Abstract. High Performance Structures and Materials III 301

Similar documents
UNIT V -CRYSTAL STRUCTURE

Citation for published version (APA): Borsa, D. M. (2004). Nitride-based insulating and magnetic thin films and multilayers s.n.

GROWTH AND CHARACTERIZATION OF NONLINEAR OPTICAL MATERIAL: ALANINE BARIUM CHLORIDE

CRYSTAL STRUCTURE TERMS

Metallic crystal structures The atomic bonding is metallic and thus non-directional in nature

Chapter 1. Crystal Structure

Structure Identification of Ni- Fe Oxide Mixture using Raman Spectroscopy, and X-ray Absorption Near-Edge Spectroscopy

Mössbauer and XRD investigations of layered double hydroxides (LDHs) with varying Mg/Fe ratios

Neutron diffraction study of Mn 3 Fe 4 V 6 O 24

Lecture C4b Microscopic to Macroscopic, Part 4: X-Ray Diffraction and Crystal Packing

Physics of Transition Metal Oxides

Solids SECTION Critical Thinking

Chapter 12 Metals. crystalline, in which particles are in highly ordered arrangement. (Have MP.)

Supplemental Exam Problems for Study

Characterization. of solid catalysts. 5. Mössbauer Spectroscopy. Prof dr J W (Hans) Niemantsverdriet.

Diffraction Basics. The qualitative basics:

Unit-1 THE SOLID STATE QUESTIONS VSA QUESTIONS (1 - MARK QUESTIONS)

CHAPTER THE SOLID STATE

Solids. The difference between crystalline and non-crystalline materials is in the extent of ordering

Lecture 3: Description crystal structures / Defects

IF YOUR ONLY SINGLE CRYSTAL IS NOT REALLY SINGLE

These metal centres interact through metallic bonding

INGE Engineering Materials. Chapter 3 (cont.)

(FeNaP) 2 O 9 Glasses Doped with Different TM Ions

MASTER'S THESIS. Synthesis of rare-earth oxide mesoporous structures by combustion synthesis

Chemistry 145 Exam number 4 name 11/19/98 # Faraday s constant is 96,500 c/mole of electrons.

From sand to silicon wafer

GROWTH AND CHARACTERIZATION OF SEMI-ORGANIC GLYCINE BARIUM CHLORIDE (GBC)

LECTURE 7. Dr. Teresa D. Golden University of North Texas Department of Chemistry

Characterization of Materials Using X-Ray Diffraction Powder Diffraction

3.3 Minerals. Describe the characteristics that define minerals.

metal-organic compounds

High H ionic conductivity in barium hydride

Order in materials. Making Solid Stuff. Primary Bonds Summary. How do they arrange themselves? Results from atomic bonding. What are they?

MAKING SOLID SOLUTIONS WITH ALKALI HALIDES (AND BREAKING THEM)

Chapter 17 Solubility and Complex Ion Equilibria

Nomenclature. A systematic method of writing chemical formulas and naming compounds

9.2.1 Similarities and trends in the properties of the Group II metals magnesium to barium and their compounds

9/29/2014 8:52 PM. Chapter 3. The structure of crystalline solids. Dr. Mohammad Abuhaiba, PE

Mössbauer and XRD study of novel quaternary Sn-Fe-Co-Ni electroplated alloy

Answer All Questions. All Questions Carry Equal Marks. Time: 20 Min. Marks: 10.

The synthesis and characterization of alkaline-earth metal doped Pr 2 Mo 2 O 9 pigments: Applications in coloring of plastics

9/28/2013 9:26 PM. Chapter 3. The structure of crystalline solids. Dr. Mohammad Abuhaiba, PE

Naldrett, D.L. (1987): Rare-earth elements, thermal history, and the colour of natural fluorites, Canadian Journal of Earth Sciences, 24:

Lecture C4a Microscopic to Macroscopic, Part 4: X-Ray Diffraction and Crystal Packing

1.10 Close packed structures cubic and hexagonal close packing

ECE440 Nanoelectronics. Lecture 08 Review of Solid State Physics

Magnetic resonance study of M 3 Fe 4 V 6 O 24 (M = Mg, Zn, Mn, Cu, Co) compounds

This lecture is part of the Basic XRD Course.

