyl phosphates.48 Both layered49 and open-framework uranyl vanadates, such as A(UO2)4(VO4)3 (A = Li, Na),28 have been shown to exhibit high cationic conductivities. We have also shown that the layered uranyl iodate, K[UO2(IO3)3], shows selective ion- exchange of K+ for Cs+.50 Given the three-dimensional framework structures of UGaP1 and UGaP2, we were interested in evaluating the potential exchange of Cs+ with H+, Na+, K+, Ca2+, Sr2+ and Ba2+ ions. The concentrations of K+, Na+, Ca2+ and Ba2+ solutions were determined by potentiometric measurements using ion-selective electrodes.32,50 This technique allows for the determination of both uptake and release of different ions by using different ion-selective electrodes. Sr2+ and duplicated Ca2+ measurements were performed using atomic adsorption spectroscopy. 76 -8 -7 -6 -5 -4 -3 -2 -1 0 0 100 200 300 400 500 600 Temperature, oC H e a t f l o w , m W . . 87oC 127oC 147oC 155oC 260oC 365oC 508oC 528oC 555oC Figure 3.8. DSC data of UGaP2. 77 Direct measurement of cation uptake often over- or underestimates actual ion- exchange values owing to the typically large concentration of the exchanging ion in solution relative to the ion initially present in the material, as well as simple sorption by the material. Therefore more realistic values can be obtained by measuring the concomitant release of Cs+ cations into solution. For UGaP1 and UGaP2 no exchange of Cs+ was noted with H+, Na+, or K+. However, the Cs+ cations in UGaP1 can be exchanged for Ca2+, Sr2+ and Ba2+ cations. Direct Ca2+ uptake measured by ion-selective electrodes of 0.63(10) meq/g was found. Cs+ cation release for this process yielded a similar value of 0.57(10) meq/g. These values are in perfect match with uptake, 0.61(10) meq/g, measured by atomic adsorption spectroscopy. Sr2+ ion-exchange capacity was determined as 0.40(7) meq/g. In the mixed 1:1 SrCl2/NaCl solutions again no Na+ uptake and reduced 0.35(5) meq/g Sr2+ uptake were detected. The smaller Sr2+ uptake value in the mixed solutions could be ascribed to the reduced concentration of Sr2+ and therefore increased volume to mass ratio during the exchange. When Cs+ cations are exchanged for Ba2+ cations a release value of 0.36(10) meq/g was found. The smaller channel size of UGaP2 apparently better accommodates Ca2+ than UGaP1 does, and a large uptake value of 1.79(30) meq/g was found; the Cs+ release was 1.06(20) meq/g. At the same time exchange of Cs+ with Sr2+ was not detected for UGaP2. UGaP1 and UGaP2 show similar ability to exchange Cs+ for Ba2+, and a release of 0.31(20) meq/g was measured for UGaP2. Crystallinity and absence of structural change in the both UGaP1 and UGaP2 sample were monitored by powder diffraction analysis of all samples before and after ion-exchange. 78 CONCLUSIONS In this work we have demonstrated that uranyl gallium phosphates can be prepared under mild hydrothermal conditions. Our original goal was to prepare three- dimensional framework compounds by incorporating octahedral GaO6 building units. The crystal structure of Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) indicates that this strategy was successful. However, the structure of Cs[UO2Ga(PO4)2] (UGaP2), which contains tetrahedral GaO4 units, demonstrates that octahedral gallium is not necessary for the construction of open-framework uranyl gallium phosphates. Thermal analysis revealed the absence of any structural of UGaP1 up to 400 ?C. Ion-exchange experiments show that the Cs+ cations in UGaP1 can be highly selectively exchanged with Sr2+ with the uptake of 0.4 meq/g in the presence of Na+ cations. Summarizing these data we can recommend UGaP1, as a potential 90Sr selective material, which should be studied further under conditions more closely approximating those present in nuclear waste. 79 REFERENCES 1. Cheetham, A. K.; F?rey, G.; Loiseau, T. Angew. Chem., Int. Ed. Engl. 1999, 38, 3268. 2. Wilson, T.; Lok, B. M.; Flanigen, E. M. 1982, U.S. patent no. 4 310 440. 3. Wilson, T.; Lok, B. M.; Messina, C. A.; Cannan, T. R.; Flanigen, E. M. J. Am. Chem. Soc. 1982, 104, 1146. 4. Parise, J. B. J. Chem. Soc., Chem. Commun. 1985, 9, 606. 5. Loiseau, T.; Serpaggi, F.; F?rey, G. J. Chem. Soc., Chem. Commun. (Cambridge) 1997, 12, 1093. 6. Serpaggi, F.; Loiseau, T.; F?rey, G. Acta Crystallogr. 1997, C53, 1568. 7. Bonhomme, F.; Thoma, S. G.; Nenoff, T. M. Micropor. Mesopor. Mater. 2002, 53, 87. 8. Parise, J. B. Acta Crystallogr. 1986, C42, 144. 9. Beitone, L.; Loiseau, T.; F?rey, G. Acta Crystallogr. 2002, C58, i103. 10. Guesdon, A.; Monnin, Y.; Raveau, B. J. Solid State Chem. 2003, 172, 237. 11. Attfield, M. P.; Morris, R. E.; Gutierrez-Puebla, E.; Monge-Bravo, A.; Cheetham, A. K. J. Chem. Soc., Chem. Commun. 1995, 8, 843. 12. Wragg, D. S.; Morris, R. E. J. Am. Chem. Soc. 2000, 122, 11246. 13. Chippindale, A. M.; Bond, A. D.; Law, A. D. Cowley, A. R. J. Solid State Chem. 1998, 136, 227. 14. Lin, H.-M.; Lii, K.-H. Inorg. Chem. 1998, 37, 4220. 15. Chippindale, A. M. Chem. Mater. 2000, 12, 227. 16. Leech, M. A.; Cowley, A. R.; Prout, K.; Chippindale, A. M. Chem. Mater. 1998, 10, 451. 80 17. Brouca-Cabarrecq, C.; Mosset, A. J. Mater. Chem. 2000, 10, 445. 18. Chippindale, A. M.; Walton, R. I. J. Chem. Soc., Chem. Commun. 1994, 21, 2453. 19. Cowley, A. R.; Jones, R. H.; Teat, S. J. Chippindale, A. M. Micropor. Mesopor. Mater. 2002, 51, 51. 20. Hammond, R. P.; Zubieta, J. A. J. Solid State Chem. 1999, 144, 442. 21. Hsu, K.-F.; Wang, S.-L. Inorg. Chem. 2000, 39, 1773. 22. Sykora, R. E.; Albrecht-Schmitt, T. E. Inorg. Chem. 2003, 42, 2179. 23. Sullens, T. A.; Jensen, R. A.; Shvareva, T. Y.; Albrecht-Schmitt, T. E. J. Am. Chem. Soc. 2004, 126, 2676. 24. a) Burns, P. C.; Miller, M. L.; Ewing, R. C. Can. Mineral. 1996, 34, 845. b) Burns, P. C. In Burns, P. C.; Finch, R., Eds. Uranium: Mineralogy, Geochemistry and the Environment; Mineralogical Society of America: Washington, DC, 1999; Chapter 1. c) Burns, P. C. Mater. Res. Soc. Symp. Proc. 2004, 802, 89. 25. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2004, 177, 2675. 26. Burns, P. C.; Alexopoulos, C. M.; Hotchkiss, P. J.; Locock, A. J. Inorg. Chem. 2004, 43, 1816. 27. Doran, M. B.; Stuart, C. L.; Norquist, A. J.; O?Hare, D. Chem. Mater. 2004, 16, 565. 28. Obbade, S.; Dion, C.; Rivenet, M.; Saadi, M.; Abraham, F. J. Solid State Chem. 2004, 177, 2058. 29. Li, Y.; Cahill, C. L.; Burns, P. C. Chem. Mater. 2001, 13, 4026. 30. Sheldrick, G. M. SHELXTL PC, Version 6.12, An Integrated System for Solving, Refining, and Displaying Crystal Structures from Diffraction Data; Siemens Analytical X-Ray Instruments, Inc.: Madison, WI 2001. 81 31. Sheldrick, G. M. SADABS 2001, Program for absorption correction using SMART CCD based on the method of Blessing: Blessing, R. H. Acta Crystallogr. 1995, A51, 33. 32. Buehlmann, P.; Pretsch, E.; Bakker, E. Chem. Rev. 1998, 98, 1593. 33. Brown, I. D.; Altermatt, D. Acta Crystallogr. 1985, B41, 244. 34. Brese, N. E.; O?Keeffe, M. Acta Crystallogr. 1991, B47, 192. 35. Ramik, R. A.; Sturman, B. D.; Dunn, P. J.; Povarennykh, A. S. Can. Mineral. 1980, 18, 185. 36. Hawthorne, F. C. Tsch. Mineral. Petro. Mitt. 1983, 31, 121. 37. Attfield, M. P.; Morris, R. E.; Burshtein, I.; Campana, C. F.; Cheetham, A. K. J. Solid State Chem. 1995, 118, 412. 38. Cavellec, M.; Riou, D.; Greneche, J.-M.; F?rey, G. Inorg. Chem. 1997, 36, 2187. 39. Lethbridge, Z. A. D.; Lightfoot, P.; Morris, R. E.; Wragg, D. S.; Wright, P. A.; Kvick, A.; Vaughan, G. J. Solid State Chem. 1999, 142, 455. 40. Walton, R. I.; Millange, F.; O?Hare D.; Paulet, C.; Loiseau, T.; F?rey, G. Chem. Mater. 2000, 12, 1977. 41. Mahesh, S.; Green, M. A.; Natarajan, S. J. Solid State Chem. 2002, 165, 334. 42. Mandal, S.; Green, M. A.; Natarajan, S. J. Solid State Chem. 2004, 177, 1117. 43. Harvey, H. G.; Teat, S. J.; Tang, C. C.; Cranswick, L. M.; Attfield, M. P. Inorg. Chem. 2003, 42, 2428. 44. Burns, P. C.; Ewing, R. C.; Hawthorne, F. C. Can. Mineral. 1997, 35, 1551. 45. Almond, P. M.; Peper, S. M.; Bakker, E.; Albrecht-Schmitt, T. E. J. Solid State Chem. 2002, 168, 358. 82 46. Burns, P. C. Am. Mineral. 2000, 85, 801. 47. Kissick, J. L.; Cowley, A. R.; Chippindale, A. M. J. Solid State Chem. 2002, 167, 17. 48. a) Dieckmann, G. H.; Ellis, A. B. Solid State Ionics 1989, 32/33, 50. b) Vochten, R. Am. Mineral. 1990, 75, 221. c) Benavente, J.; Ramos Barrado, J. R.; Cabeza, A.; Bruque, S.; Martinez, M. Colloids Surf. 1995, A97, 13. 49. Obbade, S.; Dion, C.; Saadi, M.; Abraham, F. J. Solid State Chem. 2004, 177, 1567. 50. Shraveva, T. Y.; Almond, P. M.; Albrecht-Schmitt, T. E. J. Solid State Chem. 2004, 178, 499. 83 CHAPTER 4 SYNTHESIS, STRUCTURES, AND PROPERTIES OF THE MULTIFUNCTIONAL URANYL VANADIUM PHOSPHATE FRAMEWORKS ABSTRACT Two novel framework uranyl vanadium phosphate materials were synthesized under hydrothermal conditions. Similar reactions of UO2(NO3)2?6H2O, V metal, H3PO4 and various alkaline metal chlorides CsCl, KCl, or RbCl yielded different types of structures. Three-dimensional Cs2[UO2(VO2)2(PO4)2]?nH2O (UVVP) is constructed from uranyl phosphate chains, that are joined together by chains of corner-sharing VO5 distorted square pyramids. Channels along the b and c axes are filled with Cs+ cations. Alignment of terminal V=O bonds in all VO5 units along the c axis creates the overall polarity of the structure. Due to the polar structure, UVVP exhibits non-linear optical properties. Isostructural K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP) and Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP) contain VIV in an octahedral coordination environment. Their frameworks are constructed from vanadyl phosphate layers, interconnected by uranyl tetragonal bipyramids. There are channels, filled with K+ or Rb+, in all three perpendicular directions. Antiferromagnetic ordering at 52 K and Cs+- selective ion-exchange properties were found and characterized. 84 INTRODUCTION Rich and diverse crystal chemistry of the vanadium compounds arises from the ability of trivalent, tetravalent and pentavalent vanadium to adopt a wide range of coordination environments, from tetrahedral through trigonal bipyramidal and square pyramidal to octahedral.1 For instance, the family of vanadium phosphates, most recognized for their magnetic and catalytic properties,2,3 is composed of more than 130 members with one, two or three-dimensional crystal structures.2-5 To a certain degree, this structural flexibility is replicated in the class of uranyl-containing vanadium compounds. In the layered compounds with the structure of mineral carnotite of the general formula M2(UO2)2V2O8 nH2O,6-8 (M = K, Na, Ag, etc.) the centrosymmetric V2O86- groups are formed from two edge-shared VO5 square pyramids with terminal oxygen atoms pointed in opposite directions. Three-dimensional A(UO2)2(VO4)3 compounds, (A = Li and Na),9 that have the cation conductivity properties, comparable with NaSiCON and LiSiCON materials, as well as Pb(UO2)(V2O7),10 also crystallized in the centrosymmetric space groups I41/mad and P21/n respectively, and contain vanadium tetrahedra in the structure. On the other hand, the structure of K2[(UO2)2(VO)2(IO6)2O] H2O,11 composed completely of only distorted octahedra, including distorted VO6, is polar and crystallized in the Pba2 space group. Due to the polarity of the structure this material demonstrates non-linear optical properties. In all described uranyl-containing vanadium materials, either with layered or three-dimensional structures, vanadium is present in the +5 oxidation state. Numerous vanadium phosphates, that incorporate only tetravalent vanadium, or mixed valent VIV/VV units, such as -LiVIVOPO4,12 Zn2(VIVO)(PO4)2,13 Ba3VIV/VVO3(PO4)3,14 85 Na5(VIVO)2(PO4)3 H2O,15 and many others2 demonstrate paramagnetic behavior in the range of temperatures between 50 and 300 K, due to the presence of d1 unpaired electron on the vanadium center, and antiferromagnetic or ferromagnetic ordering at temperatures 5-40 K as a consequence of different super-exchange mechanisms of the various types of V-P-O chains present in the structures. In this chapter we describe the structure of two novel framework uranyl vanadium phosphates: Cs2[UO2(VO2)2(PO4)2]?nH2O (UVVP) and isostructural K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP) and Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP), along with their physical properties: the thermal behavior, second harmonic generation (SHG) of laser light of UVVP, selective ion-exchange properties and magnetic properties for KUVIVP and RbUVIVP. EXPERIMENTAL Syntheses. UO2(NO3)2?6H2O (98%, Alfa-Aesar), V powder (99.99%, Sigma Aldrich), H3PO4 (98%, Aldrich), KCl (99.997%, Alfa-Aesar), RbCl (99.975%, Alfa- Aesar), and CsCl (99.9%, Cerac) were used as received. V powder stored in the glove box was loaded in argon atmosphere. The reactions were run in PTFE-lined Parr 4749 reaction vessels. Distilled and Millipore filtered water with a resistance of 18.2 MW?cm was used in all experiments. Cs2[UO2(VO2)2(PO4)2]?nH2O (UVVP). UO2(NO3)2?6H2O (0.289 g, 0.576 mmol), H3PO4 (0.169 g, 1.72 mmol), CsCl (0.484 g, 2.88 mmol), and 1 mL of Millipore- filtered water were mixed in the 23 mL autoclave, followed by addition of V metal powder (0.080 g, 1.16 mmol), separately weighed in the glove box. The reaction was run 86 in sealed autoclave at 190 ?C in a box furnace for 5 days and then cooled at a rate of 9 ?C/hour to room temperature. Clusters of large bright yellow crystals of UVVP, reaching several mm in length, were isolated, washed with methanol, and allowed to dry. Yield, 515 mg (98% yield based on U). K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP). UO2(NO3)2?6H2O (0.395 g, 0.787 mmol), H3PO4 (0.231 g, 2.36 mmol), KCl (0.293 g, 3.94 mmol) and V (0.080 g, 1.6 mmol) were mixed in 23 mL PTFE-lined autoclave with 1 mL of water. Then reaction was heated at 190 ?C for five days and cooled to room temperature at a rate of 9 ?C/hour. After decantation of the mother liquid the only solid product of the reaction - bright blue prismatic crystals of KUVIVP were revealed. The yield of KUVIVP was 0.340 g (40.5 % yield based on the U). Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP). UO2(NO3)2?6H2O (0.334 g, 0.665 mmol), H3PO4 (0.196 g, 1.99 mmol), RbCl (0.402 g, 3.33 mmol) and V (0.068 g, 1.3 mmol) were mixed in 23 mL PTFE-lined autoclave followed by addition of 1 mL of water. After heating at 190 ?C for five days and cooling at a 9 ?C/hour to room temperature, the reaction vessel was opened and the product mixture, containing blue prism crystals of RbUVIVP and yellow powder product, was washed with methanol and dried at the air. The yield of RbUVIVP was 0.426 mg (50.9 % yield based on the U). Crystallographic Studies. Crystals of UVVP, KUVIVP, and RbUVIVP were mounted on glass fibers and aligned on a Bruker SMART APEX CCD X-ray diffractometer. Intensity measurements were performed using graphite monochromated Mo Ka radiation from a sealed tube and monocapillary collimator. SMART (v 5.624) was used for preliminary determination of the cell constants and data collection control. 