ESQ-905511024“. SUBSTSTUTION EN CRYSTALUNE SUBSTANCES Thesis for £116 Dogma 6% Ph. Ba MICHEGAN STATE SNEVERSETY Roy Meficeim Miiler 1960 '5 £1.13“: This is to certify that the thesis entitled \ ISO- POSITIONAL SUBSTITUTION IN CRYSTALLINE SUBSTANCES presented by ROY MALCOLM MILLER has been accepted towards fulfillment of the requirements for effiwzam Major professor Date May 19, 1960 0-169 LIBRARY Michigan State University ISO-POSITIONAL SUBSTITUTION IN CRY STALLIN E SUBSTANCE S By Roy Malcolm Miller AN ABSTRACT Submitted to the School for Advanced Graduate Studies of Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Geology 1960 Approved ll 3 . 5&W ABSTRACT The iso-positional substitution of ions in crystalline sub- stances is not clearly understood. V. M. Goldschmidt’s concept of ionic size, valency, and temperature as factors controlling substitu- tion in ionic crystals is subject to investigation. AnhydrOus alumi- num phosphate was selected as a crystalline substance for experi- mental purposes due to its exceptionally close structural similarity to cx-quartz. Crystalline aluminum phosphate was produced from aqueous solution under pressure at 163°C. Minor quantities of tri- valent gallium, iron, chromium, cobalt, and vanadium were allowed to enter the crystalline framework in turn. The percentage of impurity substituted was determined by spectroscopic analysis while X-ray diffraction patterns indicated the corresponding changes in cell di- mensions. Gallium substitution caused an 3 axis contraction with increasing impurity. Iron and chromium entered the structure in amounts less than 1.0 percent, expanding the 3 axis and contracting the 3 axis. Less than 0.05 percent cobalt and vanadium were ac- cepted by the crystal. Neither the ionic radius, electronegativity, or ionization potential of the cations used in this investigation ii appear to be the controlling factors in their success in substituting iso-positionally for aluminum. The electron configuration of the trace ion attempting to proxy for the aluminum in the framework appears to be the prime factor in its occupying a given structural site. iii ISO-POSITIONAL SUBSTITUTION IN CRY STALLINE SUB STANCE S By ROY MALCOLM MILLER A THESIS Submitted to the School for Advanced Graduate Studies of Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Geology 1960 C3 20 865 5/24/55. ACKNOWLEDGMENT S The author expresses deep appreciation to Dr. H. B. Stone- house for the inspiration and direction given during the course of this investigation. The cooperation of the Department of Geology and guidance of the committee members are gratefully acknowledged. Many thanks are due the Department of Soil Science, Michi- gan State University, and the Department of Mineralogy, University of Michigan, for the use of laboratory facilities. The advice and encouragement of Professors Ralph Erickson and L. S. Ramsdell of the respective departments are accepted with gratitude. Dr. J. D. Hill and Dr. J. Zinn very kindly contributed to the progress of this investigation. The author thanks Mr. Gene Mondro for assistance with photography. The help and stimulus provided by fellow staff members of the Henry Ford Community College as the project grew are highly appre- ciated. The study was supported by a grant from the Petroleum Re- search Fund administered by the American Chemical Society. iv TABLE OF CONTENTS Page ABSTRACT .................................. 11 ACKNOWLEDGMENTS ........................... 1v LIST OF TABLES ............................. vi LIST OF FIGURES ............................. ix INTRODUCTION ............................... 1 CRYSTAL CHEMISTRY .......................... 8 Crystal Theory ........................... 9 Crystal Growth ........................... 21 EXPERIMENTAL PROCEDURE ..................... 32 Addition of Trace Ions ..................... 33 Analytic Procedure ........................ 41 Treatment of Data ......................... 46 RESULTS ................................... 55 DISCUSSION OF RESULTS AND CONCLUSIONS ........ 76 Acceptance of Impurities .................... 77 Variation of a and g with Impurities ............ 78 Reasons for Acceptance of Impurities ........... 84 Conclusions ............................. 98 SUGGESTIONS FOR FURTHER WORK ............... 99 BIBLIOGRAPHY ............................... 101 TABLE LIST OF TABLES A Comparison of Closely Measured Physical Pr0perties of oc-Quartz and Aluminum Phosphate ........................... A Comparison of Less Well Defined Physical Properties of SiSiO4 and AlPO4 ........... A Comparison of Physical Constants of Silicon and Phosphorus ....................... Physical Data Concerning Crystalline Ortho- phosphate Compounds Displaying Structures Similar to Silica ...................... A Comparison of X-Ray Diffraction Data for Commercially Prepared Aluminum Phosphate with the Standard d- Spacings for Aluminum Phosphate Hydrate ..................... A Tabulation of the Impurity Content of Ortho- phosphoric Acid as Supplied by the Manu- facturer ............................ A Comparison of X-Ray Diffraction Data for the AlPO4 Crystals with Standard Data from the ASTM Index .................. Ionic Radii Values Assigned the Cations Linking the P04 Tetrahedra in the Crystal Structure of AlPO4 ............................ A List of Substances Supplying Cation Impuri- ties for Substitution in AlPO4 ............. vi Page 11 13 16 23 24 29 34 36 TABLE 10. 11. 12. 13. 14. 15. 16. 17-A. 17-B. 17-C. 17-D. 1'7-E. 18-A. A List of Proposed Mol Ratios between Impurity Cations and AlPO4 ............... A Table of Weights of Selected Impurities Added to the Crystallizing AlPO4 Solution The Most Effective Instrumental Settings for the X-Ray Diffractometer ................ A Comparison of 20 Values for Selected Dif- fraction Peaks with Those Determined by Huttenlocher (1935) ..................... A Table of Weighting Coefficients Assigned to Values of a and g ..................... Inherent Contamination of AlPO4 Crystals Due to Trace Elements Present in the Reacting Materials ............................ The Results of Spectrosc0pic Analysis of AlPO4 Crystals in Which Specific Impurities Were Incorporated ......................... Experimental 26 Values for Crystals of AlPO4 Containing Gallium as an Impurity .......... Experimental 26 Values for Crystals of AlPO4 Containing Iron as an Impurity ............. Experimental 20 Values for Crystals of AlPO4 Containing Chromium as an Impurity ......... Experimental 29 Values for Crystals of AlPO4 Containing Cobalt as an Impurity ........... Experimental 26 Values for Crystals of AlPO4 Containing Vanadium as an Impurity ......... Unweighted a and 2/ 2 Values (in A) for Crystals of AlPO4 Containing Gallium .............. vii Page 37 38 45 47 54 56 57 63 64 65 66 67 68 TABLE Page 18-B. Unweighted a and _c_/ 2 Values (in A) for Crystals of A1P04 Containing Iron ................ 69 18-C. Unweighted _a_ and 2/2 Values (in A) for Crystals of AlPO4 Containing Chromium ............. 71 18-D. Unweighted a and 3/ 2 Values (in A) for Crystals of A1P04 Containing Cobalt ............... 72 18-E. Unweighted a and 9/ 2 Values (in A) for Crystals of A1P04 Containing Vanadium ............. 73 19. The Best Experimental 3 and 9/2 Values for AlPO4 Crystals Containing Gallium, Iron, Chromium, Cobalt, and Vanadium ........... 74 20. Electronegativity Values Assigned the Cations Linking the P04 Tetrahedra in the Crystal Structure of A1P04 .................... 37 21. Ionization Potentials Assigned the Cations Linking the P04 Tetrahedra in the Crystal Structure of A1P04 ............................ 90 22. Electron Configurations Assigned the Cations Linking the P04 Tetrahedra in the Crystal Structure of A1P04 .................... 91 viii Figure 1. 10. LIST OF FIGURES Schematic free-energy vs. temperature plot showing stable and metastable equilibria between phases in the system AlPO4 ........ A diagram of the orthophosphate ion without regard for the electron pairs involved ....... The orthophosphate ion depicted as a resonance hybrid .............................. An idealized diagram of the structural arrange- ment of the ions in crystalline A1P04 ........ An exploded drawing of the stainless steel pressure vessel used in growing crystalline AlPO4 .............................. An enlarged photograph of assorted crystals of AlPO4 ............................ An enlarged photograph of a representative single crystal of A1P04 ................. A comparison of the internal structures of aluminum phosphate and 0<~quartz ........... A graphic solution of the quadratic formula as applied to two lattice planes of a hexagonal crystal and solved simultaneously for a and g . . . A diagram of the weighting factors versus vari- ation in axial lengths for different combina- tions of the quadratic formula representing lattice planes ......................... ix Page 12 14 15 18 25 27 28 31 48 52 Figure 11. 12. 13. 14. 15. 16-A. 16-B. 16-C. 16-D. 16-E. comparison of selected diffraction peaks of crystalline A1P04 containing gallium with those of uncontaminated AlPO4 ............. comparison of selected diffraction peaks of crystalline AlPO4 containing iron with those of uncontaminated A1P04 ................. comparison of selected diffraction peaks of crystalline AlPO4 containing chromium with those of uncontaminated A1P04 ............. comparison of selected diffraction peaks of crystalline A1P04 containing cobalt with those of uncontaminated A1P04 ............. comparison of selected diffraction peaks of crystalline A1P04 containing vanadium with those of uncontaminated A1P04 ............. diagram of axial length versus percent impurity for A1P04 crystals containing gallium ............................. diagram of axial length versus percent impurity for A1P04 crystals containing iron ............................... diagram of axial length versus percent impurity for AlPO4 crystals containing chromium ............................ diagram 'of axial length versus percent impurity for AlP04 crystals containing cobalt .............................. diagram of axial length versus percent impurity for AlPO4 crystals containing vanadium ............................ Page 58 59 60 61 62 79 80 81 82 83 INTRODUCTION INTRODUCTION Fractional crystallization from the melt and substitution of one element for another appear to be the primary chemical mechan- isms creating the crust of the earth. Bowen (1922) is credited with establishing the general pattern followed in fractional crystallization, while Clarke and Washington (1922, 1924) provided quantitative evi- dence that eight elements (0, Si, Al, Fe, Ca, Na, K, and Mg) con- stitute the bulk of the earth’s crust. All other elements are called as a group, accessory, minor, or trace elements (Rankama and Sahama, 1950, p. 34). The classical investigation of the structure and properties of crystals by V. M. Goldschmidt (1926) directed attention to the prin- ciples governing distribution of trace elements among the magmatic minerals. Substances of analogous chemical formula which show similar crystal structure were described as isomorphous, a term which Goldschmidt (1931) credits to Mitcherlich. According to Gold- schmidt, isomorphism appears when the relative size of the crystal unit and the relative strength of polarization (within limits) are equal for the two substances in question. Considerable interest was de- veloped in the phenomenon in which atoms or ions of particular 2 elements occupy the structural positions of atoms or ions of other elements in a mineral. Strunz (1937) used the term “diadochic” in discussing this structural relationship and attributes the word to Professor Niggli. Making the assumption that ionic bonding prevails and the ion making the larger contribution to the energy of the crystal structure is preferred, Goldschmidt (1937) proposed a set of rules governing diadochic substitution of ions in magmatic minerals: a. For two ions to substitute diadochically one for another in a crystal structure, the ionic radii must not differ by more than 15 percent. b. When two ions possessing the same charge, but different radii, compete for a position in the crystal structure, the ion with the smaller radius is preferred. c. Ions having similar radii but different charges (of the same sign) may substitute diadochically in a crystal; in which case the ion having the larger charge has preference over the ion with the smaller charge. The discovery that ionic size of trace elements govern their substitutional behavior to a greater extent than their position in the periodic