Instrument Configuration for Powder Diffraction

CEMS study on diluted magneto titanium oxide films prepared by pulsed laser deposition

Now, let s examine how atoms are affected as liquids transform into solids.

Investigation of different iron sites in -Fe y. N (2<y<3) nanoparticles using Mössbauer spectroscopy. Journal of Physics: Conference Series

MSE 352 Engineering Ceramics II

Observation of Long-Term Corrosion at Nuclear Power Plant Bohunice (Slovakia)

Single crystal X-ray diffraction. Zsolt Kovács

Nanocrystals of the Fe(phen) 2 (NCS) 2 prototypical compound and the size-dependent spin-crossover characteristics

9/16/ :30 PM. Chapter 3. The structure of crystalline solids. Mohammad Suliman Abuhaiba, Ph.D., PE

Kiyoshi Nomura, Alexandre Rykov, Zoltán Németh & Yoshitaka Yoda

CHEMISTRY. SCIENCE Paper 2. (Two hours) You will not be allowed to write during the first 15 minutes.

Influence of preparation method on the performance of Mn Ce O catalysts

Synthesis calcium-titanate (CaTiO 3 )

Zirconium Oxide X-ray Diffraction Data Processing ByRietveld Analysis Method

Thin Film Scattering: Epitaxial Layers

Review of Metallic Structure

CHAPTER 8 SYNTHESIS AND CHARACTERIZATION OF COPPER DOPED NICKEL-MANGANESE MIXED FERRITE NANOPARTICLES BY CO-PRECIPITATION METHOD

PRECIS AND CONCLUSIONS

SOLID STATE

KEY FIRST MIDTERM EXAM Chemistry February 2009 Professor Buhro

A Designed 3D Porous Hydrogen-Bonding Network Based on a Metal-Organic Polyhedron

Chapter 5. Influence of Pb 2+ doping on photoluminescence properties CaSiO 3 : Mn 2+ nanophosphor

Department of Physics National Institute of Technology, Rourkela

CHAPTER. The Structure of Crystalline Solids

Spreadsheet Applications for Materials Science

STRUCTURE AND MAGNETIC PROPERTIES OF THE Zr 30

Crystal Defects. Perfect crystal - every atom of the same type in the correct equilibrium position (does not exist at T > 0 K)

(iii) Describe how you would use a powder diffraction pattern of this material to measure

SOLUBILITY LIMITS OF ERBIUM IN PARTIALLY DISORDERED CRYSTALS LANGANITE AND LANGATATE *

X-Ray Diffraction. Nicola Pinna


Identification of Crystal Structure and Lattice Parameter. for Metal Powders Using X-ray Diffraction. Eman Mousa Alhajji

Generation Response. (Supporting Information: 14 pages)

Chapter1: Crystal Structure 1

Our country, our future S2 CHEMISTRY DURATION: 2 HOUR

Supplementary Figures

X-ray Diffraction (XRD)

Synthesis and Characterization of Cadmium Sulfide Nanoparticles

Near-infrared luminescence of rare earth ions in oxyfluoride lead borate glasses and transparent glass-ceramic materials

CHAPTER 3: CRYSTAL STRUCTURES & PROPERTIES

GROWTH AND CHARACTERIZATION OF SEMI-ORGANIC GLYCINE POTASSIUM CHLORIDE (GPC)

Supplementary Figures

Low- and high-temperature oxidation of Mn 1.5 Al 1.5 O 4 in relation to decomposition mechanism and microstructure

Synthesis and characterization of Aln 2 O 4 indates, A ˆ Mg, Ca, Sr, Ba

Studies on the thermal expansion of nanocrystalline boron carbide

IN-SITU XRD TO OPTIMIZE POWDER SYNTHESIS OF AURIVILLIUS PHASES

Results and Characterization of Electroless Nickel Phosphorus Alloy. The micro-structure of freshly deposited EN films seems to be independent of the