87 The intensities of reflections of a sphere were collected by a combination of 3 sets of exposures (frames). Each set had a different f angle for the crystal and each exposure covered a range of 0.3? in w. A total of 1800 frames were collected with an exposure time per frame of 30 s for the crystals of UVVP, KUVIVP, and RbUVIVP. Determination of integrated intensities and global refinement were performed with the Bruker SAINT (v 6.02) software package using a narrow-frame integration algorithm. A face-indexed analytical absorption correction was initially applied using XPREP, where individual shells of unmerged data were corrected analytically.16 These files were subsequently treated with a semiempirical absorption correction by SADABS.17 The program suite SHELXTL (v 6.12) was used for space group determination (XPREP), direct methods structure solution (XS), and least-squares refinement (XL).16 The final refinements included anisotropic displacement parameters for all atoms. Secondary extinction was not noted for the crystals of UVVP, KUVIVP, and RbUVIVP. Some crystallographic details are given in Table 4.1. Atomic coordinates, symmetry information, displacement parameters, and occupancy for all atoms can be found in the Tables 4.2, 4.3, and 4.4 for UVVP, KUVIVP, and RbUVIVP respectively. Thermal analysis. The data were collected using differential scanning calorimeter TA Instrument, model 2920. The samples of known mass encapsulated in an aluminum container, were heated under nitrogen atmosphere from 30 ?C to 600 ?C with a heating rate of 10 ?C /minute. Non-linear optical properties. SHG was investigated using 1064 nm excitation laser pulses from a Q-switched Nd:YAG laser (Continuum Surelite I-10). A 88 Table 4.1. Crystallographic Data for Cs2[UO2(VO2)2(PO4)2]?nH2O (UVVP), K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP) and Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP) Compound UVVP KUVIVP RbUVIVP Formula Mass 901.99 1106.07 1243.26 Color and habit yellow prism blue plate blue plate Crystal System orthorhombic orthorhombic Orthorhombic Space group Cmc21(No. 36) Immm (No.71 ) Immm (No.71 ) a (?) 20.7116(14) 7.3803(7) 7.4370(7) b (?) 6.8564(5) 9.1577(8) 9.1662(8) c (?) 10.5497(7) 17.0898(16) 17.2415(15) V (?3) 1498.13(18) 1155.04(18) 1175.34(18) Z 4 2 2 T (K) 193 193 193 l (?) 0.71073 0.71073 0.71073 Maximum 2q (deg.) 56.62 56.58 56.74 rcalcd (g cm?3) 3.999 3.180 3.513 m(Mo Ka) (cm?1) 170.76 96.48 158.39 R(F) for Fo2 > s(Fo2) a 0.0265 0.0533 0.0551 Rw(Fo2) b 0.0602 0.1407 0.1461 a ( )R F F F F= -  o c o . b ( ) ( )R F w F F wF w o 2 o 2 c 2 2 o 4 1 2 = -    . 89 Table 4.2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs2[UO2(VO2)2(PO4)2]?nH2O (UVVP) Atom Symmetrya x y z Ueq (?2) b Occup. U(1) 4 a m 0 0.0326(5) -0.02493(4) 0.01259(10) 1 V(1) 8 b 1 -0.25770(5) -0.2961(16) 0.18303(13) 0.0134(3) 1 P(1) 8 b 1 -0.12106(9) -0.0815(3) 0.2254(2) 0.0132(4) 1 O(1) 8 b 1 -0.0809(3) -0.1215(9) 0.3441(5) 0.0215(12) 1 O(2) 8 b 1 -0.1609(2) 0.1081(8) 0.2442(6) 0.0207(11) 1 O(3) 8 b 1 -0.0767(3) -0.0678(9) 0.1113(6) 0.0241(13) 1 O(4) 8 b 1 -0.1657(2) -0.2593(8) 0.2052(5) 0.0160(10) 1 O(5) 8 b 1 -0.2709(3) -0.0641(7) 0.2323(6) 0.0195(13) 1 O(6) 8 b 1 -0.2648(3) -0.2942(9) 0.0320(6) 0.0232(13) 1 O(7) 4 a m 0 -0.2040(12) -0.0944(7) 0.0196(16) 1 O(8) 4 a m 0 0.2726(11) -0.9540(9) 0.0184(16) 1 O(9) 4 a m 0 0.544(3) 0.241(3) 0.068(11) 0.59(5) Cs(1) 8 b 1 -0.12432(2) -0.46363(7) -0.05798(4) 0.02101(15) 1 a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 90 Table 4.3. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP) Atom Symmetrya x y z Ueq (?2) b Occup. U(1) 2 a mmm ? ? ? 0.0078(3) 1 V(1) 8 m m 0.7058(3) ? 0.76487(11) 0.0092(4) 1 P(1) 8 l m ? 0.7579(3) 0.66892(17) 0.0081(6) 1 O(1) 4 e 2mm 0.7435(18) ? ? 0.016(2) 1 O(2) 8 l m ? 0.6805(10) 0.5902(5) 0.0166(18) 1 O(3) 8 l m ? 0.6429(10) 0.7368(5) 0.0123(17) 1 O(4) 16 o 1 0.6721(9) 0.8479(7) 0.6743(4) 0.0155(13) 1 O(5) 8 m m 0.8080(13) ? 0.6825(5) 0.0175(19) 1 O(6) 4 i mm2 ? ? 0.8660(9) 0.023(3) 1 K(1) 8 l m 0 0.3052(8) 0.4122(3) 0.048(2) 0.751(18) K(2) 2 d mmm ? 0 ? 0.114(9) 1 a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 91 Table 4.4. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP) Atom Symmetrya x y z Ueq (?2) b Occup. U(1) 2 a mmm ? ? ? 0.0072(3) 1 V(1) 8 m m 0.7052(3) ? 0.76599(12) 0.0083(5) 1 P(1) 8 l m ? 0.7571(3) 0.66934(17) 0.0082(6) 1 O(1) 4 e 2mm 0.7397(19) ? ? 0.015(3) 1 O(2) 8 l m ? 0.6779(11) 0.5910(5) 0.0145(19) 1 O(3) 8 l m ? 0.6432(10) 0.7367(5) 0.0108(18) 1 O(4) 16 o 1 0.6711(9) 0.8473(7) 0.6742(4) 0.0128(13) 1 O(5) 8 m m 0.8047(13) ? 0.6842(6) 0.017(2) 1 O(6) 4 I mm2 ? ? 0.8644(9) 0.023(3) 1 Rb(1) 8 l m 0 0.3083(4) 0.41202(15) 0.0313(10) 0.608(8) Rb(2) 2 d mmm ? 0 ? 0.0724(16) 1 a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 92 Scientech volume absorbing calorimeter was used to measure average laser power. SHG at 532 nm was visually observed in a darkened room from a polycrystalline sample of UVVP contained in a glass tube and was quantified using a band-pass optical filter, IR insensitive photomultiplier (1P28), and a signal averaging digital storage oscilloscope (Tektronix TDS 640A), as the intensity of the unfocused laser beam was increased above the SHG threshold. At a beam cross section averaged intensity of 3.6 MW/cm2, the SHG intensity from UVVP was 1500 times weaker that the signal observed from a commercial rare-earth-based ceramic frequency upconverting laser beam finder (Kentek View-It). Magnetic properties. Magnetic susceptibility measurements were conducted using a Quantum Design superconducting interference device (SQUID), under applied fields ranging from 100 to 10000 gauss as a function of temperature over the range of 5 ? 320 K and magnetization data were collected at 5K in fields ranging from 0 ? 10,000 gauss. Samples were doubly encapsulated in aluminum containers, which contributed significantly to the measured signal, especially at higher temperatures and for samples with a low inherent susceptibility. Empty containers were run independently under the same conditions and the data used directly for background subtraction. Diamagnetic contributions to the data were also calculated and used to correct the measured susceptibility. The temperature dependent magnetic susceptibility data were analyzed assuming Curie-Weiss behavior with an added temperature independent term, c = C / (T-q) + cTIP, where C and q are Curie and Weiss constants and cTIP is the temperature-independent contribution to the susceptibility. C is related to the effective magnetic moment 93 2/1 2 0 3  = B eff N kC mm in which N0 is Avogadro?s number and mB is the Bohr magneton (0.927 x 10-20 erg/Gauss). For cases in which cTIP is small, the effective magnetic moment can be obtained directly from the slope of the temperature dependence versus inverse magnetic susceptibility. Ion-Exchange Studies. NaCl, KCl, CsCl, CaCl2 and SrCl2 solutions with concentrations of 1?10-3 M and 1?10-2 M were used in order to estimate ion-exchange capacities of all materials. Ion-exchange properties, in actual fact, concentrations of the solutions of different cations were measured with two different methods: classic potentiometric method with ion-selective electrodes for K+, Na+ and Cs+ uptake,18 using Accumet A15 pH meter, and atomic adsorption spectroscopy, using Varian 240 AA spectrometer with hollow cathode Ca and Sr lamps for Ca2+ and Sr2+ uptake. Ion-selective electrodes were made in accordance with Fluka-recommended compositions and silver/silver chloride electrode as a reference.18 In order to prevent contamination of tested solutions with KCl the reference electrode with double junction and lithium acetate as a bridging electrolyte were utilized. In all cases samples were equilibrated in capped 5 mL vials with different cation solutions of known concentration for 48 hours at constant shaking, and then the change in the concentrations was determined. The approximate volume to mass ratio 200:1 (4 mL of the solution for 0.02 g of sample) was used in all experiments. 94 RESULTS AND DISCUSSION Syntheses. The utilization of the same reaction with only varied alkali metal chloride, CsCl - in the case of UVVP, KCl - KUVIVP, and RbCl - RbUVIVP, reveals the products, distinct not only in oxidation state and coordination number of vanadium, but also in the construction pattern of the structures. All our attempts to synthesize the identical to UVVP structure with K+ and Rb+ charge-balancing cations, as well as KUVIVP and RbUVIVP in the Cs-form, or even in the mixed-cations form failed. But at the same time KUVIVP and RbUVIVP did not undergo any structural changes after substitution of K+ or Rb+ with Cs+ by ion-exchange. We expect that the reason for this phenomena can be found in the thermodynamics of the synthetic reactions. In situ scanning microcalorimetry with slowly increasing temperature may be useful in order to solve this discrepancy, because it will give a possibility to observe all heat effects that correspond to all stages of synthesis of both UVVP and K(Rb)UVIVP. Running of multiple identical reactions also helps to identify every intermediate product of the reactions at different temperatures and make conclusions about driving forces and limitations and thus about major thermodynamic factors for both reactions.19-21 Vanadium metal powder was used as a source of vanadium. This synthetic method was previously studied with Ti and V, and in both cases yielded mixed-valent TiIII/TiIV and VIII/VIV compounds.22,23 We suggest that for oxidation of V0/VIII, protons from the phosphoric acid serve as oxidizer with subsequent release of hydrogen gas. In further oxidation of VIII to VIV and VV oxygen from air, present in the reaction vessel, or dissolved in water, can possible be involved. 95 Structures. Cs2[UO2(VO2)2(PO4)2]?nH2O (UVVP). The structure of UVVP consists of uranyl cations bound by phosphate to yield UO6 tetragonal bipyramids. These units are bridged by phosphate to yield one-dimensional chains that run down the c axis. This basic one-dimensional topology is recognized to occur in several uranyl phases including Cu2[UO2(PO4)2].24 The uranyl phosphate chains are in turn linked by chains of corner-sharing VO5 distorted square pyramids that run down the b axis into a novel open- framework structure that has intersecting channels that occur along the b and c axes as is shown in Figure 4.1. The channels running down the b axis are approximately 5.5 ? x 10.9 ?, whereas those running down the c axis are 5.0 x 9.5 ?. These channels are occupied by Cs+ cations and water molecules. One the most apparent features of this structure is that the VO5 units are aligned so that the single terminal oxo atoms of these polyhedra are all aligned along the c axis as is shown in Figure 4.1. Cs2[UO2(VO2)2(PO4)2]?nH2O crystallizes in polar orthorhombic space group Cmc21. As such, c is the polar axis, and the origin of this polarity lies in the alignment of the vanadyl units. The alignment of the vanadyl units is reminiscent of the alignment of titanyl units in the key NLO material KTiOPO4 (KTP).25, 26 The selected U-O, V-O and P-O bond distances, and angles are shown in Table 4.5. The UO6 units in Cs2[UO2(VO2)2(PO4)2]?nH2O display two short U=O bond distances of 1.772(7) and 1.781(8) ?, that define the uranyl cation. These bond distances are within the typical range for U=O distances. Longer U?O bonds to the equatorial oxygen atoms from the phosphate anions range from 2.250(6) to 2.256(5) ?. These distances were used to calculate a bond-valence sum for the U center of 6.32, which is consistent with U(VI).28 96 Figure 4.1. Depiction of the channels of UVVP, extended along b and c axes, occupied by Cs+ cations. Green ? uranium, blue ? vanadium, yellow ? phosphorus, purple ? cesium centers. 97 Table 4.5. Selected Bond Distances (?) and Angles (?) for Cs2[UO2(VO2)2(PO4)2]?nH2O (UVVP) Distances (?) U(1)-O(8) 1.772(7) V(1)-O(4) 1.936(5) U(1)-O(7) 1.781(8) V(1)-O(5?) 1.999(5) U(1)-O(3) 2.250(6) P(1)-O(3) 1.518(6) U(1)-O(1) 2.256(5) P(1)-O(1) 1.527(5) V(1)-O(6) 1.600(6) P(1)-O(4) 1.545(5) V(1)-O(5) 1.696(5) P(2)-O(2) 1.552(6) V(1)-O(2) 1.921(5) Angles (?) O(8)?U(1)?O(7) 177.5(4) O(5)?V(1)?O(2) 94.4(3) O(5)?V(1)?O(6) 106.4(3) O(5)?V(1)?O(4) 89.9(2) 98 The VO5 units contain VO2+ vanadyl cations with two short V=O bonds of 1.600(6) and 1.696(5) ?. The shorter of these bonds is to the terminal oxo atom. The remaining three V?O bonds range from 1.921(5) to 1.999(5) ?, yielding a bond-valence sum for the V atom of 5.08.28, 29 The tetrahedral phosphate anion shows typical P?O distances ranging from 1.518(6) to 1.552(6) ?. Finally, charge balance for the anionic lattice formed by the UO6, VO5, and PO4 units is maintained by the Cs+ cations that form contacts with surrounding oxygen atoms that occur from 3.105(6) to 3.395(6) ?. There are also partially occupied water molecules within smaller channels in the structure. There are 0.59 water molecules per formula unit. The fundamental building unit in UVVP is shown in Figure 4.2. K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP) and Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP). KUVIVP and RbUVIVP have similar framework structures based on the interconnection of vanadyl phosphate layers by uranyl tetragonal bipyramids. The fundamental building unit of KUVIVP and RbUVIVP is shown in Figure 4.3. Selected bond distances and angles are given in the Table 4.6. The vanadyl phosphate layer includes one crystallographically unique vanadium atom connected to six oxygen atoms. This connection leads to the formation of distorted coordination octahedra with one short vanadyl V?O(5) distance 1.598(9) ? (1.592(10) ?), one long V ? O6 bond - 2.301(11) ? (2.282(12) ?) in trans position to vanadyl, and four equatorial bonds with lengths ranging from 1.958(6) ? to 2.061(6) ? (from 1.966(7) ? to 2.075(6) ?). Distances and angles given in parenthesis represent the data for RbUVIVP.The bond valency sum calculated from these distances confirms vanadium +4 oxidation state.28 Close packing of two 99 Figure 4.2. The fundamental building unit of UVVP shown in 50% probability thermal ellipsoids. 100 Figure 4.3. Thermal ellipsoid plot of the fundamental unit of K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP) and Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP) at 50% probability. 101 Table 4.6. Selected Bond Distances (?) and Angles (?) for K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP). Data for Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP) are given in parathensis. Distances (?) U(1)-O(1) 1.797(13) (1.783(14)) V(1)-O(6) 2.301(11) (2.282(12)) U(1)-O(2) 2.261(9) (2.262(9)) V(1)-V(1) 3.038(4) (3.053(4)) V(1)-O(5) 1.598(9) (1.592(10)) P(1)-O(4) 1.517(6) (1.520(7)) V(1)-O(4) 1.958(6) (1.966(7)) P(1)-O(2) 1.520(9) (1.534(10)) V(1)-O(3) 2.061(6) (2.075(6)) P(2)-O(3) 1.567(9) (1.562(9)) Angles (?) O(8)?U(1)?O(7) 180.00(1) (180.00(2)) O(4)?V(1)?O(3) 90.6(3) (91.2(3)) O(5)?V(1)?O(6) 166.9(4) (165.7(5)) V(1)?O(6)?V(1) 82.6(5) (83.9(5)) 102 vanadium centers with V-V distance 3.038(4) ?(3.053(4) ?) results in the formation of face-shared V2O9 dimers where two equatorial O(4) and axial O(6) atoms belong to both coordination polyhedra.