table was accepted as a fundamental relationship. However, in recent years workers in the field have expressed the need for revision of the rules governing trace element distribution among common rocks and minerals. Fyfe (1951) and Ramberg (1952) called attention to necessary modifications. Shaw (1953) questioned the validity of the rules and listed exceptions to them. Ahrens (1952) points to chemical bonding as the answer to much of the questionable geochemical phenomena, with ionization potential as the clue to be investigated. Ringwood (1955) proposes to modify the Goldschmidt rules dealing with replacement of trace elements in magmatic miner- als to allow for differing electronegativities of elements which dis- play diadochy. The recent work of Kaiser and Bond (1959) on dia- mond structure is illustrative of the current interest displayed by investigators in the fields of solid state physics, crystallography, ceramics, and geochemistry, concerning the relationship between very minor amounts of impurity in crystals and the corresponding alteration in crystalline physical properties. The occurrence of silicon in the upper lithosphere makes diadochic substitution in silicates of extreme importance to mineralo- gists and geochemists. W. L. Bragg (1930) outlined the general na- ture of many crystalline silicates by means of X-ray diffraction studies. Subsequent effort has been made to extend this knowledge and to establish relationships between known silicate structures and the elements occupying specific positions in the frameworks. One recent investigation (Goldsmith, 1950) is that of iso-positional re- placement of Ga (III) and Ge (IV) for Al (III) and Si (IV) in sodium and potassium feldspars. The production of silica tetrahedral structures requires pressures and temperatures not ordinarily available in mineralogy laboratories. In one of the early studies concerning the crystallog- raphy and stereochemistry of the inorganic compounds Goldschmidt (1932) points to the possibility of YPO4, YAsO4, and CaCrO4 form- ing isomorphous structures with quartz. Research organizations en- gaged in silicate studies have continued to investigate the silicate structures evidenced. by phosphates, arsenates, and vanadates as in- dicated by the recent publications of Mooney (1956); Perloff (1956); Shafer, Shafer, and Roy (1956); and others. In the search for a substitute for crystalline quartz, in short supply during World War II, Gruner (1946) reported a ' Signal Corps investigation of synthetic aluminum phosphate. Aluminum phosphate as a mineral was first described by Blomstrand in 1868 and named berlinite (Palache, Berman, Frondel, 1951). Strunz (1941) proved the relatively rare berlinite to be es- sentially aluminum phosphate. A similarity in crystalline structure between aluminum phosphate and o<-quartz was reported by Brill and deBretteville (1948) as part of a Signal Corps project. Stanley (1954) developed a method for growing aluminum phosphate crystals from aqueous solution at near ordinary temperature and pressure. However, the piezoelectric properties of the synthetic crystal were not within the tolerance required and the Signal Corps investigation was abandoned. Interest in the similarity of aluminum phosphate and d-quartz was maintained by geochemists, and Manly (1950) emphasized the close duplication of the physical and chemical properties of the sub- stances. On the basis of this similarity it would seem logical that in an investigation of Goldschmidt’s rules for diadochic replacement of ions, synthetic aluminum phosphate might conveniently be used to provide a replica of the silicon tetrahedral structure so common in the upper lithosphere and trace element substitution made for the aluminum cation. To limit the number of variables the ions selected for substitution need to be trivalent and within the limitation of 15 percent of the size of aluminum. By standardizing the crystallization process, the manner in which the replacement ions enter the structure may be dependent upon their radii, ionization potential, electronegativity, electron configuration, or a combination of these. The number of impurity ions replacing the aluminum cations can be determined by spectro— graphic analysis. The effect of iso-positional substitution of foreign ions on the lattice parameters of the aluminum phosphate crystal may be followed by X-ray diffraction techniques. Measurement of these two variables, when correlated with data pertaining to the particular ion being substituted, bears directly on the Goldschmidt rules and the revisions being proposed by current investigators. CRYSTAL CHEMISTRY CRYSTAL CHEMISTRY Crystal theory Textbooks of geochemistry (e.g., Goldschmidt, 1954) usually list several phosphate minerals as being isotypic1 with silicate min- erals. Among these is the pair aluminum phosphate (AlPO4) and Ot-quartz (SiSiO4 gated this relationship over twenty years ago, and their publications ). Strada (1934) and Huttenlocher (1935) investi- are of considerable interest (see Table l). The quantities measured by X-ray, optical, and analytical balance methods are exact physical properties and show the isotypic qualities of the minerals. Table 2 indicates that the less well-defined characteristics of the two minerals appear even more similar. A compound with such~ extensive similarities to Temperature (°C) Figure 1. Schematic free-energy vs. temperature plot showing stable and metastable equilibria between phases in the system AlP04 (E. C. Shafer and Roy, 1957). A comparison of the physical constants in Table 3 supports the hypothesis that phosphate minerals should exist as isotypic with silica minerals (Hendricks, 1944). Van Wazer (1953) describes the orthophosphates as examples of a simple tetrahedrally bonded phosphorus molecule-ion, with four oxygen atoms arranged tetrahedrally about the phosphorous. Such P0 tetrahedrons may produce chains, rings, or branched polymers 4 by sharing oxygen atoms between tetrahedra. The electronic TAB LE 3 13 A COMPARISON OF PHYSICAL CON STANTS OF SILICON AND PHOSPHORUS Item Silicon PhOSphorus Atomic number ............ 14 15 Atomic weight ............. 28.09 30.98 (phy. units) Common isotopes ........... 27.98577 30.975 28.98566 29.98325 Common valence ........... 4 5 Ionic radius .............. 