Study of electric quadrupole interactions at 111 Cd on Zn sites in RZn (R = Ce, Gd, Tb, Dy) compounds using the PAC spectroscopy

Solid State Chemistry CHEM-E4155 (5 cr)

CHEM 200/202. Professor Gregory P. Holland Office: GMCS-213C. All s are to be sent to:

Journal of Nuclear and Radiochemical Sciences, Vol. 11, No.1, pp. 1-5, 2010

Transcription:

High Performance Structures and Materials III 30 The study of surface oxidation of tin(ii) fluoride and chloride fluoride materials by Mössbauer spectroscopy: to oxidize or not to oxidize, that is the question G. Dénès, E. Laou, M. C. Madamba & A. Muntasar Laboratory of Solid State Chemistry and Mössbauer Spectroscopy, Department of Chemistry and Biochemistry, Concordia University, Montréal, Canada Abstract Experimental methods designed to study the bulk of materials do not necessarily detect the changes taking place at the surface of the crystallites. For example, divalent tin-containing materials appear to be stable at ambient conditions in air, provided they are not hygroscopic, X-ray powder diffraction shows only the peaks of the expected tin(ii) phase. However, we have observed that the Mössbauer spectrum of polycrystalline samples contain, in addition to the expected tin(ii) peak(s), a small peak at 0 mm s relative to CaSnO 3 at ambient conditions, that can be attributed only to tin(iv) coordinated by oxygen. A detailed study of this phenomenon has shown that Mössbauer spectroscopy is quite sensitive for detecting thin layers of oxide at the surface of crystallites of tin(ii). This phenomenon has been exploited for the study of spontaneous oxidation of various tin(ii) fluoride and chloride-containing materials, some of these fluorides being the highest performance fluoride-ion conductors known to date. It was observed that passivation is quite efficient in the fluorides, and in the chloride fluorides that have all their tin(ii) covalently bonded. On the other hand, the materials containing a mixture of covalently bonded tin(ii) and the Sn 2+ stannous ion namely, the Ba -x Sn x Cl +y F -y solid solution, show a higher rate of oxidation, which is highly dependent on the method of preparation and the composition parameters, x and y. Keywords: passivation, oxidation, disordered phases, fluorite-type structure, Mössbauer spectroscopy, X-ray diffraction, ionic conductivity, BaClF. doi:0.2495/hpsm06029