(Figure 4.4) These dimers were previously described in the literature, and for the first time were reported as a part of the (VO)2H4P2O9 structure.30,31 (VO)2H4P2O9 is a key precursor for the synthesis of vanadyl pyrophosphate (VO)2P2O7, which is a well known technologically important catalyst in the oxidation of butane to the maleic anhydride3,32 that has been studied in detail. Six equatorial oxygen atoms of the V2O9 building units are bridged to the six other vanadium dimers through phosphate groups, while two vanadyl O(5) and one opposite axial O(6) atoms remain terminal. There is an undefined electron density area located at the distance 1.55(2) ? from O(6). We speculate that the density could be assigned to one more, highly disordered H2PO4 phosphate group, but there is not enough structural information to prove this suggestion. Each PO4 group is tetrahedral with three typical P-O bonds 1.517(6) ? - 1.520(9) ? (1.520(7) ? - 1.534(10) ?) and one longer 1.567(9) ? (1.562(9) ?) bond, which corresponds to the 3-oxo bridge. The interaction of V2O9 and PO4 units create the layer in the [ab] plane, as is shown in the Figure 4.4. The vanadium phosphate layers with similar topology are found in the structures of (NH4)Zn(H2O)(VO)2(PO4)2(H2PO4)33 and MZn(H2O)(VO)2(PO4)2(H2PO4) (M = K+, Rb+, and Cs+).34 It is important to note that highly disordered phosphate groups in these structures are also found at the same location relative to the vanadium phosphate layer, as undefined electron density in KUVIVP and RbUVIVP. The layers are joined together by uranyl tetragonal bipyramids. There is one 103 Figure 4.4. Depiction of essential building blocks of the structures of KUVIVP and RbUVIVP. V2O9 dimer, constructed from two vanadium octahedra, sharing two equatorial and one axial oxygen atoms (top) and vanadyl phosphate layer (bottom). Cations (purple) occupy the channels 2.8 ? ? 4.4 ? along c axis. Blue - vanadium polyhedra, yellow ? phosphate groups. 104 crystallographically unique uranium center in the structure that belongs to linear UO22+ cation and indeed has two axial uranyl oxygen atoms with a distance of 1.797(13) ? (1.783(14) ?) and an angle of 180.000(1)? (180.000(2)?). Four equatorial oxygen bonds with lengths 2.261(9) ? (2.262(9) ?) are bridged to phosphate groups of vanadium phosphate layer creating a negatively charged framework of [(UO2)(VO)4(OH)2(PO4)4]4- with a system of channels, parallel to the a, b, and c axes, filled with K+ (Rb+) cations. The channels along the a and b axes are shown in the Figures 4.5 and 4.6 respectively. There are two crystallographically distinct cationic sites in this structure. One is fully occupied with an alkali metal cation, either K+, or Rb+, coordinated to 12 oxygen atoms with coordination distances ranging from 3.163(8) ? to 3.308(9) ? (3.343(10) ? - 3.550(7) ?). The other cation is in an 8-coordinate environment. It is disordered and only partially occupied with 0.751(18) of K+ and 0.608(8) of Rb+, in the unit cell. This partial occupancy creates the small positive charge defficiency, that can be neutralized by protonation of terminal oxygen atoms of the framework and again, could be ascribed to the presence of an additional disordered phosphate group in the structure. Finally, the formula of the compound is proposed as ideal M4[(UO2)(VO)4(OH)2(PO4)4], that does not reflect the partial occupancy of the cationic sites. Thermal analysis. We have studied the thermal behavior of UVVP and KUVIVP as potential Cs+ and Sr2+ selective ion-exchange materials, in order to test their ability to preserve the structure at high-temperature conditions of actual nuclear waste. DSC data for the UVVP reveal two significant endothermic events upon an increase in temperature to 600 ?C. (Figure 4.7) The first peak occurs at 150 ?C, which can 105 Figure 4.5. Depiction of the channels in the framework of KUVIVP and RbUVIVP, filled with K+ or Rb+ cations, extended along the a axis. Blue- vanadium, yellow ? phosphorus, green ? uranium centers, purple ? K+ or Rb+ cations. 106 Figure 4.6. A view of the framework of KUVIVP and RbUVIVP along the b axis. 107 -2 -1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0 30 130 230 330 430 530 Temperature, oC H e a t f l o w , m W . . 150oC 570oC 510oC Figure 4.7. DSC data for UVVP demonstrate thermal stability up to 510 ?C. 108 be assigned to the removal of the surface and interstitial 0.59 water molecules. The second event, composed of two peaks, the smaller at 510 ?C and the larger ? at 550 ?C, corresponds to the decomposition of material. We can suggest that the structure of UVVP remains stable and maintains crystallinity at the temperatures under 500 ?C. The data for KUVIVP demonstrate the loss of water content at 160 ?C, and a series of sequential reconstructions of the framework at 310 ?C, 330 ?C, 510 ?C and 550 ?C. (Figure 4.8) Amorphization of the structure after heating to 600 ?C is confirmed by powder diffraction. Non-linear optical properties. One of the consequences of the polarity in the structure of UVVP is that the compound should exhibit the SHG of laser light; that is, it should act as a NLO material. SHG was investigated using 1064 nm excitation laser pulses from a Q-switched Nd:YAG laser. SHG at 532 nm was observed from a polycrystalline sample of UVVP. The response is substantially weaker than that of a commercial ceramic frequency doubling laser beam finder (Kentek View-It). At the same time the sample of UVVP shows a 70 times larger response than -quartz. Thus, observed non-linear optical properties can confirm the polar nature of the structure, but UVVP cannot be considered as an effective NLO material. Magnetic properties. Magnetization of UVVP varies linearly with magnetic field, as is shown in Figure 4.9. The temperature dependence of the magnetic susceptibility of UVVP, measured in a constant field of 10000 G, is shown in Figure 4.10. The magnetic signal at high temperatures is almost equal to the background, consistent with the expectation that V in this compound is present as non-magnetic V5+. Our measurements 109 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0 30 130 230 330 430 530 Temperature, oC H e a t f l o w , m W . . 160 oC 310oC 330oC 510oC 550oC Figure 4.8. DSC data for KUVIVP. Several endothermic events between 160 ?C and 550 ?C reveal gradual structural transformations with increase of temperature. 110 Figure 4.9. Magnetization of UVVP as a function of applied magnetic field at 5 K. Line is the linear fit to the data. 111 Figure 4.10. Magnetic susceptibility of UVVP as a function of temperature with a 10000 G magnetic field. 112 indicate that effective moment is smaller than 0.24 B. If this magnetic contribution is arising from a V4+, and the spin only moment of 1.73 B is assumed, then the results are consistent with less than 2% of the total V as V4+. Magnetization measurements of RbUVIVP, obtained at 5 K, increase linearly with magnetic field up to the highest measured field, 10,000 G, as shown in Figure 4.11. About 75% of the magnetic signal arises from the background at high temperatures, necessitating susceptibility data acquisition in a 10,000 G field. The temperature dependence of the magnetic susceptibility of RbUVIVP, together with its inverse, is depicted in Figure 4.12. The data show complex behavior with temperature. Susceptibility data obtained at lower fields, 500 and 2000 Gauss show a similar temperature dependence. The data indicate a long-range ordering of the moments at 52(2) K. Above that temperature the sample of RbUVIVP appears paramagnetic and the data between 100 and 320 K are fit to the modified Curie law and the effective magnetic moment is determined to be 1.73(10) mB. The effective magnetic moment obtained from a Curie-Weiss fit to the temperature dependence of inverse magnetic susceptibility is 1.82(10) mB. The difference between these two values reflects the difference in emphasis of the lower or higher temperature data respectively and reflects the inherent error in working with such small magnetic signals. cTIP and q are also determined from the modified Curie law fit as -0.00008(4) emu/mole and -41(8) K, respectively. The large, negative Weiss constant is consistent with an antiferromagnetic component to the ordering, as is the shape of the c versus T plot. There is an additional feature in the data that occurs at about 15(1) K that could be a re-ordering of the moments as the temperature is lowered through the ordered state from 52 K, or it could be a paramagnetic 113 Figure 4.11. Magnetization of RbUVIVP as a function of applied magnetic field at 5 K. Line is the linear fit to the data. 114 Figure 4.12. Magnetic susceptibility of RbUVIVP as a function of temperature. Line is the fit to Curie-Weiss behavior. Inset shows the temperature dependence of the inverse magnetic susceptibility. 115 impurity phase. The effective moment at 5 K, determined from the M vs. H curve, is 0.34 mB. It is not possible to distinguish between these two options without further studies. Ion-exchange study. Due to the framework type of the structure, the ion- exchange experiments were run for both UVVP and KUVIVP. After equilibration of all samples of UVVP with NaCl, KCl, CaCl2, SrCl2 solutions the change of the color of solutions, decrease of the sample amount, and consequently, the partial dissolution of UVVP was detected. In contrast, the structure of KUVIVP in all exchanged probes, was conserved, as was proved by the identity of powder diffraction patterns of the samples, tested before and after ion-exchange experiments. There is no uptake of the cations found from the pure 1?10-2 M and 1?10-3 M solutions of CaCl2 and SrCl2. Small exchange was observed in 1?10-2 M NaCl solutions with an uptake value 0.25(6) meq/g. In the mixed 1:1 solutions of NaCl/CaCl2 and NaCl/SrCl2 slightly smaller Na+- exchange capacities of 0.18(5) meq/g and 0.17(5) meq/g were detected, probably owing not to the competition of Na+/Ca2+ and Na+/Sr2+ cationic pairs in the probe, but to the overall dilution of the Na+ solutions. Ca2+ and Sr2+ uptake were not found in the mixed solutions either. Exchange with Cs+ from 1?10-2 M CsCl solution reveals notable uptake 0.87(10) meq/g. The amount of the exchanged Cs+ cations per taken mass of the sample is close to 50% (46.8%?0.9%) of total cation content, in other words, half of the theoretical capacity, calculated from the formula. These results are in good agreement with the structural data. We suggest that the larger channels along the a and b axes with size 5.8 ? ? 4.6 ? and 3.8 ? ?4.6 ?, respectively, are not available for exchange due to the presence of undefined electron density, pointed inside of these channels. Therefore, we 116 propose that ion-exchange occurs only through smaller channels 2.8 ? ? 4.4 ? along the c axis, occupied by one out of two K+ cations, present in the structure. The high selectivity of the framework toward Cs+ versus K+ can be explained in the radii and hydration energy terms. At the hydrated state the radii of K+ and Cs+ are the same and correspond to 3.3 ?.35 Due to the lower dehydration energy Cs+ is able to lose its hydration sphere and enter the channels easier than K+. In the mixed Cs+/Na+ solutions with equal concentrations of both cations, again only half of the total cation content is exchanged. Three quarters of exchanged cations are Cs+ and only one quarter is Na+. If the concentration of Na+ is doubled in the initial mixed solutions, the ratio of exchanged Cs+/Na+ remains the same. From this data we can conclude that KUVIVP is Cs+-selective ion-exchange material with high capacity 0.87(10) meq/g. CONCLUSIONS Two novel uranyl vanadium phosphate framework materials have been synthesized and studied. One material is thermally stable up to 500 ?C and demonstrates non-linear optical properties. The other exhibits remarkable magnetic properties and selective Cs+ uptake from mixed Cs+/Na+ solutions, as would be required by actual nuclear waste conditions. Several issues need to be explained, such as structural uncertainty of K(Rb)UVIVP or the reason for the formation of different compounds under similar conditions. We are currently working on the detailed elemental analysis of RbUVIVP that will help to solve the structural uncertainties of KUVIVP and RbUVIVP members of uranyl vanadium phosphate family. 117 REFERENCES 1. Cotton, F. A.; Wilkinson, G.; Murillo, C. A.; Bochmann, M. Advanced Inorganic Chemistry , 6th ed.; Wiley-Interscience: New-York, 1999, pp 714-736. 2. Boudin, S.; Guesdon, A.; Leclaire, A.; Borel, M.-M. Int. J. Inorg. Mater. 2000, 2, 561. 3. Hutchings, G. J. J. Mater. Chem. 2004, 14, 3385. 4. Beltran-Porter, D.; Beltran-Porter, A.; Amoros, P.; Ibanez, R.; Martinez, E.; Le Bail, A.; Ferey, G.; Villeneuve, G. Eur. J. Solid State Inorg. Chem. 1999, 28, 131. 5. Amoros, P.; Marcos, D. M.; Beltran-Porter, A.; Beltran-Porter, D. Curr. Opin. Solid State Mat. Science 1999, 4, 123. 6. Abraham, F.; Dion, C.; Saadi, M. J. Mater. Chem. 1993, 3, 459. 7. Abraham, F.; Dion, C. ; Tancret, N. ; Saadi, M. Adv. Mater. Res. 1994, 1 2, 511. 8. Dickens, P. G.; Stuttard, G. P.; Ball, R. G. J.; Powell, A. V. ; Hull, S. ; Patat, S. J. Mater. Chem. 1992, 2, 161. 9. Obbade, S.; Dion, C.; Rivenet, M. ; Saadi, M.; Abraham F. J. Solid State Chem. 2004, 177, 2058. 10. Obbade, S.; Dion, C.; Saadi, M.; Yagoubi, S.; Abraham, F. J. Solid State Chem. 2004, 177, 3909. 11. Sykora, R. E.; Albrecht-Schmitt, T. E. Inorg. Chem. 2003, 42, 2179. 12. Lii, K. H.; Li, C. H.; Cheng, C. Y.; Wang, S. L. J. Solid State Chem. 1991, 95, 352. 13. Bayi, F.; Pourroy, G.; Belaiche, M.; Legall, P.; Drillon, M.; Kuentzler, R. Eur. J. Solid State Inorg. Chem. 1993, 30, 55. 118 14. Borel, M. M.; Leclaire, A.; Chardon, J.; Michel, C.; Provost, J.; Raveau, B. J. Solid State Chem. 1998, 135, 302. 15. Roca, M.; Amoros, P.; Cano, J.; Marcos, M. D.; Alamo, J.; Beltran-Porter, A.; Beltran-Porter, D. Inorg. Chem. 1998, 37, 3167. 16. Sheldrick, G. M. SHELXTL PC, Version 6.12, An Integrated System for Solving, Refining, and Displaying Crystal Structures from Diffraction Data; Siemens Analytical X-Ray Instruments, Inc.: Madison, WI 2001. 17. Sheldrick, G. M. SADABS 2001, Program for absorption correction using SMART CCD based on the method of Blessing: Blessing, R. H. Acta Crystallogr. 1995, A51, 33. 18. Buehlmann, P.; Pretsch, E.; Bakker, E. Chem. Rev. 1998, 98, 1593. 19. Navrotsky, A. Curr. Opin. in Colloid & Interface Science 2005, 10, 195. 20. Yang, S., Li, Q.; Wang, M.; Navrotsky, A. Micropor. Mesopor. Mater. 2006, 87, 261. 21. Yang, S.; Navrotsky, A.; Phillips, B. L. J. Phys. Chem. 2000, 104, 6071. 22. Ekambaram, S.; Serre, C.; Ferey, G.; Sevov, S. C. Chem. Mater. 2000, 12, 444. 23. Calin, N.; Serre, C.; Sevov, S. C. J. Mater. Chem. 2003, 13, 531. 24. Guesdon, A.; Charon, J.; Provost, J.; Raveau, B. J. Solid State Chem. 2002, 165, 89. 25. Masse, R.; Grenier, J. C. Bull. Soc. Fran. Mineral. Cristallogr. 1971, 94, 437. 26. Tordjman, I.; Masse, R.; Guitel, J. C. Z. Kristallogr., Kristallgeom., Kristallphys., Kristallchem. 1974, 139, 103. 27. Burns, P. C.; Ewing, R. C.; Hawthorne, F. C. Can. Mineral. 1997, 35, 1551. 28. Brese, N. E.; O?Keefe, M. Acta Crystallogr. 1991, B47, 192. 29. Brown, I. D.; Altermatt, D. Acta Crystallogr. 1985, B41, 244. 119 30. Torardi C. C.; Calabrese, J. C. Inorg. Chem. 1984, 23, 1308. 31. Leonowicz, M. E.; Johnson, J. W.; Brody, J. F.; Shannon H.; Newsam J. M. J. Solid State Chem. 1985, 56, 370. 32. Bordes, E.; Courtine, P. J. Catal. 1979, 57, 236. 33. Le Fur, E.; Pea, O.; Pivan, J. Y. J. Mater. Chem. 2002, 12, 132. 34. Messaoudi, S.; Furet, E.; Gautier, R.; Le Fur, E.; Pea, O.; Pivan, J. Y. Chem. Mater. 2004, 16, 435. 35. Nightingale, E.R. J. Phys. Chem. 1959, 63, 1381. 