0.39 A 0.3—0.4 A Electronegativity ........... 1.8 (ev) 2.1 (ev) Ionization potential ......... 11.21 10.22 (ev/unit charge) 14 configuration of phosphorus (152, 232, 2p6, 3s2, 3p3, 3d0) appears to account for its bonding properties. Quantum mechanical treatment indicates the hybrid bond orbitals using _s and p electrons are such that the bonds make tetrahedral angles of 109° (Van Wazer, 1953). The five valence electrons of the atom are in the M or third quantum shell. This shell can accommodate electron pairs in _s, p, and _d_ orbitals; the _s orbitals holding one electron pair, the p orbitals three pairs, and the _d orbitals five pairs. In its normal state phosphorus has two of its valence electrons in the s orbitals with the three remaining electrons in the p orbitals. Thus three more electrons can be acquired by sharing or taking those of other atoms to fill the p orbitals. In representing the orthophosphate ion, Paul- ing (1940, p. 242) used the notation shown in Figure 2 without re- gard to the electron pairs involved. . o .___ I o—P=o I i. .0 . Figure 2. A diagram of the orthophosphate ion without regard for the electron pairs involved. A more conventional configuration of the ion as a resonance hybrid is shown in Figure 3. 15 F o '-3 ' o “-3 F o “-3 " o -3 I II | l O=P-O O-P-O O-P=O 0—P-0 I I I u _ O _ _ 0 _J - O a L O .. Figure 3. The orthophosphate ion depicted as a resonance hybrid. Interpretation of X-ray measurements (Van Wazer, 1953) shows that in the orthophosphates all four P—O bonds average near 1.56 A; whereas, Wells (1950, p. 569) lists the 81—0 bond lengths for 0(- quartz as 1.61 A. The ionic radii of Si (IV) is 0.39 A (Table 3) and Al (III) is 0.57 A (Table 8). The ability of aluminum to proxy for silicon in the silicates is well known. It can be assumed that the tightly bound PO4 tetrahedral unit of structure provides the alu- minum phosphate crystal with quartz-like properties. Aluminum phosphate is not unique as the only mineral re- sembling O(-quartz. Caglioti (1935) and Braspeur (1936) investigated ferric phosphate and noted its close resemblance to quartz while among the recent investigators Dachille and Glasser (1959) report BPO4 as a quartz analogue. However, Westbrook (1958) contributed the interesting fact that aluminum phosphate is the only known com- Pound exhibiting all the isomorphs of silica (see Table 4). 16 TABLE 4 PHYSICAL DATA CONCERNING CRYSTALLINE ORTHO- PHOSPHATE COMPOUNDS DISPLAYING STRUC- TURES SIMILAR TO SILICA Phosphate a (A) _c_ (A) 111““:ng ' Sp. Gr. AIPO4 4.93 5.47 x 2 585 :t 5 2.56 GaPO4 4.92 5.50 x 2 933 :i: 4 3.54 FePO4 5.04 5.62 707 :t 5 3.02 CrPO4 unknowna CoPO4 unknownb vpo4 unknownc MnPO4 4.94 5.48 813 :t 5 3.20 BPO4 4.47 9.926 —— 2.22 aSee M. w. Shafer and Roy (1956). bSee Rankama and Sahama (1950, p. 683). cSee Rankama and Sahama (1950, p. 595). 17 The original study of aluminum phosphate is credited to the German chemist, H. F. Huttenlocher (1935), and it is his work that is cited in reference (Table 1). Huttenlocher was of the opinion that aluminum phosphate existed as an ionic crystal containing Al (III) and P (V) cations in combination with 0 (II) anions. The crystal was indexed on a hexagonal lattice and the 3 axis length reported to be double that of quartz. Renewed interest in aluminum phosphate, created by the Signal Corps synthetic quartz investiga- tion, caused its crystallography to be re-examined (Brill and deBretteville, 1955). The essential conclusions of the recent in- vestigation support Huttenlocher’s contention that Al (III) and P (V) are the cations present. The structure is described as consisting of an alternating sequence of planes which are occupied by aluminum and phosphorus ions, respectively. The planes are perpendicular to the _9 axis and the unit cell contains 3 molecules. Brill and de- Bretteville assume the distance between 0 and Al differs from that between 0 and P, and agree with the concept of ionic bonding. The cell constants are reported by these authors as 4.97 A for a and 10.84 A for 9. The idealized aluminum phosphate crystalline framework shown in Figure 4 is patterned after fl-quartz (Bragg, 1937, p. 84) 18 AlPOh projected on 001 o 'o o Dunc; 5 O o O) O 30 O o lo 0 AlPOh projected on hkO Figure h. .An idealized diagram of’the structural arrangement of the ions in crystalline AlPOh. 19 and represents the combined crystallographic data of Huttenlocher and Brill and deBretteville. The ionic radii of oxygen (1.40 A), aluminum (0.57 A) and phosphorus (0.35 A) are scaled to their re- spective values but the framework is much expanded for illustrative purposes. The tetrahedral groups lie on three layers at different heights and probably the groups spiral as in quartz. The aluminum phosphate crystal class is reported as 32 and very likely the bond- ing shifts the cations from the idealized positions of the diagram to those of less symmetry although no links between tetrahedral groups need be broken and the spiral structure may be retained. The question of the type bonding to be assigned AlPO4 is a difficult one in the absolute sense. The coordinate structure of the analogous silicate tetrahedron is established in the literature. This tetrahedron is explained as due to the dimensions of the Si (IV) and 0 (II) ions. . In an early paper on the structure of complex ionic crystals Pauling (1929) wrote: The crystals considered are to contain only small cations, with relatively large electric charges, i.e., usually trivalent and tetravalent cations, with crystal radii not over 0.8 A. All anions are large (over 1.35 A) and univalent or divalent. The most important anions satisfying this restriction are oxygen and fluorine. Throughout our discussion the crystals will be re- ferred to as composed of ions. This does not signify that the chemical bonds in the crystal are necessarily ionic in the 20 quantum mechanical sense; they should not, however, be of the extreme non-polar or shared electron pair type. The difference in electronegativity is generally accepted as an indi- cation of the amount of ionic bonding between atoms. Wells (1950, p. 554) estimated the character of the Si—O bond as 37 percent ionic. A recent inorganic chemistry reference (Moeller, 1957, p. 723) care- fully points out there are no discrete Si (IV) and 0 (11) ions present in the silica tetrahedron but each Si—O linkage is predominantly ionic in character. Considering the nature of the Al—O and P—O bonds existing in aluminum phosphate, their ionic character may be estimated as high as 50 percent for the former and near 35 percent for the latter. It has been postulated in this discussion that the tetrahedral bonds of the P04 unit are sp3 hybrid bond orbitals and obey the directional characteristics of covalent bonding. Finding the argument on the horns of a dilemma, the author prefers to follow the more recent literature trusting future progress in bond theory will resolve the problem. Investigation of the bonding of AlPO4 by Brill and de- Bretteville (1955) produced evidence offered in support of its ionic nature. By refined X-ray techniques the investigators captured 001 reflections and concluded the reflections from the aluminum and 21 phosphorus layers were cancelled by their ionic nature and the faint line was due to the oxygen layers. On this basis, aluminum phosphate meets the Goldschmidt re- quirement that ionic bonding prevail in minerals to which the rules of diadochic replacement apply and provides a crystalline structure similar to that of o(-quartz. There remain the problems of producing pure aluminum phosphate crystals and introducing impurities under conditions normally found in a chemical laboratory. Crystal growth The following method for producing crystalline aluminum phos- phate by a reaction between sodium aluminate and concentrated ortho- phosphoric acid is described by Huttenlocher (1935). NaAlO2 + 2H3PO4 ——> AlPO4 + NaH2P04 + 2H20 The product was formed by heating the reactants in a closed pipe for several hours at 250°C. In attempting to produce AlPO4 by this method, problems of contamination from the container surfaces, main- tenance of the pressure seal at high temperature and the formation of hydroxides were met with. Other methods were investigated (Hummel, 1949) and the technique used by Stanley (1954) has been 22 selected as most advantageous for this particular study. In his re- port, Stanley states: Good quality aluminum phosphate crystals were produced from a solution of 1.45 moles aluminum phosphate per liter of 6.5 M phosphoric acid when held in a glass lined stainless steel auto- clave for 48 hours at 167°C. A commercial oven controlled by a thermo-regulator to 42°C was available. BIOS Laboratories, 17 West 60th Street, New York 23, New York, supplied precipitated aluminum phosphate. An X—ray powder analysis of this aluminum phosphate was made using a 57.3 mm Buerger camera, Cu target and Ni filter. The film was meas- ured to the nearest 0.1 mm and no attempt made to correct for shrinkage. A comparison (Table 5) with the ASTM card file indi- cated the sample to be AlPO4-2H20. Upon establishing the true molecular nature of the specimen as an aluminum phosphate hydrate (AlPO4-2H20), the decision was made to delay determination of its purity until after the anhydrous form was crystallized. Orthophosphoric acid (H3PO4) meeting American Chemical Society specifications was obtained from J. T. Baker Chemical Company, Phillipsburg, New Jersey. The acid analysis supplied by the manufacturer appears in Table 6. A stainless steel pressure vessel was constructed by the uni- versity machine shop (Figure 5) to contain a platinum crucible of 23 TABLE 5 A COMPARISON OF X-RAY DIFFRACTION DATA FOR COMNIERCIALLY PREPARED ALUMINUM PHOS- PHATE WITH THE STANDARD d-SPACINGS FOR ALUMINUM PHOSPHATE HYDRATE Intensities 51385:???) AligfaIC-lgdisate (1) 2.72 A 2.71 A (2) 1.96 1 -95 (3) 2.22 2.21 approximately 20 ml capacity. 'The pressure vessel lid, secured by 6 machine screws to a flange, was sealed by a lead gasket. The crucible was covered with a platinum lid and no attempt made to confine the reactants inside the crucible. The phosphoric acid (6.5 M) was mixed in 100 ml volumes (73.2 g. of 877° H3PO4 per 100 ml soln) as needed. A 1.5 moi/liter concentration of aluminum phosphate in the acid was desired and to produce this concentration 23.6 grams of AlPO ~2H 0 were mixed 4 2 with each 100 ml of 6.5 M phosphoric acid. Aluminum phosphate 24 TABLE 6 A TABULATION OF THE IMPURITY CONTENT OF ORTHOPHOSPHORIC ACID AS SUPPLIED BY THE MANUFACTURER Lot Number 90782 Percent Assay (H3PO4) ............................. 87.0 Chloride (C1) .............................. 0.0002 Nitrate (N03) .............................. 0.0005 Sulfate (s04) .............................. 0.002 Volatile acids (as HC2H302) ................... 0.001 Alkali and other phosphates .................... 0.05 Arsenic (As) .............................. 0.00003 Heavy metals (as Pb) ........................ 0.0005 Iron (Fe) ................................. 0.003 L DART? LIST NC. DART NAME SIZE RES} 1" _ ' "37555 H: ‘ :45331 'mewmx§” 6 LID STA/NLESS' STEEL f3 052/55 6 H0155 £3.5P5CED ® VESSEL 57A/NLE55 STEEL ”25(749) DRILL —--1 '/0(./90).24 NC" 28 6 HOLES EQUALLV ® FLANGE STA/NLESS STEEL A SPACED Figure 5. An exploded drawing of the stainless 3555:- «'6 i’ES‘F‘E.’ steel pressure vessel used in growing D“ “A av R ~ 575 PM k l_/ _(A I ~ ~ A E 3'." crystalline AlPOh . 26 hydrate is not immediately soluble in the acid and considerable heat is evolved in mixing. If allowed to stand undisturbed for 6 or 8 days the mixture will reach equilibrium, and with the exception of a small amount of unidentified brown precipitate appears similar to pure orthophosphoric acid. The precipitate was avoided in this experiment by drawing off su— pernatant liquid by means of a pipette. Samples of the reactants (15—20 ml) were placed in the platinum crucible, sealed in the pressure vessel, and raised to 163°C. After standing 24 to 48 hours at this temperature, the pressure vessel was rapidly cooled by plunging into cold water and then unsealed. The walls and bottom of the platinum crucible supported the yield of crystalline aluminum phosphate. When separated from the mother liquor, washed, and dried, the crystals appear as in Figures 6 and 7. They are uniformly small, clear, well-formed, and up to 1 mm in length. The scaleno- hedron form is noticed and fully developed rhombohedral faces are the rule. Despite the similarities of aluminum phosphate to cot-quartz, the crystal habits do not bear a close resemblance. A sample