302 High Performance Structures and Materials III Introduction The MF 2 fluorides of large divalent metals (M = Ca, Sr, Ba & Pb) have the well known fluorite type structure, with a cubic unit-cell, space group Fm3m, and a cubic coordination of the metal ion (fig. ) []. This contrasts with SnF 2 tin(ii) fluoride (stannous fluoride) that has a molecular structure; it is made of Sn 4 F 8 tetramers and the Sn-F bonds are strongly covalent [2]. When SnF 2 and MF 2 are combined together, some new materials are produced, some of which are ordered, others disordered. The MSnF 4 compounds have a M M Sn Sn order parallel to the c axis of the unit-cell (fig. ) [3]. Tin and lead belong to group 4, and therefore they have four valence electrons, and in the +2 suboxidation state, the 5s 2 (Sn) or 6s 2 (Pb) electrons are unused and form a nonbonding electron pair, also called a lone pair. When bonding is ionic, the orbitals are not hybridized and the lone pair is located on the ns native orbital that has spherical symmetry, unless it is distorted by polarization, and a quite regular coordination is observed, and the lone pair is said to be non-stereoactive since it does not modify the coordination. In fluorite-type β PbF 2, lead has a cubic coordination, therefore its lone pair is located on the native 6s orbital and bonding is ionic, like in BaF 2 (fig. ). In tetragonal MSnF 4, the coordination of M is a distortion of the MF 8 cube of the fluorite structure, with in addition, four longer interactions, to form an overall MF 4 F 4 F 4 where the bond length increases as follows: M-F (from M-F-M bridges) < M-F (from M-F eq -Sn bridges) < M-F (from M-F ax -Sn bridges), where F eq and F ax form equatorial (4) and axial () bonds with tin, respectively (fig. ). For the sake of clarity, the Ba- F bonds are not shown on figure. The lead coordination in β PbF 2, is still ionic since Sr 2+ and Ba 2+ take the same coordination in SrSnF 4 and in BaSnF 4, respectively. In contrast, tin forms covalent bonding in MSnF 4, with a very short axial Sn-F bond, and the lone pair is located on a hybrid orbital, in transposition to F (fig. ). The tin lone pair is said to be stereoactive since it changes drastically the tin stereochemistry (lower coordination number, highly distorted coordination). A consequence of the M/Sn order and of the stereoactivity of the tin lone pair is that the lone pairs cluster in sheets that are cleavage planes, which make the structure very highly two-dimensional. Examples of disordered structures are PbSn 4 F 0, the M -x Sn x F 2 solid solution (M = Ca and Pb), and µγ- MSnF 4 (M = Ba and Pb) obtained by ball-milling. The disordered phases have an Fm3m cubic unit-cell like the fluorite-type MF 2, and their diffraction pattern shows no peak splitting or additional low angle peak that would indicate the presence of a lattice distortion or a superstructure. Therefore, Sn and the other metal M are fully disordered on the metal ion site of the MF 2 structure (fig. 2). The M bonding is ionic and the coordination is cubic with some disordered local distortions in the neighborhood of tin, while tin bind to only one of the faces of the F 8 cubes, with its stereoactive lone pair pointing towards the center of the cube, and thus its bonding is covalent. BaClF crystallizes in the PbClF structure (space group P4/nmm). Its structure is also a tetragonal distortion of the fluoritetype BaF 2, with layers of F and Cl ordered parallel to the c axis of the unit-cell, making it a layered structure, in contrast with the three-dimensional cubic

High Performance Structures and Materials III 303 structure of BaF 2 (fig. ). We have recently found two methods for substituting substantial amounts of Ba by Sn in BaClF, and also some F by Cl (y>0), or some Cl by F (y<0), making it a doubly disordered Ba x Sn x Cl +y F -y solid solution. X- ray diffraction shows that Ba and Sn are fully disordered on the Ba site of the BaClF structure. In addition, while the two anionic sites remain ordered, for y>0 the y Cl in excess are disordered with the remaining ( y) F on the F site, whereas for y<0, the ( y) excess F are disordered with the remaining (+y) Cl on the Cl site. However, the electronic situation at tin cannot be assessed by X-ray diffraction. All the tin(ii)-containing fluorides studied here are high performance fluoride ion conductors (about three orders of magnitude higher than the conductivity of the corresponding MF 2 ). Figure : Projection of the crystal structure of BaSnF 4, BaF 2 and BaClF. All these materials appear stable in air at ambient conditions, and X-ray diffraction shows no sign of oxidation to tin(iv). The present study shows that tin-9 Mössbauer spectroscopy is an excellent method for detecting minute amounts of tin(iv) oxide in tin(ii) compounds, and that all materials show at the minimum some surface oxidation, and some passivate very well while other do so less efficiently. 2 Experimental Fluoride materials synthesis was carried out either in aqueous medium or at high temperature under dry conditions. Tetragonal α-pbsnf 4 was obtained by precipitation on addition of an aqueous solution of lead(ii) nitrate to a solution of SnF 2. Ca -x Sn x F 2 precipitated when a solution of calcium nitrate was added to a solution of SnF 2 at high Ca/Sn ratio, or by leaching, on stirring, CaSn 2 F 6 in a large amount of water. The synthesis of Pb -x Sn x F 2 and BaSnF 4 was carried out under nitrogen in sealed copper tubes heated at 500 o C and quenched. Barium