120 CHAPTER 5 HYDROTHERMAL SYNTHESIS AND STRUCTURAL CHARACTERIZATION OF Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O and Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O ABSTRACT Two novel uranyl-containing three-dimensional materials Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O (UTiP) and Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O (UGeP) were synthesized under hydrothermal conditions. The materials are isostructural and have identical frameworks, constructed from the perpendicular intersection of one- dimensional uranyl phosphate and titanium or germanium phosphate chains. The channels are occupied by disordered Cs+ cations and disordered water molecules. The compounds crystallize in monoclinic space group C2/m. Crystallographic data: UTiP: a = 19.715(4) ?, b = 7.1207(14) ?, c = 9.4756(19) ?,  = 115.93(3)?, V = 1196.3(4) ?3, Z = 4, T = 193K. UGeP: a = 19.5123(17) ?, b = 6.8639(6)?, c = 9.3731(8) ?,  = 116.125(2)?, V = 1127.09(17) ?3, Z = 4, T = 193K. 121 INTRODUCTION Potential technical applications of inorganic crystalline three-dimensional structures, due to their non-linear optical properties, magnetism, conductivity, or ion- exchange, leads to a continued search for novel compounds and methods for their preparation.1 Three-dimensional uranyl-containing materials have come to the forefront because there are thousands of tons of depleted uranium, with very low level radioactivity stored at the Savannah River Site and several other locations. The incorporation of UVI in the form of uranyl, UO22+ cations, into crystalline frameworks has the potential of yielding functional materials while at the same time expanding our knowledge on the crystal chemistry of UVI. The formation of uranyl-containing three-dimensional structures is a very challenging task due to the reduced dimensionality along axial uranyl-oxygen double bonds.29-30 There are several synthetic strategies known today to facilitate the formation of frameworks. We focused our study on the incorporation of octahedral building units, such as IO65?,28 in order to allow the bond formation in the direction parallel to uranyl axis. In previous chapters we have already shown that, indeed, the structure of UGaP125 and KUVIVP (RbUVIVP) are formed according to this new synthetic method. In this chapter we continue to investigate the conditions of the formation of frameworks by the introduction of Ti and Ge coordination polyhedra. The formation of six-coordinate complexes is common in the structural chemistry of Ti(IV). In the structures of ETS-4 and ETS-10 molecular sieves,31 Cs+ and Sr2+- selective titanium silicates Na2Ti2O3(SiO4) H2O and HM3Ti4O4(SiO4) H2O32 and many other compounds, as well as in recently synthesized organically templated framework 122 mixed-valent Ti(III)/Ti(IV) phosphates33,34 and mixed-valent Ti(III)/Ti(IV) gallium phosphate,35 Ti metal centers form octahedra with different degrees of distortion. In the latter structures of TiIIITiIV(PO4)(HPO4)2(H2O)2?0.5NH2CH2CH2CH2NH2, TiIIITiIV(HPO4)4?C2N2H9?H2O33,34 and [C5NH6][Ti3Ga(PO4)5(H2O)2]35 both Ti(III) and Ti(IV) have a six-folded coordination environment, with small differences in the bond length range. The knowledge about crystal chemistry of uranium-titanium oxocompounds is restricted to only a series of structures of UTixNb3-xO10 (x = 0.33, 1) and its intercalated compound Li0.9UTiNb2O10, where Ti plays only a substitution role,36 and in several mineral structures.37 The most abundant uranium-titanium mineral brannerite, UTi2O6, is an important component of the Synroc pyrochlore-type ceramics,38 studied as a storage material for Pu isotopes in a geological environment.39, 40 At room temperature naturally occurring brannerite is amorphous. It crystallizes only at temperatures above 1000 ?C in monoclinic C2/m structure, consisting of layers of titanium edge-sharing octahedra, linked together with uranium tetragonal bipyramids.41 One of the proposed mechanisms for the stabilization of brannerite-phase formation could be incorporation of doping cations such as Ca2+, La3+, and Ga3+ in the structure.42-44 We suggest that the formation of crystal structures based on uranium and titanium at hydrothermal temperatures can be reached through addition of phosphate groups, which improve the solubility of components and stabilize the resultant structure. The coordination polyhedra of germanium can also be used as potential octahedral building unit, due to the ability to have coordination numbers 4, 5, or 6, unlike its closest neighbor Si.45 This originates from larger Ge, then Si, atomic radius and 123 therefore longer Ge?O bonds and more flexible Ge-O-Ge angle. The variations in the coordination modes of germanium allows it to form structures incorporating tetrahedral,46,47 trigonal bipyramidal or square pyramidal48,49 and octahedral50-52 coordination environments, as well as structures where all three types of polyhedra are combined together53-55 and frameworks entirely constructed from germanate units.56-58 There are a number of features of the germanate frameworks that make this class of structures extremely diverse and attractive for detailed studies. First is the ability of germanate units to form clusters. [Ge9O18(OH)4] clusters synthesized with different organic templates57,59 and Ge9O19(OH)2(N2C2H10)2(N2C2H8)0.5H2O60 are just a few examples of three-dimensional structures with remarkable porosity due to the cluster construction. Second, the germanate framework compounds with odd-membered rings can be formed.50,52,61,62 Finally, the capacity to build the porous structures with chiral channels makes germanates attractive for enantioselective catalysis and separation.60,63,64 All known uranyl germanates have layered structures.65,66 The germanate anions there exhibit only tetrahedral arrangements. In this chapter, as a continuation of previous work, we intend to investigate the potential of Ti and Ge centers to form octahedral polyhedra in order to synthesize the first uranyl-containing titanium and germanium phosphate framework materials. EXPERIMENTAL Syntheses. UO2(NO3)2?6H2O (98%, Alfa-Aesar), Ti powder (99.99%, Sigma Aldrich), Ge powder (99.99%, Sigma Aldrich), H3PO4 (98%, Aldrich), and CsCl (99.9%, Cerac) were used as received. Ti and Ge powder were handled in the glovebox under an 124 Ar atmosphere due to the potential rapid oxidation in air. For all reactions PTFE-lined Parr 4749 vessels were loaded with reaction mixture and 1 mL of distilled and Millipore filtered water with a resistance of 18.2 MW?cm. Despite the presence of only depleted uranium in UO2(NO3)2?6H2O and products of the reactions, and therefore very low level of radioactivity, all studies were carried out with precautions. Semi-quantitative SEM/EDX analyses were performed using a JEOL JSM-7000F instrument. Cs, U, Ti, Ge, and P percentages were calibrated against standards. Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O (UTiP). UO2(NO3)2?6H2O (0.276 g, 0.549 mmol), H3PO4 (0.161 g, 1.64 mmol) and CsCl (0.554 g, 2.75 mmol) were combined together in 23 mL PTFE liner with 1 mL of water with stirring until homogenous. The Ti (0.026g, 0.55 mmol) was quickly added to the reaction mixture, the autoclave was rapidly sealed, placed in a box furnace, and heated at 195 ?C for seven days. Then reaction was slowly cooled with an average rate 1 ?C/hour. The solid part of the product mixture was washed with water and methanol and dried in air. The final composition of the mixture was identified as small fractions of pale yellow crystals of UTiP, mixed together with yellow crystals of uranyl phosphate and excess Ti metal, and larger portions of amorphous UTiP. The isolated yield of crystalline UTiP was 0.028 g (7.1 % yield based on the U). EDX analysis provided Cs:U:Ti:P ratio of 1.5:1:1:2. Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O (UGeP). UO2(NO3)2?6H2O (0.282 g, 0.561 mmol), H3PO4 (0.165 g, 1.64 mmol) and CsCl (0.472 g, 2.81 mmol) were mixed in a 23 mL PTFE-lined autoclave with 1 mL of water and Ge (0.081 g, 1.1 mmol) powder, weighed separately in the glovebox.. The sealed vessel was heated at 195 ?C for five days and cooled to room temperature at the rate of 9?C/hour. After decantation of the mother 125 liquid small yellow crystals of UGeP, grown on the surface of non-crystalline product were found. The yield of UGeP was near quantitative. EDX analysis provided Cs:U:Ge:P ratio of 1.5:1:1:2. Crystallographic Studies. Crystals of UTiP and UGeP were mounted on glass fibers and aligned on a Bruker SMART APEX CCD X-ray diffractometer. Intensity measurements were performed using graphite monochromated Mo Ka radiation from a sealed tube and monocapillary collimator. SMART (v 5.624) was used for preliminary determination of the cell constants and data collection controls. The intensities of reflections of a sphere were collected by a combination of 3 sets of exposures (frames). Each set had a different f angle for the crystal and each exposure covered a range of 0.3? in w. A total of 1800 frames were collected with an exposure time per frame of 120 s for both UTiP and UGeP due to the extremely small dimensions of the crystals. Determination of integrated intensities and global refinement were performed with the Bruker SAINT (v 6.02) software package using a narrow-frame integration algorithm. A face-indexed analytical absorption correction was initially applied using XPREP, where individual shells of unmerged data were corrected analytically.67 These files were subsequently treated with a semiempirical absorption correction by SADABS.68 The program suite SHELXTL (v 6.12) was used for space group determination (XPREP), direct methods structure solution (XS), and least-squares refinement (XL).67 The final refinements included anisotropic displacement parameters for all atoms. Secondary extinction was not noted for either crystal. Some crystallographic details are given in Table 5.1. Atomic coordinates and displacement parameters for all atoms of UTiP can be found in Table 5.2, UGeP ? in Table 5.3. 126 Table 5.1. Crystallographic Data for Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O (UTiP) and Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O (UGeP). Compound UTiP UGeP Formula Mass 740.16 742.28 Color and habit yellow plate yellow plate Crystal System Monoclinic Monoclinic Space group C2/m (No. 12) C2/m (No. 12) a (?) 19.715(4) 19.5123(17) b (?) 7.1207(14) 6.8639(6) c (?) 9.4756(19) 9.3731(8)  (?) 115.93(3) 116.125(2) V (?3) 1196.3(4) 1127.09(17) Z 4 4 T (K) 193 193 l (?) 0.71073 0.71073 Maximum 2q (deg.) 56.64 56.52 rcalcd (g cm?3) 4.110 4.374 m(Mo Ka) (cm?1) 190.8 213.6 R(F) for Fo2 > 2s(Fo2) 0.0371 0.0412 Rw(Fo2) b 0.0747 0.0977 a ( )R F F F F= -  o c o . b ( ) ( )R F w F F wF w o 2 o 2 c 2 2 o 4 1 2 = -    . 127 Table 5.2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O (UTiP). Atom Symmetrya x y z Ueq (?2) b s.o.f. U(1) 4 i m 0.04052(2) 0 -0.66748(5) 0.0109(1) 1 Ti(1) 4 e 1 ? ? 0 0.0156(4) 1 P(1) 4 i m 0.12604(16) 0 -0.9415(3)) 0.0115(6) 1 P(2) 4 i m 0.15447(16) 0 -0.3148(3) 0.0115(6) 1 O(1) 4 i m 0.1665(5) 0 -0.4635(9) 0.027(2) 1 O(2) 8 j 1 0.1866(4) 0.1742(8) -0.2172(7) 0.0287(15) 1 O(3) 4 i m 0.0670(4) 0 -0.3835(9)) 0.024(2) 1 O(4) 4 i m -0.0517(5) 0 -0.9137(10) 0.037(2) 1 O(5) 4 i m 0.1134(5) 0 -0.7976(10) 0.037(3) 1 O(6) 8 j 1 0.0413(4) 0.2490(8) -0.6652(7) 0.0269(14) 1 O(7) 8 j 1 0.1715(3) 0.1769(8) -0.9408(8) 0.0295(16) 1 O(8) 4 i m 0.2912(4) 0 -0.9540(9) 0.0139(16) 1 O(9) 4 i m 0.3640(7) 0 -0.2041(17) 0.047(5) 0.85(3) Cs(1) 4 i m 0.30542(5) 0 -0.5527(1) 0.0289(3) 0.874(3) Cs(2) 4 i m 0.4233(3) 0 -0.6394(7) 0.046(2) 0.194(4) Cs(3) 4 i m 0.45520(11) 0 0.0264(3) 0.0405(8) 0.457(3) a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 128 Table 5.3. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O (UGeP). Atom Symmetrya x y z Ueq (?2) b s.o.f. U(1) 4 i m 0.04126(2) 0.0000 -0.66805(5) 0.00962(15) 1 Ge(1) 4 e 1 ? ? 0 0.0174(3) 1 P(1) 4 i m 0.12860(19) 0 -0.9417(4) 0.0128(6) 1 P(2) 4 i m 0.15654(18) 0 -0.3081(4) 0.0110(6) 1 O(1) 4 i m 0.1692(6) 0 -0.4552(11) 0.027(2) 1 O(2) 8 j 1 0.1886(6) 0.1811(11) -0.2092(9) 0.050(3) 1 O(3) 4 i m 0.0689(6) 0 -0.3787(12) 0.055(5) 1 O(4) 4 i m -0.0523(7) 0 -0.9147(13) 0.052(4) 1 O(5) 4 i m 0.1180(6) 0 -0.7930(12 0.057(5) 1 O(6) 8 j 1 0.0404(4) 0.2571(12) -0.6643(11) 0.042(2) 1 O(7) 8 j 1 0.1756(6) 0.1797(12) -0.9392(17) 0.060(3) 1 O(8) 4 i m 0.2929(6) 0 -0.9512(11) 0.020(2) 1 O(9) 4 i m 0.3707(17) 0 -0.186(3) 0.062(10) 0.63(7) Cs(1) 4 i m 0.30512(7) 0 -0.54392(16) 0.0383(5) 0.812(6) Cs(2) 4 i m 0.4159(11) 0 -0.682(2) 0.069(7) 0.100(7) Cs(3) 4 i m 0.45559(14) 0 0.0046(4) 0.0396(10) 0.435(5) a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 129 RESULTS AND DISCUSSION Syntheses. The syntheses of UTiP and UGeP appear to be similar and yield isostructural compounds. Both Ti and Ge starting materials are pure elements. Ti powder was previously used in the production of several titanium phosphates,33-35 which often led to the partial oxidation of titanium and synthesis of mixed-valent materials. UTiP also could be synthesized from titanium isopropoxide Ti(OC3H7)4 as a source of Ti, but crystals were very small and could be seen only using SEM. We speculated that the choice of the source material can play a crucial role in crystal growth. In the study of synthetic conditions for the formation of jarosites, Grohol and Nocera68 suggested that slow oxidation of vanadium metal can be the rate-determining step in the crystallization of the final vanadium sulfate mineral. This could also be true for the development of UTiP and UGeP as well. Slow delivery of oxidized reactants to the reaction mixture of uranyl nitrate and phosphoric acid aids in the formation of a multicomponent system instead of separate titanium or germanium phosphates and uranyl phosphate, and gives the possibility of growing crystals with dimensions acceptable for single-crystal X-ray diffraction. Structure. The crystallographic data (Table 5.1) and atomic coordinates (Tables 5.2 and 5.3) of Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O (UTiP) and Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O (UGeP) are closely related. This indicates that UTiP and UGeP are isostructural compounds. The variation in the b axis of the unit cell (7.1207 ? for UTiP and 6.8639 ? for UGeP) can be explained in the terms of ionic radii of Ti+4 (0.61 ?) and Ge+4 (0.39 ? ) and, indeed, corresponds to the difference in the radii and average difference between Ti-O and Ge-O bond length (1.939 ? and 1.856 ? 130 respectively). The structures of UTiP and UGeP are closely related to the structure of Cs4[(UO2)2(GaOH)2(PO4)4]?H2O,25 described in Chapter 3. In general, it is a framework structure, built out of perpendicularly intersecting one-dimensional uranyl phosphate and titanium or germanium phosphate chains. The channels, formed along the c axis, are filled with Cs+ cations. The fundamental building unit of UTiP (or UGeP) is shown in Figure 5.1. Uranyl phosphate chains of UTiP and UGeP extended in the c direction, and have the same topology as uranyl phosphate chains in the structures of Cs4[(UO2)2(GaOH)2(PO4)4]?H2O and Cs[UO2Ga(PO4)2]25. (Figure 5.2) These chains have been previously described as 1? [UO2(PO4)2]4-25,70 They differ from Cs4[(UO2)2(GaOH)2(PO4)4]?H2O, in that there are two crystallographically unique uranium centers. In UTiP and UGeP structures only one unique uranium is present, and therefore only one uranyl phosphate chain is formed. This phenomenon can also be secondarily confirmed by the unit cell comparison. In Cs4[(UO2)2(GaOH)2(PO4)4]?H2O, the dimension of the unit cell axis, containing two crystallographycally unique uranyl phosphate chains, is doubled compare to UTiP and UGeP . The uranium center is connected to five phosphate oxygen atoms in the equatorial plane and two uranyl oxygen atoms, almost linearly with angles 178.7(4)? and 177.1(7)? for UTiP and UGeP, respectively, in the axial direction, to form a pentagonal bipyramid environment. The U-O distances in the equatorial plane are long and range between 131 Figure 5.1. Thermal ellipsoid plot of Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O (UTiP) shown at 50% probability. 