of the crystalline material (powdered in an agate mortar) was mounted in a 114.6 mm Buerger camera, placed in an X-ray beam (1.5418 A) and exposed for 6 hours to capture the lines in the back reflection re- gion. The measurements were corrected for film shrinkage and the d-spacings compared (Table 7) with those listed in the ASTM index for crystalline AlPO4. Figure 6. Km enlarged photograph of assorted crystals of «‘l .a. 28 0.25 mm Fizure T. An enlarged photograph of a representative sinrle crystal of 51“ 29 TABLE 7 A COMPARISON OF X-RAY DIFFRACTION DATA FOR THE A1P04 CRYSTALS WITH STANDARD DATA FROM THE ASTM INDEX Intensity ASTM Experimental (100) 3.32 A 3.363 A (100) 1.38 1.387 (80) 4.27 4.275 For the 29 values less than 90° listed in the ASTM index, the film was overexposed and the discrepancy invalues is due to inability to locate the centers of the diffraction lines. Comparison of the weaker lines in the back reflection region with those of the original work (Huttenlocher, 1935) confirm the crystals as A1P04. In order to compare the internal crystal structure of alumi- num phosphate with cat-quartz, samples of each were powdered and subjected to X-ray analysis. The machine used was an X-ray spectrometer (to be described) and a strip-recorder charted the reflections between 29 angles of 20° and 60°. The close 30 resemblance of the internal structures (Figure 8) confirms the work of earlier investigators reporting similarities in physical properties of the two substances. 31 .upuenutxo one ovennaogn annaeuao «o consensus. "echoes“ can no nooaaeaeeo 4 .o oasmwm 2.: 5.. .« mm o. «055 mien O C O bm .3 .5 .3 .o N .o n .«n .3 .3 .o n .0: .9. .cv .3 .9 .0... . 3 .3 .w... .8 .3 . . . . _ . . . . . . . . . . . . . _ . . ‘ g5:§§23§§3::§§. ass. Snsggassxssssssassass xx 0. o. E a 1. my on c... . to .m.w C... 03 CV 01 .. .. /\ss£3555§341(2 - tanO<2 In order to solve for AA and AC the expressions for 0(1 and «2 are combined to produce: x tan 0(2 (cot O(1 + cot 0(2) -x (tan 0C2 coto<2 + 1) AA = 1 - tano<2 cotc>(z ’ AC = tanOC2 - tan<>(1 Since 0(1 and «2 are controlled by hkl values and x is as- sumed to be a constant, AA and AC may be evaluated for any spe- cific set of lines. Each lattice plane represented by a line is as- signed (Eq. 3) a value for 0(1. Holding OS and x fixed, AA and AC are determined by using any other line producing an angle 0(2 with the axis and intersecting this reference line. Repeating the pro- cedure, in turn, for each of the Six reflection peaks provides data used in preparing the graphs required to weight the relative impor- tance of each intersection (Figure 10). 52 413-331, 413—421 x 4 13—404 3! 404—331, 404—421 AA or AC x 10-4 25—413, 225—404 225-421 413-331, 413—421— 14 404-331, 404-421/ 225—331, 225-421/ i 3 4 5 6 7 8 ‘9 Weighting Factor 5.1-1 Figure 10. A diagram of the weighting factors versus variation in axial lengths for different combinations of the quadratic formula representing lattice planes in crystalline AlPO4. 53 The 420 plane fails to fit the graphic scheme for testing the solutions and is omitted from further consideration. The intersec- tion of lines representing 421—331 planes is considered indeterminate and this pair is rejected. Each of the remaining intersections is significant and a coefficient assigned to the importance of its solu- tion. Note that an intersection of two given lines need not provide solutions of equal Significance for both a and g. The magnitudes of the weighting coefficients are listed in Table 14. The X-ray measurements were reproducible and any variation in results appears experimental in nature and primarily due to in- strumental and propagation errors. The diffraction angles (29) are presumed to be correct to 40.01°. The hkl planes, 413—225, were selected to provide numbers tending to create a maximum propaga- tion error when used with the quadratic formula. A value 0.01° above, or below, the true 29 value causes 1/d2 to vary 40.0001 A (NBS Applied Math Series 10, 1950). Solving the quadratic formula (Eq. 1) for 3 produces an error of 40.0004 A for the hkl combina- tion chosen. However, when solved for c this same hkl combination provides the limiting error of 40.0024 A. In the weighting process the final values of _a_ and 3 obtained are considered the best possible and the attached error a maximum. 54 TABLE 14 A TABLE OF WEIGHTING COEFFICIENTS ASSIGNED TO VALUES OF _a_ AND 9 Intersecting “’23::ng wg‘ifltggg hkl Planes (3) (53) 225—404 x 1 x 2.8 413—404 x 1.7 x 15 413—225 x 2.5 x 2.8 413-331 x 3.6 x 1 413—421 x 3.6 x 1 404—331 x 3.8 x 2.4 404—421 x 3.8 x 2.4 225-331 x 4.2 x 4.1 225-421 x 4.2 x 4.1 RE SULT S 55 56 TABLE 15 INHERENT CONTAMINATION OF AlPO4 CRYSTALS DUE TO TRACE ELEMENTS PRESENT IN THE REACTING MATERIALS Trace ppm as Element Metals Gallium 35 Iron 125 Chromium 6 Cobalt < 10 Vanadium < 10 TABLE 16 57 THE RESULTS OF SPECTROSCOPIC ANALYSIS OF AlPO4 CRYSTALS IN WHICH SPECIFIC IMPURITIES WERE INCORPORATED Sample Percent Percent Percent Percent Percent Ga Fe Cr Co V A 2.0 0.16 0.10 0.01 0.01 B 3.0 0.15 -— 0.03 —— C 4.0 0.25 0.13 0.02 0.03 D 5.6 0.23 0.11 0.04 0.035 E 10.6 0.32 0.17 0.05 0.04 F 18.4 0.60 0.18 0.33"‘ 0.05 G —— 0.57 0.19 —— __ H ——- 0.61 -- __ ._ aConsidered in error due to metallic cobalt mixed with sample. M ALPO4 ' I8.4v.6a . 106% Ga ALP04 r 5.S7. Ga ALP04 * 43% Ga ‘ I ALPO“ + 2.07. Ga W‘MW ["W I W «37 I38 I39 I40 MI 142 I43 I44 ALP04 29 133 I34 |45 146 I47 I48 I45 r A‘k‘“Y—‘ SCALE FACTOR 8 1/4- PER MIN Figure 11. A comparison of“ selected diffraction peaks of crystalline AlPO4 containing gallium with those of uncontaminated AlPO4. WW WWW W WWWM NAWWWMWMW ALPqu WFWW WWMW CCCCCCCCCCC AAAAAAA 4* M WWWMWWWW/M . 'I WWW WWW W W MIWMICS‘ C WWWWWWMWWWM M/ “ V WMWWwawW WY” J W WNW mm (1:;- w {I WWW»? l MMWMWW WMWM WWO W WWW WWW AAAAA Figure 15 A comparison of selected diffraction peaks of crystalline AlPO4 containing chromium with those 0 cont inate A P0 . I33 I34 Figure 149 IILPO4 0 03374:. ALP!)4 v 0.057. Cum ALPO4 9 0.04% Co ALI'C4 + 0.03% Cu ' ALPS4 + 0.013;“ ALPO“ + «3.on co A comparison of selected diffra 61 I39 MD MI I42 I43 1" 2 9 SCALE FACTOR 8 ction peaks of crystaIline AIPO4 containing cobalt + W I45 I46 |47 MB “9 ”4' PER MIN with those of uncontaminated AlPO4. I N I. l ADI .IY[.. . .. I f .h‘...’ I J MINI ' MI... MW ALPO4‘ 4' 0.035% V . < ill-Pep 0.037. v ALPOq 9 (3.le V WI I33 MO ‘3‘ I33 Figure 15. A comparison of selected diffraction peaks of crystal ' 1 1 I L . WWW‘WW W \ I42 I43 «44 I45 I46 I47 I49 I49 SCALE FACTOR 3 IM- PER MIN line AlPO4 containing vanadium with thoce of uncontaminated AlPO4. TAB LE 17 -A 63 EXPERIMENTAL 28 VALUES FOR CRYSTALS OF AlPO4 CONTAINING GALLIUM AS AN IMPURITY Percent Ga hkl L L 2 3 4 5.6 10.6 18.4 421 149.75° 149.79° 149.75" 149.72° 149.84° 149.82° 420 146.24 146.29 146.19 146.29 146.23 146.24 331 144.39 144.37 144.37 144.42 144.46 144.51 225 140.95 141.03 140.96 141.01 141.06 141.03 413 135.72 135.75 135.76 135.77 135.75 135.80 404 133.15 133.18 133.20 133.15 133.12 133.12 TABLE 17-B EXPERIMENTAL 28 VALUES FOR CRYSTALS OF AlPO4 CONTAINING IRON AS AN IMPURITY 64 Percent Fe hkl 0.15 0.16 0.23 0.25 421 149.67° 149.65° 149.69° 149.63° 420 146.23 146.22 146.22 146.20 331 144.31 -— 144.30 144.30 225 140.94 140.85 140.85 140.90 413 135.68 135.64 135.67 135.64 404 133.15 133.11 133.13 133.15 Percent Fe hkl 0.32 0.57 0.60 0.61 421 149.69" 149.64° 149.60” 149.56” 420 146.19 146.21 146.12 146.15 331 -— 144.25 144.26 144.18 225 140.99 140.87 140.85 140.81 413 135.67 135.65 135.60 135.55 404 133.18 133.10 133.10 133.05 TABLE 17-C EXPERIMENTAL 26 VALUES FOR CRYSTALS OF AlPO4 CONTAINING CHROMIUM AS AN IMPURITY 65 Percent Cr hkl 0.10 0.11 0.13 0.17 0.18 0.19 421 149.66° 149.63" 149.64° 149.63° 149.64° 149.59° 420 146.16 146.14 -— 146.14 146.13 146.18 331 144.33 144.26 144.27 144.26 144.26 144.31 225 140.88 140.87 140.83 140.87 140.91 140.80 413 135.62 135.68 135.67 135.68 135.67 135.67 404 133.11 133.11 133.13 133.11 133.11 133.09 TABLE 17-D 66 EXPERIMENTAL 28 VALUES FOR CRYSTALS OF AlPO4 CONTAINING COBALT AS AN IMPURITY Percent Co hkl 0.01 0.02 0.03 0.04 0.05 0.33 421 149.73° 149.65° 149.65° 149.67° 149.62° 149.68° 420 144.26 146.13 146.18 146.20 146.17 146.17 331 144.37 144.23 144.30 144.37 144.34 144.32 225 140.39 140.90 140.33 140.39 140.90 140.91 413 135.73 135.67 135.63 135.63 135.63 135.67 404 133.19 133.17 133.10 133.14 133.11 133.17 TABLE 17-E EXPERIMENTAL 28 VALUES FOR CRYSTALS OF AlPO4 CONTAINING VANADIUM AS AN IMPURITY 67 Percent V hkl 0.01 0.03 0.035 0.04 0.05 421 149.67° 149.71° 149.69° 149.720 1 49.67" 420 146.21 146.31 146.28 146.21 146.16 331 144.28 144.38 144.34 144.34 144.40 225 140.92 140.98 140.90 140.98 140.92 413 135.64 135.68 135.70 135.72 135.67 404 133.13 133.17 133.17 133.20 133.14 TABLE 18-A 68 UNWEIGHTED _a_ AND 9/2 VALUES (IN A) FOR CRYSTALS OF AlPO4 CONTAINING GALLIUM Percent Ga hkl’s 2 3 4 5.6 10.6 18.4 421—225 4.9280 4.9278 4.9280 4.9286 4.9271 4.9275 5.4608 5.4584 5.4605 5.4583 5.4583 5.4586 331—225 4.9426 4.9415 4.9414 4.9407 4.9401 4.9394 5.4481 5.4466 5.4489 5.4479 5.4470 5.4483 413—225 4.9516 4.9515 4.9505 4.9508 4.9517 4.9502 5.4400 5.4385 5.4417 5.4393 5.4369 5.4385 404—225 4.9045 4.9072 4.9033 4.9082 4.9111 4.9104 5.4813 5.4763 5.4821 5.4761 5.4722 5.4735 421-404 4.9285 4.9282 4.9285 4.9290 4.9275 4.9278 5.4393 5.4393 5.4393 5.4312 5.4312 5.4393 404-331 4.9094 4.9083 4.9083 4.9073 4.9068 4.9061 5.4723 5.4744 5.4728 5.4776 5.4801 5.4814 404—413 4.9702 4.9691 4.9693 4.9677 4.9678 4.9659 5.3653 5 .3671 5.3654 5.3709 5.3722 5.3754 331—413 4.9062 4.9050 4.9050 4.9042 4.9036 4.9031 5.6407 5.6433 5.6424 5.6451 5.6505 5.6470 421—413 4.9264 4.9260 4.9264 4.9270 4.9255 4.9258 5.5479 5.5470 5.5445 5.5410 5.5496 5.5427 TABLE 18-B UNWEIGHTED a AND _c_/2 VALUES (IN A) FOR CRYSTALS OF AlPO4 CONTAINING IRON Percent Fe hkl’S 0.015 0.016 0.23 0.25 421—225 4.9293 4.9292 4.9294 4.9287 5.4626 5.4600 5.4670 5.4631 331—225 -— 4.9422 4.9424 4.9423 -—— 5.4489 5.4558 5.4514 413—225 4.9529 4.9526 4.9535 4.9521 5.4417 5.4401 5.4393 5.4425 404—225 4.9051 4.9057 4.9047 4.9041 5.4838 5.4806 5.4825 5.4847 421—404 4.9298 4.9296 4.9300 4.9293 5.4312 5.4393 5.4312 5.4393 404—331 -——~ 4.9091 4.9093 4.9092 -— 5.4744 5.4739 5.4753 404-413 4.9718 4.9711 4.9726 4.9711 5.3659 5.3652 5.3627 5.3662 331-413 -—- 4.9058 4.9058 4.9060 —— 5.4489 5.6505 5.4514 421—413 4.9276 4.9276 4.9278 4.9272 5.5504 5.5470 5.5496 5.5496 TABLE 18-B (Continued) 7O Percent Fe hkl’s 0.32 0.57 0.60 0.61 421-225 4.9289 4.9299 4.9294 4.9303 5.4587 5.4621 5.4618 5.4630 331-225 -— 4.9428 4.9430 4.9440 — 5.4509 5.4501 5.4511 413—225 4.9532 4.9542 4.9527 4.9549 5.4385 5.4417 5.4417 5.4417 404—225 4.9058 4.9056 4.9063 4.9074 5.4789 5.4834 5.4821 5.4831 421-404 4.9293 4.9305 4.9300 4.9308 5.4312 5.4312 5.4312 5.4393 404—331 — 4.9097 4.9099 4.9108 — 5.4759 5.4755 5.4769 404-413 4.9721 4.9735 4.9712 4.9737 5.3621 5.3636 5.3674 5.3661 331-413 — 4.9064 4.9066 4.9077 —— 5.6532 5.6451 5.6505 421—413 4.9272 4.9281 4.9278 4.9287 5.5496 5.5530 5.5487 5.5538 TABLE 18-C 71 UNWEIGHTED E AND 3/2 VALUES (IN A) FOR CRYSTALS OF AlPO4 CONTAINING CHRONIIUM Percent Cr hkl’s 0.10 0.11 0.13 0.17 0.18 0.19 421—225 4.9292 4.9294 4.9295 4.9294 4.9301 4.9300 5.4611 5.4619 5.4631 5.4606 5.4609 5.4636 331—225 4.9419 4.9428 4.9427 4.9429 4.9429 4.9422 5.4508 5.4502 5.4516 5.4489 5.4499 5.4531 413—225 4.9536 4.9519 4.9518 4.9526 4.9544 4.9515 5.4409 5.4425 5.4441 5.4409 5.4409 5.4457 404—225 4.9061 4.9057 4.9034 4.9071 4.9056 4.9045 5.4820 5.4828 5.4861 5.4801 5.4823 5.4860 421—404 4.9296 4.9300 4.9300 4.9300 4.9306 4.9306 5.4393 5.4312 5.4312 5.4312 5.4312 5.4312 404-331 4.9087 4.9097 4.9096 4.9097 4.9097 4.9091 5.4771 5.4753 5.4742 5.4753 5.4748 5.4775 404—413 4.9725 4.9703 4.9711 4.9706 4.9739 4.9702 5.3647 5.3684 5.3662 5.3679 5.3619 5.3696 331—413 4.9052 4.9066 4.9064 4.9066 4.9064 4.9058 5.6550 5.6424 5.6450 5.6433 5.6532 5.6478 421—413 4.9274 4.9279 4.9279 4.9279 4.9283 4.9285 5.5530 5.5453 5.5462 5.5462 5.5521 5.5436 TABLE 18-D 72 U'NWEIGHTED a AND c/2 VALUES (IN A) FOR CRYSTALS OF AIPOZ CONTAINING COBALT Percent Co hkl’s 0.01 0.02 0.03 0.04 0.05 0.33 225-421 4.9283 4.9293 4.9293 4.9292 4.9296 4.9290 5.4621 5.4613 5.4616 5.4614 5.4610 5.4609 225-331 4.9414 4.9428 4.9424 4.9414 4.9417 4.9420 5.4509 5.4496 - 5.4503 5.4509 5.4506 5.4497 225-413 4.9288 4.9298 4.9298 4.9296 4.9301 4.9295 5.4425 5.4417 5.4409 5.4400 5.4400 5 .4409 404-225 4.9083 4.9096 4.9092 4.9083 4.9086 4.9090 5.4850 5.4842 5.4815 5.4828 5.4813 5.4830 404-421 4.9288 4.9298 4.9298 4.9296 4.9301 4.9295 5.4393 5.4312 5.4312 5.4393 5.4393 5.4312 404-331 4.9083 4.9096 4.9092 4.9083 4.9086 4.9090 5.4738 5.4724 