304 High Performance Structures and Materials III tin(ii) chloride fluorides were prepared by reaction of barium nitrate and tin(ii) fluoride in aqueous solutions versus the X = BaCl 2 /(SnF 2 + BaCl 2 ) molar ratio. Three new materials were obtained: BaSn 2 Cl 2 F 4 (0.25<X<0.35), BaSnClF 3.0.8H 2 O (0.50<X<0.80), and a material that has the diffraction pattern of BaClF (X>0.87). Chemical analysis showed that it is a Ba -x Sn x Cl +y F -y solid solution (x = 0.5, y = 0.). A very similar BaClF phase was obtained by synthesis in dry conditions when mixtures calculated according to eqn () were heated three days under nitrogen in sealed copper tubes at temperatures varying from 350 to 800 o C. 2x y BaF 2 2 + xsnf 2 + y + BaCl Ba x Sn x Cl + y F 2 2 y () Figure 2: Model of a disordered fluorite type M -x Sn x F 2 structure. X-ray powder diffraction was carried out on a PW-050/25 Philips diffractometer automated with the SIE RAY 22 system from Diffraction Technology, using the K α radiation of copper (λ Kα =.5405 Å) and a monochromator. Mössbauer spectra were recorded in a TN7200 multichannel analyzer from Tracor Northern, using an Elscint driving system, a Harshaw (Tl)NaI detector and a 0 mci CaSnO 3 γ-ray source. All chemical isomer shifts were referenced relative to a standard CaSnO 3 absorber. All spectra were collected at ambient temperature and fitted using the GMFP software. Samples were studied when young (freshly prepared) and old (after several years storage in air at ambient temperature).

High Performance Structures and Materials III 305 0.96 0.92 0.88 (a) 0.995 0.975 0.955 0.935 (b) Velocity (mm/s) Figure 3: Ambient temperature 9 Sn Mössbauer of BaSnF 4 prepared by the dry method at 500 o C for 2 hours: (a) young sample, (b) sample aged for 69 months at ambient conditions. 3 Results and discussion 3. Oxidation of the fluorides Various ordered and disordered tin(ii)-containing fluorides were checked for oxidation, using X-ray diffraction and Mössbauer spectroscopy. Figure 3 shows the spectrum of BaSnF 4 when freshly prepared and after 69 months of age. Both spectra have a large quadrupole doublet in the 2-4 mm/s velocity range, which is characteristic of the presence of a stereoactive lone pair on tin, and confirms that bonding is covalent (fig. ). In addition, a weak peak is observed at ca. 0 mm/s. The peaks positions are given relative to a tin(iv) oxide standard, CaSnO 3, and therefore, the peak at 0 mm/s in BaSnF 4 shows some tin(iv) oxide is contained in the ample, even when freshly prepared. This can be assigned to Sn(II) to Sn(IV) oxidation of the surface of the solid particles by atmospheric oxygen. After aging 69 months in air at ambient conditions, the tin(iv) signal at ca. 0 mm/s has not increased significantly, therefore no much further oxidation occurred during the nearly 6 years of aging. It is therefore established that there is an initial very rapid surface oxidation of the particles that creates a passivating layer, which protects the bulk of the particles against further oxidation. X-ray diffraction does not show any peak foreign to BaSnF 4, and more particularly no

306 High Performance Structures and Materials III SnO 2 peak. This suggests that either the amount of SnO 2 is too small to be detected by diffraction, or the oxide is amorphous, or the particles are too small to diffract. This establishes the superiority of Mössbauer spectroscopy to detect small amounts of oxide in tin(ii) halides. The relative intensity of Mössbauer peaks is a function of both the amount of each kind of tin and their recoil-free fraction f a. The recoil-free fraction is the fraction of 9 Sn nuclei that absorb γ rays without dissipating the recoil energy through phonons (lattice vibrations). The stronger the lattice, the higher the recoil-free fraction, and the stronger the Mössbauer peaks at equal amount of tin. Since Sn(IV) forms very strong bonds to oxygen in the highly tight rutile structure of SnO 2, it has a high recoil-free fraction and gives a disproportionably high signal, in comparison with small recoil-free fraction tin(ii) halides, that have much weaker bonds. It results that the fraction of SnO 2 in a tin(ii) halide is more than ten times smaller than its relative signal. Therefore, the amount of SnO 2 in BaSnF 4 is much smaller than the intensity of it signal suggests. Disordered systems, even very poor in tin, like the M -x Sn x F 2 at small tin content (x = 0.27 for M = Ca, x = 0.0 for M = Pb) also show no SnO 2 diffraction peak (fig. 4Ab), and the tin(iv) Mössbauer peak is barely detectable (fig. 4Cb), showing excellent passivation. (A) (B) (C) Figure 4: (A) X-ray powder diffraction pattern of CaF 2 and Ca -x Sn x F 2 (x = 0.27), (B) X-ray powder diffraction pattern of BaClF and Ba x Sn x Cl +y F -y (x = 0.5, y = 0.094), (C) 9 Sn ambient temperature Mössbauer spectrum of Ba -x Sn x Cl +y F -y (x = 0.5, y = 0.094) and Ca -x Sn x F 2 (x = 0.27).