132 Figure 5.2 Depiction of one-dimensional titanium phosphate or germanium phosphate (top) and uranyl phosphate (bottom) chains in UTiP or UGeP. Violet - titanium or germanium polyhedra, yellow-phosphate tetrahedra, green-uranyl building units. 133 2.245(9) ? and 2.507(8) ? for UTiP and between 2.227(11) ? and 2.521(10) ? for UGeP, compared to the short uranyl distances of 1.773(6) ? for UTiP and 1.765(8) ? for UGeP. Two uranyl pentagonal bipyramids share edges to form dimers, connected to the chain by phosphate anions. The bond-valence sums, calculated according to these data, correspond to uranium in the +6 oxidation state. Titanium and germanium phosphate chains 1? [Ti(OH0.48)(PO4)2]2? and 1 ? [Ge(OH0.65)(PO4)2] 2?, are also very similar to gallium phosphate chain in Cs4[(UO2)2(GaOH)2(PO4)4]?H2O.25 Titanium and germanium each form bonds with six oxygen atoms, four from phosphate anions and two from the neighbouring titanium or germanium unit, in octahedral environments. Due to the difference in the ionic size of Ti(IV) and Ge(IV), as well as the nature of the orbitals, participating in the bond formation, Ti-O and Ge-O distances are different. For Ti-O, the distances range from 1.926 ? to 1.960 ?, which is within the normal distance range for Ti in an octahedral environment. In the different Ge octahedra, Ge-O bonds vary from 1.77 ? to 2.06 ?, and again bond lengths in UGeP fit in this range. These distances are also much longer than Ge-O distances in the tetrahedra, where they can range between 1.70 ? and 1.79 ?. The bond-valence sums are consistent with Ti and Ge present in a +4 oxidation state. Selected bond distances and angles for UTiP and UGeP are summarized in Tables 5.4 and 5.5. The intersection of the uranyl phosphate chains, extended in the c direction, and titanium or germanium phosphate chains, run along the b axis, giving rise to the tunnels filled with Cs+ cations. (Figure 5.3) There are four strongly disordered atomic sites in the channels. According to the refinement of occupancy of these atomic sites, we suggest that 134 Table 5.4. Selected Bond Distances (?) and Angles (?) for Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O (UTiP) Distances (?) U(1)-O(6) 1.773(6) Ti(1)-O(2) 1.960(6) U(1)-O(4) 2.245(9) P(1)-O(5) 1.490(9) U(1)-O(5) 2.265(8) P(1)-O(4) 1.508(9) U(1)-O(3) 2.371(8) P(1)-O(7) 1.545(6) U(1)-O(1) 2.389(8) P(2)-O(1) 1.511(6) U(1)-O(3?) 2.507(8) P(2)-O(2) 1.528(8) Ti(1)-O(8) 1.926(3) P(2)-O(3) 1.555(8) Ti(1)-O(7) 1.933(6) Angles (?) O(6)?U(1)?O(6?) 178.7(4) O(7)?Ti(1)?O(7?) 180.00(15) O(8)?Ti(1)?O(8?) 180.0(5) O(2)?Ti(1)?O(2?) 180.000(2) 135 Table 5.5. Selected Bond Distances (?) and Angles (?) for Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O (UGeP) Distances (?) U(1)-O(6) 1.765(8) Ge(1)-O(2) 1.853(7) U(1)-O(4) 2.227(11) P(1)-O(5) 1.497(10) U(1)-O(5) 2.271(10) P(1)-O(4) 1.505(12) U(1)-O(3) 2.376(10) P(1)-O(7) 1.531(8) U(1)-O(1) 2.414(10) P(2)-O(2) 1.512(8) U(1)-O(3?) 2.521(10) P(2)-O(1) 1.504(10) Ge(1)-O(8) 1.875(4) P(2)-O(3) 1.538(11) Ge(1)-O(7) 1.842(8) Angles (?) O(6)?U(1)?O(6?) 177.1(7) O(7)?Ge(1)?O(7?) 180.0(3) O(8)?Ge(1)?O(8?) 180.000(1) O(2)?Ge(1)?O(2?) 180.000(2) 136 Figure 5.3. The framework of UTiP and UGeP. The channels along the c axis are filled with Cs+ cations (purple) and water molecules (red). Violet octahedra represent Ti or Ge, yellow ? phosphate, and green ? uranium centers. 137 three partially occupied Cs+ cations reside in the channels with Cs-O distances ranging from 3.029(8) ? to 3.6172(17) ? for UTiP and from 2.856(15) ? to 3.631(10) ? for UGeP. The fourth electron density peak is assigned to the partially occupied water molecule. It is 0.85 water molecules per unit cell for UTiP and 0.63 for UGeP. Due to the incompletely filled Cs positions, the overall structure has a positive charge deficiency. This can be balanced by protonation of the titanium phosphate or germanium phosphate backbone. The bond-valence sum for O(8) corresponds to 1.38 and 1.42 for UTiP and UGeP, respectively, which is consistent with the degree of protonation in these structures. This type of protonation is also found in aluminophosphate mineral tancoite, LiNa2H[Al(PO4)2(OH)],71,72 Cs4[(UO2)2(GaOH)2(PO4)4]?H2O25 and large number of aluminum, gallium, and transition metal phosphates.73-80 It is important to note that in Cs4[(UO2)2(GaOH)2(PO4)4]?H2O25 and LiNa2H[Al(PO4)2(OH)],71,72 protons are assigned to stabilize the structure with Ga and Al in oxidation state +3, but in the case of UTiP and UGeP, the proton?s role is restricted only to equilibrate the Cs+ cations deficiency. CONCLUSIONS In this chapter we have described two novel isostructural uranium-containing framework compounds incorporating Ti or Ge polyhedra. According to the structural characterization, we can imply that these compounds can have promising ion-exchange and conductivity properties due to the presence of highly disordered cationic sites. For further investigations, more detailed analysis of the Cs and water content, as well as knowledge about physical properties of structure are necessary. We are currently 138 investigating the methods for the synthesis of UTiP and UGeP with higher yields that will allow us to conduct detailed studies of the compounds. 139 REFERENCES 1. Chetham, A. K.; Ferey, G.; Loiseau, T. Angew. Chem., Int. Ed. Engl. 1999, 38, 3268. 2. Li, Y.; Cahill, C. L.; Burns, P. Chem. Mater. 2001, 13, 4026. 3. Halasyamani, P. S.; Walker, S. M.; O'Hare, D. J. Amer. Chem. Soc. 1999, 121, 7415. 4. Doran, M.; Walker, S. M.; O'Hare, D. Chem. Commun. (Cambridge) 2001, 19, 1988. 5. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2002, 163, 275. 6. Locock, A. J.; Burns, P. C J. Solid State Chem. 2002, 167, 226. 7. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2003, 175, 372. 8. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2003, 176, 18. 9. Burns, P. C.; Alexopoulos, C. M.; Hotchkiss, P. J.; Locock, A. J. Inorg. Chem. 2004, 43, 1816. 10. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2004, 177, 2675. 11. Doran, M.; Norquist, A. J.; O'Hare, D. Chem. Commun. (Cambridge) 2002, 24, 2946. 12. Li, Y.; Burns, P. C. J. Nucl. Materials 2001, 299, 219. 13. Wang, X.; Huang, J.; Liu, L.; Jacobson, A. J. J. Mater. Chem. 2002, 12, 406. 14. Wang, X.; Huang, J.; Jacobson, A. J. J. Amer. Chem. Soc. 2002, 124, 15190. 15. Huang, J.; Wang, X.; Jacobson, A. J. J. Mater. Chem. 2003, 13, 19115. 16. Chen, C.-S.; Kao, H.-M.; Lii, K.-H. Inorg. Chem. 2005, 44, 935. 17. Chen, C.-S.; Chiang, R.-K.; Kao, H.-M.; Lii, K.-H. Inorg. Chem. 2005, 44, 3914. 18. Obbade, S.; Dion, C.; Rivenet, M.; Saadi, M.; Abraham, F. J. Solid State Chem. 2004, 177, 2058. 140 19. Obbade, S.; Dion, C.; Saadi, M.; Yagoubi, S.; Abraham, F. J. Solid State Chem. 2004, 177, 3909. 20. Obbade, S.; Yagoubi, S.; Dion, C.; Saadi, M.; Abraham, F. J. Solid State Chem. 2004, 177, 1681. 21. Krivovichev, S. V.; Burns, P. C.; Armbruster, Th.; Nazarchuk, E. V.; Depmeier, W. Micropor. Mesopor. Mater. 2005, 78, 217. 22. Krivovichev, S. V.; Cahill, C. L.; Nazarchuk, E. V.; Burns, P. C.; Armbruster, Th.; Depmeier, W. Micropor. Mesopor. Mater. 2005, 78, 209. 23. Krivovichev, S. V.; Cahill, C. L.; Burns, P. C. Inorg. Chem.. 2003, 42, 2459. 24. Krivovichev, S. V.; Armbruster, Th.; Chernorukov, D. Yu.; Burns, P. C.; Nazarchuk, E. V.; Depmeier, W. Micropor. Mesopor. Mater. 2005, 78, 225. 25. Shvareva, T. Y.; Sullens, T. A.; Shehee, T. C.; Albrecht-Schmitt, T. E. Inorg. Chem. 2005, 44, 300. 26. Shvareva, T. Y.; Beitz, J. V.; Duin, E.C.; Albrecht-Schmitt, T. E. Chem. Mater. 2005, 17, 6219. 27. Shvareva, T. Y.; Albrecht-Schmitt, T. E. Inorg. Chem. 2005, 45, 1900. 28. Sykora, R. E.; Albrecht-Schmitt, T. E. Inorg. Chem. 2003, 42, 2179. 29. Burns, P. C.; Miller, M. L.; Ewing, R. C. Can. Mineral. 1996, 34, 845. 30. Burns, P. C. Can. Mineral. 2005, 43, 1839. 31. Kuznicki, S. M.; Thrush, K. A.; Allen, F. M.; Levine, S. M.; Hamil, M. M.; Hayhurst, D. T.; Mansour, M. Synth. Microporous Mater. 1992, 1, 427. 32. Clearfield, A. Solid State Sci. 2001, 3, 103. 33. Ekambaram, S.; Sevov, S. C. Angew. Chem., Int. Ed. Engl. 1999, 38, 372. 141 34. Ekambaram, S.; Serre, C.; Ferey, G.; Sevov, S. C. Chem. Mater. 2000, 12, 444. 35. Chippindale, A. M.; Grimshaw, M. R.; Powell, A. V.; Cowley, A. R. Inorg. Chem. 2005, 44, 4121. 36. Dickens, P. G.; Flynn,G. J.; Patat, S.; Stuttart, G. P. J. Mater. Chem. 1997, 7, 537 37. Burns, P. C. Rev. Mineral. 1999, 38, 23. 38. Ringwood, A. E.; Kesson, S. E.; Ware, N. G.; Hibberson, W. ; Major, A. Nature (London), 1979, 278, 219. 39. Szymanski, J. T.; Scott, J. D. Can. Mineral. 1982, 20, 271. 40. Vance, E. R.; Watson, J.N.; Carter, M.L.; Day, R.A., Begg, B.D.; J. Am. Ceram. Soc. 2001, 84, 141. 41. James, M.; Watson, J. N. J. Solid State Chem. 2002, 165, 261. 42. James, M.; Carter, M. L.; Watson, J. N J. Solid State Chem. 2003, 174, 329. 43. Wells, A. F. Structural Inorganic Chemistry, 5th ed., Oxford, 1984, p 1173. 44. Li, H.; Yaghi, O. M. J. Am. Chem. Soc. 1998, 120, 10569. 45. Conradsson, T. M.; Dadachov, S.; Zou, X. Micropor. Mesopor. Mater. 2000, 41, 183. 46. Cascales, C.; Gutierez-Puebla, E.; Iglesia, M.; Monge, M.A.; Ruiz-Valero, C.; Snejko, N. Chem. Commun. (Cambridge) 2000, 56, 1092. 47. Cascales, C.; Gomez-Lor, B.; Gutierez-Puebla, E.; Iglesia, M.; Monge, M. A.; Ruiz- Valero, C.; Snejko, N. Chem. Mater. 2002, 14, 677. 48. Cascales, C.; Gutierez-Puebla, E.; Monge, M.A.; Ruiz-Valero, C. Angew. Chem. ., Int. Ed. Engl. 1998, 37, 129. 49. Cascales, C.; Gutierez-Puebla, E.; Iglesia, M.; Monge, M. A.; Ruiz-Valero, C. Angew. Chem., Int. Ed. Engl. 1999, 38, 2436. 142 50. Feng, P. Y.; Bu, X. H.; Stucky, G. D. Chem. Mater. 1999, 11, 3025. 51. Li, H. L.; Eddaoudi, M.; Yaghi, O. M. Angew. Chem., Int. Ed. Engl. 1999, 38, 653. 52. Pl?vert, J.; Gentz, T. M.; Laine, A. ; Li, H. L.; Young, V. G.; Yaghi, O. M.; O?Keefe, M. J. Am. Chem. Soc. 2001, 123, 12706. 53. Pl?vert, J.; Gentz, T. M.; Groy, T. L.; O?Keefe, M.; Yaghi, O. M. Chem. Mater. 2003, 15, 714. 54. Zhou, Y.; Zhu, H.; Chen, Z.; Chen, M.; Xu, Y.; Zhang, H.; Zhao, D. Angew. Chem., Int. Ed. Engl. 2001, 40, 2166. 55. Xu, Y.; Ogura, M.; Okubo, T. Micropor. Mesopor. Mater. 2004, 70, 1. 56. Snejko, N.; Medina, M. E.; Gutierrez-Puebla, E.; Monge, M. A. Inorg. Chem. 2006, 45, 1591. 57. Lin, Z.-E.; Zhang, J.; Zheng, S.-T.; Yang, G.-Y. Micropor. Mesopor. Mater. 2004, 74, 205. 58. Medina, M.E.; Iglesias, M.; Snejko, N.; Gutierrez-Puebla, E.; Monge, M. A. Chem. Mater. 2004, 16, 594. 59. Bu, X.; Feng, P; Gier, T. E.; Zhao, D.; Stucky, G. D. J. Am. Chem. Soc. 1998, 120, 13389. 60. Lin, Z.-E.; Zhang, J.; Yang, G.-Y. Inorg. Chem. 2003, 42, 1797. 61. Zhang, H.-X.; Zhang, J.; Zheng, S.-T.; Yang, G.-Y. Inorg. Chem. 2003, 42, 6595. 62. Lin, Z.-E.; Zhang, J.; Zhao, J.-T.; Zheng, S.-T.; Pan, C.-Y.; Wang, G.-M.; Yang, G.- Y. Angew. Chem., Int. Ed. Engl. 2005, 44, 6884. 63. Legros, J. P.; Jeannin, Y. Acta Crystallogr. 1975, B31, 1133. 64. Legros, J. P.; Jeannin, Y. Acta Crystallogr. 1975, B31, 1140. 143 65. Sheldrick, G. M. SHELXTL PC, Version 6.12, An Integrated System for Solving, Refining, and Displaying Crystal Structures from Diffraction Data; Siemens Analytical X-Ray Instruments, Inc.: Madison, WI 2001. 66. Sheldrick, G. M. SADABS 2001, Program for absorption correction using SMART CCD based on the method of Blessing: Blessing, R. H. Acta Cryst. 1995, A51, 33. 67. Grohol, D.; Nocera, D. G. J. Am. Chem. Soc. 2002, 124, 2641 68. Almond, P. M.; Peper, S. M.; Bakker, E.; Albrecht-Schmitt, T. E. J. Solid State Chem. 2002, 168, 358. 69. Ramik, R. A.; Sturman, B. D.; Dunn, P. J.; Povarennykh, A. S. Can. Mineral. 1980, 18, 185. 70. Hawthorne, F. C. Tsch. Mineral. Petro. Mitt. 1983, 31, 121. 71. Lin, H.-M.; Lii, K.-H. Inorg. Chem. 1998, 37, 4220. 72. Attfield, M. P.; Morris, R. E.; Burshtein, I.; Campana, C. F.; Cheetham, A. K. J. Solid State Chem. 1995, 118, 412. 73. Cavellec, M.; Riou, D.; Greneche, J.-M.; F?rey, G. Inorg. Chem. 1997, 36, 2187. 74. Lethbridge, Z. A. D.; Lightfoot, P.; Morris, R. E.; Wragg, D. S.; Wright, P. A.; Kvick, A.; Vaughan, G. J. Solid State Chem. 1999, 142, 455. 75. Walton, R. I.; Millange, F.; O?Hare D.; Paulet, C.; Loiseau, T.; F?rey, G. Chem. Mater. 2000, 12, 1977. 76. Mahesh, S.; Green, M. A.; Natarajan, S. J. Solid State Chem. 2002, 165, 334. 77. Mandal, S.; Green, M. A.; Natarajan, S. J. Solid State Chem. 2004, 177, 1117. 78. Harvey, H. G.; Teat, S. J.; Tang, C. C.; Cranswick, L. M.; Attfield, M. P. Inorg. Chem. 2003, 42, 2428. 144 CHAPTER 6 GENERAL ROUTE TO THREE-DIMENSIONAL FRAMEWORK URANYL TRANSITION METAL PHOSPHATES WITH ATYPICAL STRUCTURAL MOTIFS: THE CASE EXAMPLES OF Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4}, Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} AND Cs3+x[(UO2)3CuH4-x(PO4)5]?H2O ABSTRACT The reaction of UO2(NO3)2?6H2O with Co, Ni, or Cu metal, phosphoric acid, and CsCl under mild hydrothermal conditions results in the formation of Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4} (UCoP), Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} (UNiP) or Cs3+x[(UO2)3-xCuH4-x(PO4)5]?H2O (UCuP). UCoP and UNiP are isostructural compounds. Their structures contain uranium atoms in pentagonal bipyramidal and hexagonal bipyramidal environments. The interaction of the uranyl cations and phosphate anions creates layers in the [ab] plane. The uranyl phosphate layers are joined together by octahedral Co or Ni centers wherein the Co and Ni are bound by phosphate and two cis water molecules. In addition, the Co and Ni ions are also ligated by a uranyl oxo atom. The presence of these octahedral building units stitches the structure together into a three-dimensional framework where void spaces are filled by Cs+ cations. The structure of UCuP contains uranium centers in UO6 tetragonal bipyramidal and UO7 pentagonal bipyramidal geometries. The uranyl moieties are bridged by 145 phosphate anions into sinusoidal sheets that extend into the [bc] plane and are linked into a three-dimensional structure by CuII. The Cu centers reside in square planar environments. Charge balance is maintained by Cs+ cations. Both the structure types and the uranyl phosphate layers in UCoP, UNiP, and UCuP are novel. INTRODUCTION The structural chemistry of UVI is dominated by the formation of layered compounds containing uranyl, UO22+, cations wherein the uranyl units are aligned in a roughly parallel fashion.1,2 These layered compounds have been shown to display several important and potentially useful properties including fast proton conductivity and ion exchange that is particularly well-known for uranyl phosphates.3-8 More recently, attention has been focused on preparing uranyl phases that adopt three-dimensional networks that may find applications in the storage of key radionuclides from spent nuclear fuel and as new selective ion-exchange materials.9 One of the most appealing features of constructing new solids from uranyl-containing polyhedra is that the uranium centers in these compounds can adopt coordination environments seldom seen in transition metal chemistry such as UO7 pentagonal bipyramids and UO8 hexagonal bipyramids.1,2 In previous chapters we have reported the syntheses and structures of the first uranyl gallium,9 vanadium,10 titanium and germanium phosphates with open-framework structures. As we demonstrated, UGaP1, UGaP2, and KUVIVP (RbUVIVP) have ion- exchange properties. Herein we have devoted our research to preparing additional first row transition metal uranyl phosphates where the gallium, germanium, vanadium, or 146 titanium centers have been replaced by other transition metals, which may, in turn, lead to new architectures, and, therefore, to properties, not previously found. We also report a relatively general and quite facile route to preparing first row transition metal uranyl phosphates with three-dimensional structures with varying degrees of openness. We illustrate this with the examples of the CoII, NiII, and CuII compounds, Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4} (UCoP), Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} (UNiP) or Cs3+x[(UO2)3-CuH4-x(PO4)5]?H2O (UCuP), all three of which display new methods of constructing three-dimensional uranyl-containing frameworks.11-22 EXPERIMENTAL Syntheses. UO2(NO3)2?6H2O (98%, Alfa-Aesar), Co powder (99.99%, Sigma Aldrich), Ni powder (99.99%, Sigma Aldrich), Cu powder (99.99%, Sigma Aldrich), H3PO4 (98%, Aldrich), and CsCl (99.9%, Cerac) were used as received. Co, Ni, and Cu powders were weighed separately under the inert atmosphere due to their potential rapid oxidation in the air. The PTFE liners were used for all reactions. While the UO3 contains depleted U, standard precautions for handling radioactive materials should be followed. Old sources of depleted U should not be used, as the daughter elements of natural decay are highly radioactive and present serious health risks. Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4} (UCoP). UO2(NO3)2?6H2O (0.285 g, 0.568 mmol), Co powder (0.067 g, 1.1 mmol), H3PO4 (0.167 g, 1.72 mmol), CsCl (0.479 g, 2.85 mmol), and 1.5 mL of Millipore-filtered water were loaded into a 23-mL autoclave. The reaction was run at 195 ?C in a box furnace for 7 days and then cooled at a rate of 9 ?C/h to room temperature. The product mixture contained bright-orange crystals 147 of UCoP and water-soluble blue crystals of cesium tetrachlorocobalt(II). The desired product was isolated by washing the mixture with water and methanol. Yield: 0.286 mg (100% based on U). Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} (UNiP). UO2(NO3)2?6H2O (0.286 g, 0.57 mmol), Ni powder (0.067 g, 1.1 mmol), H3PO4 (0.167 g, 1.72 mmol) and CsCl (0.479 g, 2.85 mmol), were mixed together in a 23-mL autoclave and followed by addition of 1.5 mL of Millipore-filtered water. The reaction was run at 195 ?C in a box furnace for 5 days and then slowly cooled to room temperature with average rate 3?C/hour. The product mixture contained precipitate of greenish-yellow plates of UNiP and bright green mother liquid. UNiP precipitate was washed with water, methanol and dried by air. Yield: 285 mg (82.4 % based on U). Cs3+x[(UO2)3CuH4-x(PO4)5]?H2O (UCuP). UO2(NO3)2?6H2O (0.285 g, 0.567 mmol), Cu powder (0.072 g, 1.1 mmol), H3PO4 (0.167 g, 1.72 mmol) and CsCl (0.477 g, 2.85 mmol), were mixed together in a 23-mL autoclave followed by addition of 1.5 mL of Millipore-filtered water. The reaction was run at 220 ?C in a box furnace for 7 days and then cooled at a rate of 9 ?C/h to room temperature. Due to the large size of the crystals, green UCuP product was easily separated from the product mixture, that also included also crystals of cesium tetrachlorocopper and cesium uranyl pyrophosphate. Yield: 330 mg (97.6 % based on U). X-ray structural analysis. Crystals of UCoP, UNiP, and UCuP were mounted on glass fibers and aligned on a Bruker SMART APEX CCD X-ray diffractometer. Intensity measurements were performed using graphite monochromated Mo Ka radiation from a sealed tube and monocapillary collimator. SMART (v 5.624) was used for 148 preliminary determination of the cell constants and data collection control. The intensities of reflections of a sphere were collected by a combination of 3 sets of exposures (frames). Each set had a different f angle for the crystal and each exposure covered a range of 0.3? in w. A total of 1800 frames were collected with an exposure time per frame of 30 s for the crystals of UCoP, UNiP, and UCuP. Determination of integrated intensities and global refinement were performed with the Bruker SAINT (v 6.02) software package using a narrow-frame integration algorithm. A face-indexed analytical absorption correction was initially applied using XPREP, where individual shells of unmerged data were corrected analytically.23 These files were subsequently treated with a semiempirical absorption correction by SADABS.24 The program suite SHELXTL (v 6.12) was used for space group determination (XPREP), direct methods structure solution (XS), and least-squares refinement (XL).23 The final refinements included anisotropic displacement parameters for all atoms. Secondary extinction was not noted for either crystal. Crystallographic details are given in Table 6.1. Atomic coordinates, equivalent isotropic displacement parameters, symmetry and occupancy for UCoP, UNiP, and UCuP are given in Tables 6.2, 6.3, and 6.4 respectively. Thermal analysis. The data were collected using differential scanning calorimeter TA Instrument, model 2920. The samples of known mass encapsulated in the aluminum container, was heated under nitrogen atmosphere from 30 ?C to 600 ?C with heating rate of 10?C /minute. Table 6.1. Crystallographic Data for Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4} (UCoP), Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} (UNiP) and Cs3+x[(UO2)3CuH4-x(PO4)5]?H2O 149 (UCuP). Compound UCoP UNiP UCuP Formula Mass 2003.65 2003.21 1788.35 Color and habit orange plate yellow-green green plate Crystal System monoclinic monoclinic orthorhombic Space group C2/c (No.15 ) C2/c (No.15 ) Pbcm (No. 57) a (?) 18.0551(8) 18.0443(13) 7.5867(4) b (?) 10.7478(5) 10.7375(8) 19.9574(10) c (?) 15.3504(7) 15.2268(11) 17.9726(9)  ?, C 99.2420(10) 99.1520(10) 90 V (?3) 2940.1(2) 2912.6(4) 2721.2(2) Z 4 4 4 T (K) 193 193 193 l (?) 0.71073 0.71073 0.71073 Maximum 2q(deg.) 56.56 56.58 56.58 rcalcd (g cm?3) 4.524 4.595 4.348 m(Mo Ka) (cm?1) 258.86 262.82 231.04 R(F) for Fo2 > s(Fo2) 0.0281 0.0319 0.0354 Rw(Fo2) b 0.0692 0.0709 0.0835 a ( )R F F F F= -  o c o . b ( ) ( )R F w F F wF w o 2 o 2 c 2 2 o 4 1 2 = -    . Table 6.2. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for 150 Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4} (UCoP). Atom Symmetrya x y z Ueq (?2) b Occup. U(1) 4 e 2 0 -0.28631(3) ? 0.00974(9) 1 U(2) 4 e 2 0 0.09385(3) ? 0.00892(9) 1 U(3) 8 f 1 0.20409(1) 0.28203(2) 0.29034(2) 0.00778(7) 1 Co(1) 8 f 1 0.21057(5) 0.33986(9) 0.05694(6) 0.01051(19) 1 P(1) 8 f 1 0.0001(2) 0.3800(3) 0.2841(3) 0.0104(7) 0.50 P(2) 8 f 1 0.34785(10) 0.44837(16) 0.20623(12) 0.0085(3) 1 P(3) 8 f 1 0.31032(10) 0.10909(16) 0.15358(12) 0.0079(3) 1 O(1) 4 e 2 0 0.4966(8) ? 0.069(4) 1 O(2) 8 f 1 -0.0192(6) 0.4143(9) 0.3795(7) 0.012(2) 0.50 O(3) 8 f 1 0.0647(3) 0.2934(5) 0.2844(6) 0.045(2) 1 O(4) 8 f 1 0.4309(3) 0.4041(4) 0.2155(3) 0.0102(10) 1 O(5) 8 f 1 0.3640(3) 0.5903(4) 0.2212(3) 0.0096(10) 1 O(6) 8 f 1 0.3115(3) 0.3939(4) 0.2801(3) 0.0091(10) 1 O(7) 8 f 1 0.3057(3) 0.4264(5) 0.1153(3) 0.0125(10) 1 O(8) 8 f 1 0.2775(3) 0.1403(5) 0.2362(3) 0.0113(10) 1 O(9) 8 f 1 0.3140(3) -0.0325(4) 0.1410(3) 0.0111(10) 1 O(10) 8 f 1 0.2604(3) 0.1628(5) 0.0701(3) 0.0102(10) 1 O(11) 8 f 1 0.3897(3) 0.1638(5) 0.1593(3) 0.0118(10) 1 O(12) 8 f 1 0.0407(3) -0.2874(6) 0.1515(4) 0.0260(14) 1 O(13) 8 f 1 0.1595(3) 0.5213(5) 0.0361(4) 0.0232(13) 1 O(14) 8 f 1 0.1051(3) 0.2584(5) 0.0071(3) 0.0164(11) 1 O(15) 8 f 1 0.2387(3) 0.2303(5) 0.3998(3) 0.0129(10) 1 O(16) 8 f 1 0.1722(3) 0.3329(5) 0.1793(3) 0.0141(10) 1 O(17) 8 f 1 0.0092(3) 0.0926(5) 0.1371(4) 0.0160(11) 1 151 Cs(1) 8 f 1 0.37456(3) 0.44843(6) 0.47121(3) 0.02451(13) 1 a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 152 Table 6.3. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} (UNiP). Atom Symmetrya x y z Ueq (?2) b occup. U(1) 4 e 2 0 -0.28780(4) ? 0.01120(11) 1 U(2) 4 e 2 0 0.09224(4) ? 0.00986(11) 1 U(3) 8 f 1 0.20419(2) 0.28060(3) 0.29040(2) 0.00887(9) 1 Ni(1) 8 f 1 0.21041(6) 0.33766(10) 0.05657(7) 0.0109(2) 1 P(1) 8 f 1 0.0001(2) 0.3792(4) 0.2849(3) 0.0130(9) 0.50 P(2) 8 f 1 0.34763(12) 0.4467(2) 0.20535(15) 0.0094(4) 1 P(3) 8 f 1 0.31001(12) 0.1083(2) 0.15224(15) 0.0093(4) 1 O(1) 4 e 2 0 0.4951(11) ? 0.062(4) 1 O(2) 8 f 1 -0.0180(7) 0.4136(10) 0.3857(11) 0.026(3) 0.50 O(3) 8 f 1 0.0647(4) 0.2910(6) 0.2841(7) 0.044(3) 1 O(4) 8 f 1 0.4311(3) 0.4027(5) 0.2147(4) 0.0115(12) 1 O(5) 8 f 1 0.3642(3) 0.5885(5) 0.2204(4) 0.0123(13) 1 O(6) 8 f 1 0.3121(3) 0.3919(5) 0.2799(4) 0.0113(12) 1 O(7) 8 f 1 0.3053(3) 0.4240(5) 0.1136(4) 0.0139(13) 1 O(8) 8 f 1 0.2771(3) 0.1380(5) 0.2358(4) 0.0133(13) 1 O(9) 8 f 1 0.3143(3) -0.0340(5) 0.1390(4) 0.0112(12) 1 O(10) 8 f 1 0.2600(3) 0.1629(5) 0.0691(4) 0.0104(12) 1 O(11) 8 f 1 0.3894(3) 0.1632(5) 0.1582(4) 0.0119(12) 1 O(12) 8 f 1 0.0405(4) -0.2884(7) 0.1508(4) 0.0247(16) 1 O(13) 8 f 1 0.1591(3) 0.5150(6) 0.0383(4) 0.0185(14) 1 O(14) 8 f 1 0.1075(3) 0.2553(6) 0.0084(4) 0.0159(14) 1 O(15) 8 f 1 0.2389(3) 0.2285(5) 0.4004(4) 0.0155(13) 1 O(16) 8 f 1 0.1732(3) 0.3316(6) 0.1779(4) 0.0151(13) 1 153 O(17) 8 f 1 0.0102(3) 0.0921(6) 0.1360(4) 0.0165(14) 1 Cs(1) 8 f 1 0.37476(3) 0.44631(6) 0.47172(4) 0.02446(15) 1 a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 154 Table 6.4. Atomic Coordinates and Equivalent Isotropic Displacement Parameters for Cs3+x[(UO2)3CuH4x(PO4)5]?H2O (UCuP). Atom Symmetrya x y z Ueq (?2) b Occup. U(1) 4 d ..m -0.50309(5) 0.39763(2) ? 0.01050(10) 1 U(2) 8 e 1 0.21145(4) 0.65898(1) 0.55377(2) 0.00997(9) 1 Cu(1) 4 d ..m 0.00099(18) 0.40714(6) ? 0.0115(3) 1 P(1) 4 d ..m 0.1054(4) 0.66578(14) ? 0.0140(6) 1 P(2) 8 e 1 0.2132(3) 0.48119(10) 0.62264(11) 0.0132(4) 1 P(3) 8 e 1 0.2080(3) 0.66171(10) 0.38077(11) 0.0121(4) 1 O(1) 4 d ..m -0.0435(13) 0.7196(4) ? 0.038(3) 1 O(2) 8 e 1 0.2183(8) 0.6708(3) 0.6816(3) 0.0227(14) 1 O(3) 4 d ..m -0.0119(10) 0.6016(4) ? 0.0195(18) 1 O(4) 8 e 1 0.3853(8) 0.4639(3) 0.6600(4) 0.0293(16) 1 O(5) 8 e 1 0.1881(15) 0.4311(4) 0.5578(4) 0.069(3) 1 O(6) 8 e 1 0.0592(9) 0.4756(3) 0.6755(4) 0.0366(18) 1 O(7) 8 e 1 0.2257(9) 0.5501(3) 0.5885(4) 0.0258(14) 1 O(8) 8 e 1 0.1900(7) 0.7231(3) 0.4337(3) 0.0142(12) 1 O(9) 8 e 1 0.3862(7) 0.6660(3) 0.3397(3) 0.0159(12) 1 O(10) 8 e 1 0.0542(8) 0.6601(3) 0.3261(3) 0.0182(12) 1 O(11) 8 e 1 0.2038(8) 0.6030(3) 0.4357(3) 0.0173(12) 1 O(12) 8 e 1 -0.0215(8) 0.6564(3) 0.5581(3) 0.0163(12) 1 O(13) 8 e 1 0.4452(8) 0.6641(3) 0.5502(3) 0.0183(12) 1 O(14) 4 d ..m 0.3061(10) 0.4482(4) ? 0.0178(17) 1 O(15) 4 d ..m -0.6992(11) 0.3485(4) ? 0.0170(17) 1 O(16) 4 d ..m -0.394(2) 0.6575(10) ? 0.132(9) 1 Cs(1) 8 e 1 -0.28867(8) 0.54778(3) 0.60995(3) 0.02581(15) 1 155 Cs(2) 8 e 1 -0.2761(1) 0.77973(5) 0.57516(6) 0.0260(4) 0.569(3) a Symmetry describes multiplicity, Wyckoff parameter and symmetry of the atomic site. b Ueq is defined as one-third of the trace of the orthogonalized Uij tensor. 156 UV-vis-NIR Diffuse Reflectance Spectra. The diffuse reflectance spectra of UCoP, UNiP, and UCuP were measured from 1800 to 200 nm on a Shimadzu UV3100 spectrophotometer equipped with an integrating sphere attachment with BaSO4 as the standard. Kubelka-Monk function was used in order to transfer reflectance data to arbitrary absorbance units. Magnetic properties. Magnetic susceptibility measurements were conducted using a Quantum Design superconducting interference device (SQUID), under applied fields ranging from 100 to 10000 gauss as a function of temperature over the range of 5 ? 320 K and magnetization data were collected at 5 K in fields ranging from 0 ? 10,000 gauss. Samples were doubly encapsulated in aluminum containers, which contributed significantly to the measured signal, especially at higher temperatures and for samples with a low inherent susceptibility. Empty containers were run independently under the same conditions and the data used directly for background subtraction. Diamagnetic contributions to the data were also calculated and used to correct the measured susceptibility. The temperature dependent magnetic susceptibility data were analyzed assuming Curie-Weiss behavior with an added temperature independent term, c = C / (T-q) + cTIP, where C and q are Curie and Weiss constants and cTIP is the temperature-independent contribution to the susceptibility. C is related to the effective magnetic moment 2/1 2 0 3  = B eff N kC mm in which N0 is Avogadro?s number and cB is the Bohr magneton (0.927 x 10-20 erg/Gauss). For cases in which cTIP is small, the effective magnetic moment can 157 be obtained directly from the slope of the temperature dependence versus inverse magnetic susceptibility. Ion-exchange Studies. NaCl, KCl, CsCl, CaCl2, and SrCl2 solutions were used in order to estimate ion-exchange capacities of UCoP and UCuP. Concentrations of the solutions of different cations were measured with classic potentiometric methods with ion-selective electrodes for K+, Na+, and Cs+ uptake,25 using an Accumet A15 pH meter, and atomic adsorption spectroscopy, using a Varian 240 AA spectrometer with hollow cathode Ca and Sr lamps for Ca2+ and Sr2+ uptake. K+-, Na+- and Cs+-selective membranes were made in according with Fluka recommended compositions. Silver/silver chloride electrode was used as a reference.25 In order to prevent contamination of tested solutions with KCl the reference electrode with double junction and lithium acetate as a bridging electrolyte were utilized. In all cases samples were equilibrated in the 5 ml capped vials with different cation solutions of known concentration for 48 hours at constant shaking, and then the change in the concentrations were determined. The approximate volume to mass ratio 200:1 (4 ml of the solution for 0.02 g of sample) was used in all experiments. Additionally the kinetic of Sr2+ uptake was studied using a 90Sr source. 