High Performance Structures and Materials III 307 3.2 Oxidation of the chloride fluorides The stoichiometric chloride fluorides, BaSn 2 Cl 2 F 4 and BaSnClF 3.0.8H 2 O have a behavior similar to that of the fluorides. Freshly prepared samples have a tin(iv) oxide peak, and therefore there is a rapid initial surface oxidation (fig. 5), and it also does not show any significant increase after aging, therefore efficient passivation occurs. X-ray diffraction does not detect the oxide. The stronger tin(iv) Mössbauer signal of BaSn 2 Cl 2 F 4, in comparison to BaSnClF 3.0.8H 2 O, does not mean that the former compound contains more SnO 2. It could just mean that tin(ii) in BaSn 2 Cl 2 F 4 has a lower recoil-free fraction than in BaSnClF 3.0.8H 2 O. Very low temperature spectra would be required to determine the reason for their unequal tin(iv) peak intensity. 0.93 (a) 0.86 0.97 0.94 0.9 0.88 (b) 0.85 Velocity (mm/s) Figure 5: Ambient temperature 9 Sn Mössbauer spectrum of: (a) BaSn 2 Cl 2 F 4 and (b) BaSnClF 3.0.8H 2 O. The behavior of the non-stoichiometric Ba -x Sn x Cl +y F -y is much more complex. The Mössbauer spectrum of the precipitated solid solution is made of a single tin(ii) line at ca. 4 mm/s, characteristic of ionic bonding, and of a tin(iv) oxide peak at ca. 0 mm/s (figs. 4Cb & 6). Like in the materials above, no SnO 2 could be detected by X-ray diffraction (fig. 4Bb). Contrary to the fluorides and the stoichiometric chloride fluorides, a significant increase of the intensity of the tin(iv) oxide Mössbauer peak takes place on aging for 4 years. Therefore,

308 High Performance Structures and Materials III although the material is still quite well protected against oxidation, it is not fully passivated. 0.99 0.98 0.97 0.96 0.95 0.94 (AM-234, X = 0.860, aged for 2.5 m) (a) (d) 0.93 Velocity (mm/s) 0.985 0.965 0.945 (AM-234, X = 0.860, aged for 48.0 m ) (b) Figure 6: 0.925 Velocity in (m m /s) Ambient temperature 9 Sn Mössbauer spectrum of the precipitated Ba x Sn x Cl +y F -y versus age: (a) 2.5 month old, (b) 48.0 months old. The rate of oxidation of the Ba -x Sn x Cl +y F -y solid solution prepared by high temperature direct reactions in dry conditions is highly dependent on the x and y composition parameters (fig. 7). At low tin content (low x), freshly prepared samples have a tin(iv) oxide Mössbauer peak quite stronger than the tin(ii) peak (fig. 7a ). This shows that initial resistance to oxidation is much weaker than in the low tin content fluoride solid solutions. For example, the tin(iv) peak of Pb -x Sn x F 2 at low tin content (x = 0.0) is barely detectable. After less than three years aging, the tin(iv) peak of figure 7a has much increased and the tin(ii) peak is now a minority, therefore passivation is much poorer. When the tin content increases to x = 0.2 at y = 0 (Cl/F = ), the

High Performance Structures and Materials III 309 tin(iv) peak is weaker than the tin(ii) peak, however, it is stronger than that of stoichiometric chloride fluorides. (AM-80, x = 0.05, y = 0.0) 0.999 (AM-80, x = 0.05, y = 0.0) 0.995 0.994 0.99 0.989 0.985 (a ) 0.984 (a 2 ) 0.98 0.979 (AM-35, x = 0.2, y = 0.000) AM-35, x = 0.2, y = 0.00 0.99 0.99 0.98 0.97 0.96 0.95 (b ) 0.98 0.97 0.96 0.95 (b 2 ) 0.94 0.94 (AM-55, x = 0.225, y = 0.200) (AM-55, x = 0.225, y = 0.2, RT aged) 0.998 0.995 0.988 (c ) 0.99 0.985 (c 2 ) 0.978 Velocity (mm/s) 0.98 Velocity (mm/s) Figure 7: 9 Sn Mössbauer spectra of Ba -x Sn x Cl +y F -y at RT, prepared by direct reactions under dry conditions (freshly prepared and aged): (a & a 2 ) AM-80, x = 0.050, y = 0.0 [freshly prepared (a ) and aged for 32 months (a 2 )], (b & b 2 ): AM-35, x = 0.2, y = 0.000 [freshly prepared (b ) and aged for 36.7 months (b 2 )], (c & c 2 ): AM-55, x = 0.225, y = 0.200 [freshly prepared (c ) and aged 3.0 months (c 2 )].

30 High Performance Structures and Materials III Upon aging to three years, the tin(iv) peak dominates slightly. This shows that both the oxidation is slower at higher tin content, and passivation is somewhat better, although not great. Near the upper limit of the solid solution in tin (x = 0.225) and at high chlorine content (y = 0.200), the tin(ii) signal predominates by far, however, it is also very weak, which shows that its recoilfree fraction is extremely small. After aging two and a half years, the tin(iv) signal is twice as strong as the tin(ii) signal, therefore initial oxidation was not very large, however, passivation was not very good. These results can be understood, at least in part, from the bonding model of tin in the solid solution. At low x, Sn is diluted in the solid and there is little chance of a Sn being neighbor to another Sn. The coordination of an isolated Sn is mostly determined by the size of the Ba site it occupies. Since Ba 2+ is quite larger than Sn 2+, the latter is loose in its site, i.e. its bonds are weak, resulting in a weak recoil-free fraction and fast oxidation (figs. 7a & 7a 2 ). At higher tin content, the rest of the solid network has to accept adjustments to satisfy the tin bonding requirements, which allows for stronger bonds, hence a higher recoilfree fraction and some improvement of passivation (figs. 7b & 7b 2 ). However, at high x and high y, the excess chlorine, probably clusters around tin than around barium because there is more room around the smaller tin. However, since Sn-Cl bonds are weaker than Sn-F bonds, it results in a decrease of recoilfree fraction and passivating power (figs. 7c & 7c 2 ). 4 Conclusion Mössbauer spectroscopy was an essential method for understanding the complex bonding of tin(ii) and its oxidation to tin(iv) in the fluorides and chloride fluorides. Acknowledgements This work was made possible by the support of Concordia University and the Natural Science and Engineering Research Council of Canada. Grateful thanks are also due to the Procter and Gamble Co. (Mason, Ohio) for supporting our Mössbauer laboratory. References [] Rayner-Canham, G. & Overton, T., Descriptive Inorganic Chemistry, 3 rd ed., Freeman, New York, pp. 78-80, 2002. [2] Dénès, G., Pannetier, J., Lucas, J. & Le Marouille, J.Y., About SnF 2 stannous fluoride. I. Crystallochemistry of α-snf 2. J. Solid State Chem., 30(3), pp. 335-343, 979. [3] Birchall, T., Dénès, G., Ruebenbauer, K. & Pannetier G., A neutron diffraction and 9 Sn Mössbauer study of PbSnF 4 and BaSnF 4. Hyperf. 29, pp. 33-334, 986.