10 mL of 90Sr2+ with initial activity 20 mCu/milliliter was added to the Sr2+ solution. One milliliter aliquots were taken at different time intervals, mixed with 19 mL of scintillation cocktail, placed in a scintillation counter and counted for 200 sec. 158 RESULTS AND DISCUSSION Syntheses. In order to investigate new pathways to the formation of uranyl- containing framework materials we have studied the interaction of pure transition metals with an acidic reaction mixture that contains uranyl nitrate, phosphoric acid, and alkali metal chloride under hydrothermal conditions. As Grohol and Nocera discussed for vanadium jarosites formation,26 slow oxidation of metallic vanadium, due to the passivation of the vanadium surface, can be the rate-limiting step in the formation of the product. We suggest that rate-limiting oxidation and therefore the dissolving step provides slow introduction of the transition metal to the reaction and enforces the formation of the multicomponent system instead of a mixture of separate metal and uranyl phosphates. But due to passivation of the metal surface the reaction is not completed and traces of metal particles can be found in the product mixture. In the same study the authors have searched for potential oxidants for the vanadium metal. They proposed that oxygen, dissolved in water, or protons could be simultaneously reduced, and therefore could serve as oxidants. But reaction under a nitrogen atmosphere yielded the same product as under oxygen atmosphere. That result revealed the oxidation nature of the protons.26 The oxidation potential of the Co0/Co2+, Ni0/Ni2+, and Cu0/Cu2+ are +0.28 V, +0.25 V and +0.15 V respectively and therefore metals can extrude the hydrogen from the phosphoric acid. Hence we can summarize that protons act as oxidizing agents in our uranyl-transition metal-phosphate formation reaction. This statement has three important practical consequences. At first, we can conclude that the formation of all uranyl-transition metal-phosphates UCoP, UNiP, UCuP, UTiP, UVVP and KUVIVP (RbUVIVP) could proceed only in acidic conditions 159 and an increase in pH usually led to the lower product content. In addition, according to suggested rate-limiting oxidation of transition metals, we could obtain near quantitative yield of the desired products by increasing the temperature and time of the reaction. Finally these general rules can be applied to most members of first-row transition metals, enabling uranyl-first row transition metal phosphate framework products with different structures and therefore with different properties to be produced in high yield. Structure. Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4} (UCoP) and Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} (UNiP). The structure of UCoP and UNiP contain three crystallographically unique uranium atoms, two of which are in pentagonal bipyramidal environments and one of which is in a hexagonal bipyramidal geometry. The fundamental building unit of UCoP and UNiP is shown in Figure 6.1. In each case, there is a central uranyl core designated by two short U=O distances. The uranyl bond distances are within expected limits and range from 1.767(5) to 1.800(6) ?. The uranyl units form longer interactions with the oxygen atoms from the phosphate anions. These distances range from 2.263(5) ? to 2.507(6) ? for UCoP and from 2.308(6) ? to 2.505(6) ? for UNiP. The interaction of the uranyl cations and phosphate anions creates layers in the [ab] plane. These layers are a new topology for a uranium oxide.1,2 (Figure 6.2) There is one crystallographically unique disordered phosphate anion present in the layer. Other than the disordered unit, the phosphate units are approximately tetrahedral with normal P-O bond distances. The uranyl phosphate layers are joined together by octahedral Co or Ni centers wherein the Co (Ni) is bound by phosphate and two cis water molecules. In addition, the Co or Ni ions are also ligated by a uranyl oxo atom from the 160 Figure 6.1. The fundamental building unit of UCoP includes three crystallographically unique U centers, two in pentagonal bipyramidal and one in hexagonal bipyramidal environment, three tetrahedral P centers and Co octahedra. 50% probability thermal ellipsoids are represented. UNiP is isostructural with UCoP. 161 Figure 6.2. View down the c axis of the uranyl phosphate layers in UCoP and UNiP. One of the phosphate anions is disordered and is shown in ball-and-stick format. Uranium polyhedra are in green and phosphate are in yellow. 162 U(3) uranyl unit. This Co-O bond distance is 2.105(5) ? and Ni-O distance is 2.065(6) ?, and it fits well within the other Co-O or Ni-O bond distances, which range from 2.031(5) ? to 2.159 (6) ? in UCoP and from 2.018(6) ? to 2.116(6) ? in UNiP. Bond distances and angles are summarized in Tables 6.5 and 6.6. This kind of interaction is rare,27-32 but is known from compounds such as [Cu(H2O)4](UO2HGeO4)2?2H2O.32 The Co and Ni centers form dimers via two 3-oxo atoms from two phosphate anions. The presence of these octahedral building units stitches the structure together into a three- dimensional framework where void spaces are filled by Cs+ cations. (Figure 6.3) There are channels 5.0 ? ? 5.0 ? in this structure that extend down the b axis. The method of construction of this mixed metal phosphate differs from that of UGaP1, UGaP2, UTiP, UGeP, UVVP, KUVIVP, and RbUVIVP. Bond-valence sum calculations provide values of 6.08 (6.10), 6.24 (6.22), and 6.07 (6.04) for U(1), U(2), and U(3), respectively, confirming the expected oxidation state of +6.33 More importantly, the sum for the Co and Ni center is 1.98 and 1.95 respectively, consistent with that of CoII and NiII.34,35 Totalling the charges of the metal centers and phosphate anions reveals that there must be one proton within the structure. This proton most likely resides on the terminal oxygen atom of the phosphate anion containing P(1). Unfortunately, the disorder of this anion does not allow for conclusive bond distance comparisons. Cs3+x[(UO2)3CuH4x(PO4)5]?H2O (UCuP). Similar to the structure of UCoP and UNiP, UCuP contains two-dimensional uranyl phosphate layers that are linked into a three-dimensional structure by a transition metal center. Beyond this general similarity, the structures of UCoP (UNiP) and UCuP are dramatically different for a number of reasons. One of these is that there are two crystallographically unique uranium centers in 163 Figure 6.3. An illustration of the three-dimensional structure of UCoP and UNiP as viewed down the b axis. Uranium polyhedra are in green, phosphate in yellow, cobalt or nickel in orange. Purple centers represent the Cs+ cations. 164 Table 6.5. Selected Bond Distances (?) and Angles (?) for Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4} (UCoP). Distances (?) U(1)-O(12) 1.785(6) Co(1)-O(10) 2.101(5) U(1)-O(11) 2.302(5) Co(1)-O(16) 2.105(5) U(1)-O(1) 2.333(9) Co(1)-O(14) 2.123(5) U(1)-O(4) 2.411(5) Co(1)-O(13) 2.159(6) U(2)-O(17) 1.767(5) P(1)-O(1) 1.358(9) U(2)-O(4) 2.405(5) P(1)-O(2) 1.602(11) U(2)-O(5) 2.425(5) P(1)-O(3) 1.491(7) U(2)-O(3) 2.459(6) P(2)-O(7) 1.497(5) U(3)-O(15) 1.786(5) P(2)-O(6) 1.516(5) U(3)-O(16) 1.794(5) P(2)-O(4) 1.557(5) U(3)-O(8) 2.263(5) P(2)-O(5) 1.563(5) U(3)-O(9) 2.302(5) P(3)-O(8) 1.521(5) U(3)-O(6) 2.308(5) P(3)-O(9) 1.536(5) U(3)-O(5) 2.391(5) P(3)-O(10) 1.555(5) U(3)-O(3) 2.507(6) P(3)-O(11) 1.539(5) Co(1)-O(7) 2.031(5) P(1)-P(1) 1.045(8) Co(1)-O(10) 2.099(5) Angles (?) O(12)?U(1)?O(12) 179.2(4) O(7)?Co(1)?O(14) 173.6(2) O(17)?U(2)?O(17) 179.2(3) O(10)?Co(1)?O(13) 177.0(2) O(15)?U(3)?O(16) 178.1(2) O(10)?Co(1)?O(16) 174.5(2) 165 Table 6.6. Selected Bond Distances (?) and Angles (?) for Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} (UNiP). Distances (?) U(1)-O(12) 1.780(7) Ni(1)-O(10) 2.067(6) U(1)-O(11) 2.308(6) Ni(1)-O(16) 2.065(6) U(1)-O(1) 2.331(11) Ni(1)-O(14) 2.083(6) U(1)-O(4) 2.410(5) Ni(1)-O(13) 2.116(6) U(2)-O(17) 1.774(6) P(1)-O(1) 1.352(11) U(2)-O(4) 2.401(6) P(1)-O(2) 1.661(15) U(2)-O(5) 2.420(6) P(1)-O(3) 1.504(8) U(2)-O(3) 2.449(6) P(2)-O(7) 1.501(6) U(3)-O(15) 1.783(6) P(2)-O(6) 1.509(6) U(3)-O(16) 1.800(6) P(2)-O(4) 1.562(6) U(3)-O(8) 2.261(6) P(2)-O(5) 1.562(6) U(3)-O(9) 2.312(6) P(3)-O(8) 1.521(6) U(3)-O(6) 2.311(6) P(3)-O(9) 1.544(6) U(3)-O(5) 2.396(5) P(3)-O(10) 1.548(6) U(3)-O(3) 2.505(6) P(3)-O(11) 1.538(6) Ni(1)-O(10) 2.074(6) P(1)-P(1) 1.061(9) Ni(1)-O(7) 2.018(6) Angles (?) O(12)?U(1)?O(12) 179.6(5) O(7)?Co(1)?O(14) 174.2(2) O(17)?U(2)?O(17) 179.9(4) O(10)?Co(1)?O(13) 177.8(2) O(15)?U(3)?O(16) 177.4(3) O(10)?Co(1)?O(16) 175.6(2) 166 UCuP that are found as UO6 tetragonal bipyramids and UO7 pentagonal bipyramids. (Figure 6.4) The uranyl U=O bond distances for these two polyhedra range from 1.770(6) ? to 1.803(8) ?. The equatorial U-O distances range from 2.236(5) ? to 2.514(5) ?. Selected bond distances and angles for UCuP are given in Table 6.7. Taking into account the differences in coordination number, these distances were used to calculate bond-valence sums of 5.96 and 6.08 for U(1) and U(2), respectively.33 The uranyl moieties are bridged by phosphate anions into sinusoidal sheets that extend into the [bc] plane. (Figure 6.5) Again, this is apparently a new layered topology for a uranium oxide compound.1,2 The uranyl phosphate layers in UCuP are interconnected by Cu centers that reside in square planar environments, being bound by four phosphate anions with two Cu-O distances of 1.962(6) ? and 1.964(6) ?. The bond-valence sum for the Cu site is 1.85 and is consistent with that of CuII.34,35 The overall structure is three-dimensional. However, the structure is much less open than that of UCoP and UNiP or other mixed-metal uranyl transition metal phosphates prepared thus far. Channels for the Cs+ cations and water molecules that are observed in other mixed-metal uranyl phosphate structures are not as distinct in UCuP, although there are found channels 3.2 ? ? 4.1 ? extending down the c axis. (Figure 6.6) Of interest, however, is that there are two Cs+ sites within the channels. One of these sites is fully occupied. The second site is occupied by 0.57. This has important consequences for the degree of protonation for UCuP. There are two phosphate anions with terminal oxo atoms in the structure. For the phosphate anion containing P(1), there 167 Figure 6.4. A view of the uranyl tetragonal and pentagonal bipyramids, square planar Cu and phosphate tetrahedra in UCuP with 50% probability ellipsoids. 168 Table 6.7. Selected Bond Distances (?) and Angles (?) for Cs3+x[(UO2)3CuH4x(PO4)5]?H2O (UCuP). Distances (?) U(1)-O(15) 1.782(8) Cu(1)-O(6) 1.964(6) U(1)-O(14) 1.803(8) P(1)-O(2) 1.501(6) U(1)-O(9) 2.236(5) P(1)-O(1) 1.560(9) U(1)-O(4) 2.254(6) P(1)-O(3) 1.560(8) U(2)-O(12) 1.770(6) P(2)-O(4) 1.509(6) U(2)-O(13) 1.777(6) P(2)-O(6) 1.509(7) U(2)-O(7) 2.263(6) P(2)-O(7) 1.509(7) U(2)-O(2) 2.310(6) P(2)-O(5) 1.548(7) U(2)-O(8) 2.369(5) P(3)-O(10) 1.525(6) U(2)-O(11) 2.399(5) P(3)-O(11) 1.531(6) U(2)-O(8) 2.514(5) P(3)-O(9) 1.543(6) Cu(1)-O(10) 1.962(6) P(3)-O(8) 1.557(6) Angles (?) O(15)?U(1)?O(14) 179.4(3) O(10)?Cu(1)?O(6) 178.7(3) O(9)?U(2)?O(4) 178.4(2) O(6)?Cu(1)?O(6) 86.0(5) O(10)?Cu(1)?O(10) 88.4(3) 169 Figure 6.5. Depiction of uranyl phosphate layers in UCuP. Uranyl polyhedra are in green and phosphates are in yellow. 170 Figure 6.6. View of the three-dimensional structure of UCuP wherein the CuII centers link uranyl phosphate layers together. Uranium polyhedra are in green, phosphate in yellow, and copper in blue. Some of the Cs+ sites have been omitted for clarity. 171 are two terminal oxo atoms with P-O bond distances of 1.560(9) ?, as compared to the bridging P-O bond distances of 1.501(6) ?. We propose that both of these aforementioned sites are protonated. A second phosphate anion containing P(2) has three bridging P-O bond distances of 1.509(7) ? and one terminal distance of 1.548(7) ?. This latter site is likely to be protonated also. However, if all three of these terminal sites are fully protonated, then there is an excess positive charge of 0.14. A closer look at the terminal oxygen atom around P(2) reveals an elongated thermal ellipsoid for this atom. We speculate that this elongation represents an average of protonated and deprotonated oxygen positions. Therefore, a potentially correct formula for UCuP can be written as Cs3.14[(UO2)3Cu(H2PO4)(H1.93PO4)2(PO4)2]?H2O. UV-vis-NIR diffuse reflectance study. Spectroscopic verification of the oxidation state for Co, Ni and Cu was provided by UV-vis-NIR diffuse reflectance spectroscopy. In the case of UCoP (Figure 6.8) the spectrum yields three main bands that are expected for octahedral CoII.36 4T2g---4T1g transition is located at 9700 cm-1 (1030 nm). The band at 19100 cm-1 (525 nm) matches with 4T1g(P)---4T1g, that is usually seen in the visible region. Additionally, a very weak band located around 13000 cm-1 (770 nm), corresponds to an 4A2g transition, which usually appears as a shoulder. The spectrum of UNiP (Figure 6.9) reveals three main bands at 10100 cm-1 (990 nm), 16700 cm-1 (600 nm) and 21200 cm-1 (470 nm), that correspond to spin allowed transitions from 3A2g to 3T2g, 3T1g 3T1g(P) respectively. Also two weak features of spin forbidden transitions to 1Eg and 1T2g at 14100 cm-1 (710 nm) and 18000 cm-1 (560 nm) might be seen at the spectra. These data are consistent with octahedral NiII.36 172 0 10 20 30 40 50 60 0 10000 20000 30000 40000 50000 Wavenumber, cm-1 F ( R ) 9700 13000 19100 25100 Figure 6.7. UV-vis-NIR diffuse reflectance spectrum of UCoP. F(R) corresponds to Kubelka-Monk function and relates reflectance data to the absorption spectra.37 173 0 10 20 30 40 50 60 0 10000 20000 30000 40000 50000 Wavenumber, cm-1 F ( R ) 10100 14100 16700 18000 21200 25600 Figure 6. 8. UV-vis-NIR diffuse reflectance spectrum of UNiP. F(R) corresponds to Kubelka-Monk function and relates reflectance data to the absorption spectra.37 174 There is a broad band with a maximum at 17800 cm-1 (560 nm) in the spectra of UCuP corresponding to the d-d z2 to x2 - y2 transition expected for square planar CuII.36 (Figure 6.9) In addition in all spectra, absorption features centered between 25100 cm-1 ? 26000 cm-1 (380 ? 400 nm) are present with considerable fine structure. These features are assigned to absorption by the uranyl cation.38,39 Magnetization Measurements. The magnetic susceptibility of UCoP was measured as a function of temperature under a constant field of 2000 G. Such a large field was necessary to assure adequate counting statistics at higher temperature. Even at such a high field, only about 40% of the measured signal arises from the sample at higher temperatures. The linearity of the susceptibility of UCoP was verified by independent measurements of the magnetization at 5 K as a function of field at fields up to 2000 G. (Figure 6.10) The temperature dependence of the magnetic susceptibility of UCoP, along with its inverse, is shown in Figure 6.11. From the temperature dependence, there is no indication of any magnetic ordering down to 5 K, the lowest temperature measured. The absence of significant ferromagnetic or spin-glass behavior is supported by a comparison of field-cooled and zero-field-cooled data, which showed no significant differences. The effective magnetic moment, per mole of UCoP sample, was determined to be 4.95(10) B from the fit to the Curie-Weiss law, and measured TIP and  are -0.00012(2) emu/mole Co and 0.3(5) K, respectively. The effective magnetic moment obtained from the temperature dependence of inverse magnetic susceptibility between 100 and 300K is 4.99(5), which is consistent with moments observed in other 175 0 5 10 15 20 25 30 35 40 45 10000 15000 20000 25000 30000 35000 40000 45000 50000 Wavenumber, cm-1 F ( R ) 17200 26000 Figure 6.9. UV-vis-NIR diffuse reflectance spectrum for UCuP. F(R) corresponds to Kubelka-Monk function and relates reflectance data to the absorption spectra.37 176 Figure 6.10. Magnetization of UCoP as a function of applied magnetic field at 5 K. Line is the linear fit to the data. 177 Figure 6.11. Magnetic susceptibility of UCoP as a function of temperature. Line is the fit to Curie-Weiss behavior. Inset shows the temperature dependence of the inverse magnetic susceptibility with a linear fit. 178 systems with octahedrally coordinated Co2+ ion. The TIP is small and is consistent with isolated Co2+ spins, as is the Weiss constant. The magnetization and the temperature dependence of the magnetic susceptibility of UCuP (Figure 6.12 and 6.13) show no evidence of magnetic ordering. Both field cooled and zero field cooled magnetic susceptibilities of UCuP were measured and are statistically indistinguishable, adding support to the absence of magnetic ordering seen from the temperature dependence. The effective magnetic moment for Cu2+ ion in this compound was determined to be 1.9(1) B from the fit to the Curie-Weiss law, and measured TIP and  are 0.0004(1) emu/mole Co and -0.3(5) K, respectively. The effective magnetic moment obtained is consistent with moments observed in other systems with Cu2+ ions. Thermal analysis. In order to estimate the thermal stability of samples we have collected DSC data for UCoP and UCuP compounds. Both structural types demonstrate complicated thermal behavior. Lost of surface water of UCoP at 155 ?C is followed by a series of the endothermic events corresponding to the structural rearrangements, beginning at 273 ?C. (Figure 6.14) Two major events at 400 ?C and 414 ?C could be assigned to the loss of two crystallographically distinct water molecules bonded to the Co center. UCuP consequently loses surface and interstitial water at 130 ?C and 160 ?C. (Figure 6.15) The series of peaks, exothermic - at 290 ?C and endothermic ? at the temperature above 400 ?C, clearly indicate decomposition of the structure and lost of crystallinity at high temperatures. 179 Figure 6.12. Magnetization of UCuP as a function of applied magnetic field at 5 K. Line is the linear fit to the data. 180 Figure 6.13. Magnetic susceptibility of UCuP as a function of temperature. Line is the fit to Curie-Weiss behavior. Inset shows the temperature dependence of the inverse magnetic susceptibility. 181 -1.6 -1.1 -0.6 -0.1 0.4 0.9 0 100 200 300 400 500 600 Temperature, oC H e a t f l o w , m W . . 155oC 273oC 400oC 414oC 592oC536 oC Figure 6.14. DSC data for UCoP. 182 -0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0 0 100 200 300 400 500 600 Temperature, oC H e a t f l o w , m w . . 130oC 160oC 290oC 400oC 510oC 540oC Figure 6.15. DSC data for UCuP. 183 Ion exchange properties. The difference in the structures of UCoP and UCuP leads to their different ion-exchange capabilities. The highest uptake from 1?10-2 M solutions of UCoP demonstrates toward Na+ cations. The value is 0.6(1) meq/g that roughly corresponds to one half of the theoretical capacity and is consistent with the exchange of one out of two Cs+ cations that occupy the channels. The pH of the solutions is not changed significantly during the exchange experiments that suggests protons do not participate in the exchange. The equilibration of the UCoP with CsCl solutions has also revealed no change of pH and Cs+ concentration, which implies that there are no exchangeable protons in the structure. The capability toward Sr2+ and Ca2+ is smaller (0.28(6) meq/g and 0.20(8) meq/g, respectively). Assuming that 2 moles of Cs+ are necessary for exchange with 1 mol of Sr2+ or Ca2+, Sr2+ exchange corresponds only to one quarter of the calculated capacity. These results are significantly higher than Sr2+ uptake 0.1(2) meq/g, measured by radiotracer method. The measurements of radioactivity of Sr solution, containing 90Sr isotope, after different time intervals of exchange, allow one to assume that 80% of exchange occurs within first 10 minutes after mixing. The lower uptake values can be explained by the much higher volume to mass ratio, exceeding 3000:1. In the mixed solutions Ca2+/Na+, Sr2+/Na+ and Ca2+/Sr2+, the uptake of alkaline earth cations by UCoP is highly preferable with no exchange of Na+ occurring. This selectivity can result from the relatively large size of the channels (5.0 ? ? 5.0 ?). Those channels are large enough to accommodate even hydrated cations, and radii of Ca2+ (4.1 ?) and Sr2+ (3.9 ?)40 in their hydrated shells fit better than smaller Na+ (3.6 ?).40 Finally, 184 there is no selectivity found between Sr2+ and Ca2+. They are taken from the mixed solutions in a 1:1 ratio. UCuP ion-exchange experiments showed strong ion-exchange capacities toward double-charged cations with uptake values of 0.49 meq/g and 0.51 meq/g for Ca2+ and Sr2+ respectively and no exchange for Na+ and K+. Again no selectivity between Ca2+ and Sr2+ is found in the mixed solutions. During the experiments it was noted that the pH of the solutions at ion-exchange equilibrium drops by approximately 3 units in all solutions. This might be explained either as participation of the structural protons that belong to the phosphoric acid in the exchange or presence of the excess of phosphoric acid on the surface of material. But due to the very similar change of pH, ranging within 0.4 units for all solutions and no notable dependence of this change from the concentration of present cation, we can rather accept that pH of the solutions is defined by the amount of phosphoric acid remaining on the surface after the synthesis. CONCLUSION The combination of the work reported herein along with our previous disclosures on Cs4[(UO2)2(GaOH)2(PO4)4]?H2O,9 Cs[UO2Ga(PO4)2],9 Cs1.52[(UO2)(TiOH0.48)(PO4)2] ?nH2O, Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O, Cs2[(UO2(VO2)2(PO4)2]?nH2O10 and K4(UO2)(VO)4(OH)2(PO4)4 (Rb4(UO2)(VO)4(OH)2(PO4)4) demonstrates that the construction of mixed-metal uranyl phosphates with three-dimensional framework structures can be accomplished by the interconnection of one-dimensional uranyl phosphate and main group or transition metal phosphate substructures, the linking of uranyl phosphate layers by transition metal centers, or linking of metal-phosphate layers 185 by uranyl centers. The main group or metal centers contained in these structures span from tetrahedral environments in UGaP2 to square planar geometries in UCuP, to square pyramidal coordination in UVVP to octahedral UGaP1, KUVIVP, RbUVIVP, UCoP, and UNiP. Likewise, the UO6 tetragonal bipyramid, the UO7 pentagonal bipyramid, and the UO8 hexagonal bipyramid all occur in this small group of compounds, illustrating the remarkable structural diversity of UVI. It is now clear that the incorporation of additional metal centers into uranyl phosphates provides for substantial enhancement in their ion- exchange, nonlinear optical, and magnetic properties. 186 REFERENCES 1. Burns, P. C.; Miller, M. L.; Ewing, R.C. Can. Mineral. 1996, 34,845. 2. Burns, P. C. Can. Mineral. 2005, 43, 1839. 3. Johnson, C. H.; Shilton, M. G.; Howe, A. T. J. Solid State Chem. 1981, 37, 37. 4. Pozas-Tormo, R.; Moreno-Real, L.; Martinez-Lara, M.; Rodriguez-Castellon, E. Can. J. Chem. 1986, 64, 35. 5. Moreno-Real, L.; Pozas-Tormo, R.; Martinez-Lara, M.; Bruque-Gamez, S. Mater. Res. Bull. 1987, 22, 19. 6. Dieckmann, G. H.; Ellis, A. B. Solid State Ionics 1989, 32/33, 50. 7. Grohol, D.; Blinn, E. L. Inorg. Chem. 1997, 36, 3422. 8. Shvareva, T. Y.; Almond, P. M.; Albrecht-Schmitt, T. E. J. Solid State Chem. 2005, 178, 499. 9. Shvareva, T. Y.; Sullens, T. A.; Shehee, T. C.; Albrecht-Schmitt, T. E. Inorg. Chem. 2005, 44, 300. 10. Shvareva, T. Y.; Beitz, J. V.; Duin, E. C.; Albrecht-Schmitt, T. E. Chem. Mater. 2005, 17, 6219. 11. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2004, 177, 2675. 12. Burns, P. C.; Alexopoulos, C. M.; Hotchkiss, P. J.; Locock, A. Inorg. Chem. 2004, 43, 1816. 13. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2003, 176, 18. 14. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2003, 175, 372. 15. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2002, 167, 226. 16. Locock, A. J.; Burns, P. C. J. Solid State Chem. 2002, 163, 275. 187 17. Doran, M. B.; Stuart, C. L.; Norquist, A. J.; O?Hare, D. Chem. Mater. 2004, 16, 565. 18. Huang, J.; Wang, X.; Jacobson, A. J. J. Mater. Chem. 2003, 13, 191. 19. Krivovichev, S. V.; Burns, P. C.; Armbruster, Th.; Nazarchuk, E. V.; Depmeier, W. Micropor. Mesopor. Mater. 2005, 78, 217. 20. Doran, M.; Norquist, A. J.; O?Hare, D. Chem. Commun. 2002, 24, 2946. 21. Doran, M.; Walker, S. M.; O?Hare, D. Chem. Commun. 2001, 19, 1988. 22. Walker, S. M.; Halasyamani, P. S.; Allen, S.; O?Hare, D. J. Am. Chem. Soc. 1999, 121, 10513. 23. Sheldrick, G. M. SHELXTL PC, Version 6.12, An Integrated System for Solving, Refining, and Displaying Crystal Structures from Diffraction Data; Siemens Analytical X-Ray Instruments, Inc.: Madison, WI 2001. 24. Sheldrick, G. M. SADABS 2001, Program for absorption correction using SMART CCD based on the method of Blessing: Blessing, R. H. Acta Crystallogr. 1995, A51, 33. 25. Buehlmann, P.; Pretsch, E.; Bakker, E. Chem. Rev. 1998, 98, 1593. 26. Grohol, D.; Nocera, D. G. J. Amer. Chem. Soc. 2002, 124, 2641 27. Rose, D.; Chang, Y.-D.; Chen, Q.; Zubieta, J. Inorg. Chem. 1994, 33, 5167. 28. Thue?ry, P.; Nierlich, M.; Souley, B.; Asfari, Z.; Vicens, J. J. Chem. Soc., Dalton Trans. 1999, 15, 2589. 29. Taylor, J. C.; Ekstrom, A.; Randall, C. H. Inorg. Chem. 1978, 17, 3285. 30. Brandenburg, N. P.; Loopstra, B. Acta Crystallogr. 1978, B34, 3734. 31. Sullens, T. A.; Jensen, R. A.; Shvareva, T. Y.; Albrecht-Schmitt, T. E. J. Am. Chem. Soc, 2004, 126, 2676. 188 32. Legros, J. P.; Jeannin, Y. Acta Crystallogr. 1975, B31, 1133. 33. Burns, P. C.; Ewing, R. C.; Hawthorne, F. C. Can. Mineral. 1997, 35, 1551. 34. Brown, I. D.; Altermatt, D. Acta Crystallogr. 1985, B41, 244. 35. Brese, N. E.; O?Keeffe, M. Acta Crystallogr. 1991, B47, 192. 36. Lever, A. B. P. Inorganic Electronic Spectroscopy, 2nd ed.; Elsevier: New York, 1984. 37. Kortum, G. Reflectance Spectroscopy, Springer-Verlag: New York, 1969. 38. Carnall, W. T.; Crosswhite, H. M. In The Chemistry of the Actinide Elements; Katz, J. J., Seaborg, G. T., Morss, J. R., Eds.; Chapman and Hall: London, 1986; Chapter 16. 39. Denning, R. G.; Norris, J. O. W.; Short, I. G.; Snellgrove, T. R.; Woodwark, D. R. In Lanthanide and Actinide Chemistry and Spectroscopy; Edelstein, N. M., Ed.; ACS Symposium Series 131; American Chemical Society: Washington, DC, 1980; Chapter 15. 40. Nightingale, E.R. J. Phys. Chem. 1959, 63, 1381. 189 CHAPTER 7 SUMMARY This chapter is a summary of data and discussions reported in the dissertation, as well as general conclusions. All reported compounds are crystalline solids synthesized under hydrothermal conditions at the temperatures between 180 ?C and 220 ?C. The structures were determined using single crystal X-ray diffraction. Spectroscopic data, second-harmonic generation of light, thermal analysis, magnetic susceptibility, and ion-exchange were employed in order to characterize these materials. Chapter 2 describes the structure and ion-exchange properties of the potassium uranyl iodate, K[UO2(IO3)3] (KUI). This compound was prepared as a continuation of the search for novel uranyl iodates, conducted in our group. KUI has a structure consisting of layers of uranyl iodate with potassium cations located between the layers. The K+ can be exchanged with Cs+ with uptake value of 1.7 meq/g. Two novel uranyl gallium phosphates, Cs4[(UO2)2(GaOH)2(PO4)4]?H2O (UGaP1) and Cs[UO2Ga(PO4)2] (UGaP2) with open-framework structures are discussed in Chapter 3. These two compounds are the first result of our program for preparing uranyl compounds with three-dimensional open-framework structures with a particular focus on the use of octahedral centers, as building units. We suggested that incorporation of the octahedral building blocks in the structure, where bond formation is allowed in axial and 190 equatorial planes, will aid in the formation of frameworks. The three-dimensional structure of UGaP1 is constructed in accordance with this hypothesis from uranyl pentagonal bipyramids, phosphate tetrahedra, and gallium octahedra. The UGaP2 open- framework structure contains uranyl pentagonal bipyramids, phosphate and gallium tetrahedra, and no octahedral building blocks. Both UGaP1 and UGaP2 structures have channels filled with Cs+ cations. Ion- exchange experiments show the ability of the Cs+ cations in UGaP1 to be exchanged with divalent Ca2+, Sr2+, and Ba2+, in UGaP2 - with Ca2+and Ba2+. We were focused on the measurements of Sr2+ uptake in the presence of Na+ due to the potential application of the ion-exchange materials for the removal of 90Sr2+ from HLNW. The Sr2+ uptake in the presence of Na+ cations is 0.4 meq/g. The thermal analysis revealed the absence of any structural changes of UGaP1 up to 400 ?C. These data indicate that UGaP1 could be a potential 90Sr selective material and should be studied further. In Chapter 4,5, and 6 we have proposed a new synthetic approach of utilizing pure element powders as starting materials for hydrothermal synthesis. UVVP, KUVIVP (RbUVIVP), UTiP, UGeP, UCoP, UNiP, and UCuP are obtained using this method. The presence of protons from phosphoric acid and oxygen in reaction vessel allow for slow oxidation of transition metals or germanium and rate-limiting delivery of oxidized reactant to the reaction mixture. In Chapter 4 we reported three novel open-framework uranyl vanadium phosphates Cs2[UO2(VO2)2(PO4)2]?nH2O (UVVP), and isostructural K4(UO2)(VO)4(OH)2(PO4)4 (KUVIVP) and Rb4(UO2)(VO)4(OH)2(PO4)4 (RbUVIVP). The structure of UVVP is built of tetragonal bipyramidal uranium, tetrahedral phosphate, 191 and square pyramidal vanadium building units. The terminal oxo atoms of vanadyl are aligned in one direction, causing the polarity of UVVP (space group Cmc21) and, as result, non-linear optical properties. KUVIVP and RbUVIVP are constructed from uranyl tetragonal bipyramids, tetrahedral phosphate and vanadyl octahedral building blocks. K+ or Rb+ cations are located in the channels and can be exchanged by Cs+ in the presence of Na+. Again, this material can be proposed for removal of 137Cs+ from nuclear waste. KUVIVP and RbUVIVP also demonstrate paramagnetic behavior at the temperatures between 100 K and 320 K and antiferromagnetic ordering at 52 K. The isostructural compounds Cs1.52[(UO2)(TiOH0.48)(PO4)2]?nH2O (UTiP) and Cs1.35[(UO2)(GeOH0.65)(PO4)2]?nH2O (UGeP) are described in Chapter 5. The open- framework structures of these compounds include Ti or Ge in octahedral coordination environments. Ti and Ge octahedra form corner-sharing chains along the same direction as the terminal uranyl oxygen atoms are aligned, providing for the formation of three- dimensional structures. In Chapter 6 we present the structures and properties of Cs2{(UO2)4[Co(H2O)2]2(HPO4)(PO4)4} (UCoP), Cs2{(UO2)4[Ni(H2O)2]2(HPO4)(PO4)4} (UNiP), isostructural with UCoP, and Cs3+x[(UO2)3-xCuH4-x(PO4)5]?H2O (UCuP). Co and Ni in octahedral coordination environment are utilized in order to build UCoP and UNiP open-framework structures. Cu is incorporated in the three-dimensional structure of UCuP in square planar environments. All three open-framework structures have channels, occupied by Cs+ cations. Cs+ can be replaced by Sr2+ in the presence of Na+ with uptake values 0.28 meq/g for UCoP and 0.51 meq/g for UCuP. 192 Thus, we can summarize that ten uranyl-containing open-framework and one layered material have been produced. The novel strategy of incorporation of octahedral building units in order to synthesize open-framework structures was successful. UGaP1, KUVIVP, RbUVIVP, UCoP, and UNiP are constructed in accordance with this strategy. Two of the materials (KUI and KUVIVP (RbUVIVP)) possess ion-exchange properties, selective for Cs+. Cations in four materials (UGaP1, UCoP, UNiP, and UCuP) can be selectivity exchanged with Sr2+. These properties make KUI, KUVIVP (RbUVIVP), UGaP1, UCoP, UNiP, and UCuP promising materials for the selective removal of 137Cs+ and 90Sr2+ from the highly radioactive nuclear waste. Thermally stable UGaP1 could be recommended for the further investigations under the conditions simulating HLNW.