5.4769 5.4763 5.4773 5.4736 404—413 4.9701 4.9719 4.9719 4.9728 4.9722 4.9719 5.3649 5.3628 5.3663 5.3627 5.3653 5.3628 331—413 4.9050 4.9064 4.9058 4.9048 4.9052 5.9056 5.6451 5.6451 5.6515 5.6560 5.6541 5.5487 421-413 4.9268 4.9278 4.9276 4.9274 4.9279 4.9274 5.5453 5.5470 5.5513 5.5521 5.5496 5.5487 TABLE 18-E 73 UNWEIGHTED 3 AND 9/2 VALUES (IN A) FOR CRYSTALS OF AlPO4 CONTAINING VANADIUM Percent V hkl’s 0.01 0.03 0.035 0.04 0.05 225-421 4.9291 4.9287 4.9288 4.9285 4.9291 5.4608 5.4592 5.4618 5.4593 5.4608 225431 4.9427 4.9412 4.9417 4.9418 4.9409 5.4490 5.4484 5.4506 5.4479 5.4507 225—413 4.9535 4.9529 4.9515 4.9518 4.9526 5.4393 5.4385 5.4425 5.4393 5.4409 404—225 4.9061 4.9059 4.9033 4.9039 4.9055 5.4810 5.4792 5.4842 5.4810 5.4815 404—421 4.9296 4.9291 4.9293 4.9290 4.9296 5.4393 5.4393 5.4393 5.4312 5.4393 404—331 4.9096 4.9081 4.9086 4.9087 4.9078 5.4745 5.4753 5.4743 5.4721 5.4774 404-413 4.9722 4.9715 4.9708 4.9709 4.9713 5 .3643 5 .3636 5 .3647 5 .3627 5 .3654 331—413 4.9062 4.9048 4.9052 4.9054 4.9044 5.6487 5.6515 5.6470 5.6442 5.6541 421—413 4.9274 4.9270 4.9272 4.9270 4.9276 5.5513 5.5496 5.5462 5.5453 5.5479 ..... TABLE 19 THE BEST EXPERIMENTAL E AND 3/2 VALUES FOR AlPO4 CRYSTALS CONTAINING GALLIUM, IRON, CHROMIUM, COBALT, AND VANADIUMa {55532; _a_ (A) 3/2 (A) Gallium 0.0 4.9292 5.4633 2.0 4.9285 5.4629 3.0 4.9282 5.4615 4.0 4.9278 5.4633 5.6 4.9278 5.4615 10.6 4.9271 5.4615 18.4 4.9268 5.4629 11313 0.0 4.9292 5.4633 0.15 4.9354 5.4536 0.16 4.9289 5.4543 0.23 4.9292 5.4652 0.25 4.9289 5.4566 0.32 4.9352 5.4505 0.57 4.9299 5.4642 0.60 4.9296 5.4638 0.61 4.9306 5.4651 aError considered maximum for each value in column 2 is 10.0004; error considered maximum for each value in column 3 is 3:0.0024. TABLE 19 (Continued) Percent Impurity _a_ (A) _C_/ 2 (A) Chromium 0.0 4.9292 5.4633 0.10 4.9289 5.4651 0.11 4.9292 5.4624 0.13 4.9292 5.4647 0.17 4.9296 5.4624 0.18 4.9299 5.4633 0.19 4.9292 5.4661 Cobalt 0.0 4.9292 5.4633 0.01 4.9278 5.4647 0.02 4.9292 5.4629 0.03 4.9292 5.4638 0.04 4.9289 5.4647 0.05 4.9292 5.4647 0.33 4.9289 5.4629 Vanadium 0.0 4.9292 5.4633 0.01 4.9292 5.4638 0.03 4.9257 5.4629 0.035 4.9285 5.4647 0.04 4.9285 5.4615 0.05 4.9285 5.4647 _v--——L DISCUSSION OF RESULTS AND CONCLUSIONS 76 DISCUSSION OF RESULTS AND CONCLUSIONS Accgptance of impurities The spectroscopic data (Table 16) indicate gallium entered the crystalline structure of aluminum phosphate in amounts near those predicted by stoichiometric calculation, whereas iron, chro- mium, cobalt, and vanadium were not accepted as anticipated. The iron and chromium ions entered the structure sufficiently to alter the color but the inclusion of less than 1.0 percent places them in a category different from gallium. The induced coloration is weak and Zinn (1959) reported no significant change in refractive index with addition of chromium. Cobalt and vanadium ions were the least acceptable with a range of substitution from 0.0 to 0.05 percent. It is possible that the cobalt and vanadium impurities were not incor- porated in the crystal structure but were absorbed on the surface or carried along as contamination. The iron and chromium ions were incorporated in the crys- tallizing structure in amounts ranging approximately 0.1 percent for chromium and 0.5 percent for iron, 3 factor of ten more than cobalt and vanadium. The number of foreign ions taken up increased with the amount of impurity added to the crystallizing solution and the 77 78 crystal appearance altered to opaque or colored. Spectroscopic analysis of the pure crystalline aluminum phosphate (Table 15) indi- cates the inherent trace element concentrations are 100-fold less than the amount Of added impurity and may be considered negligible. The ease with which the gallium ion entered the structure leads to the question whether it prefers to substitute diadochically for aluminum or to form mixed crystals. There is no indication that the upper limit of gallium substitution reached in this experiment could not be exceeded by increasing the gallium content of the crys- tallizing solution. Variation of a and c with impurities The replacement of aluminum ions in the crystalline structure by ions Of different, but similar, physical properties altered the lattice dimensions as anticipated. Both the amount and kind of im- purity substituted affects the _a_ and _g axial lengths (Figures 16 A—E). Gallium ions decreased the a dimension in a regular way correspond- ing to increased gallium substitution. Change in the g dimension ac- companying the decrease in a is either masked by the magnitude of the propagated error or does not exist. The substitution of iron for aluminum created the greatest structural distortion. Both _a_ and _c are Observed to vary in a 79 L ........... 4.9295 - It I. error :1: 0.0004 A 4.9285 «- € 57 “I 4.9275 . 4.9265 1 ‘ . I 5 10 15 20 Percent Ga 5.4655 fun-”"n-“nn'ql 5.4645 - 62‘ 7; 5'4635 ‘ e 0 0024 A , a: x rror :i: . A \,. gl 5.4625 .- 5.4615 -‘ u ‘4 1k ..................... Y 5.4605 1 ' ‘ ‘ 5 10 15 20 Percent Ga Figure 16-A. A diagram of axial length versus percent impurity for AlPO4 crystals containing gallium. 80 4.9340 « 4.9320 - a axis (A) 4.9300 .. I ------ -- ”\I x/\x ierror 4 0.0004 A 4. 9280 .10 .20 .30 .40 .50 .60 .70 Percent Fe 5.4675 I .---- ---------------- 5'46“: error 4 0.0024 A ”‘5 A) 5.4605 -" """“““ " "" c/2 axis ( 5.4570 ~ 5.4535 - \ 5.4500 . , “ . .20 30 40 .15 .5'0 .6'0 .70 Percent Fe Figure 16-B. A diagram Of axial length versus percent impurity for AlPO4 crystals containing iron. (3) 4.92977 4.9293- 81 error :1: 0.0004 A g 1 \ /x 3‘ x (8| \ 4.9239. L x 4.9285 , , . , L1 .04 .08 .12 .16 .20 Percent Cr 5.4660- " ”"1 5.4650. " A X ‘5, § 5.4640. :3 H/ error :I: 0.0024 A x BI 5.4630- / X 5.4620- 5.4610 X U V .04 .03 .1'2 .16 .20 Percent Cr Figure 16-C. A diagram of axial length versus percent impurity for AlPO4 crystals containing chromium. 82 ‘61..“ ‘m—u gusto \ In.) an...._-‘1 I . 4.9299. 0:5 error 4: 0.004 A V 4.92913 \ " 3 ,/ N “I 4.9233- 4.9275 , , , , ' .01 .02 .03 .04 .05 Percent Co 5.4660I --------------7I~ 5.46504 X X X a / ":1; 5.4640- .. N 5 4630:. ,/ EI ' 3 error 4. 0.0024 A 5.4620“ 5.4610 V .0'1 .0'2 .63 .04 .05 Percent CO Figure 16-D. A diagram of axial length versus percent im— purity for AlPO4 crystals containing cobalt. 83 4.9305. 4. 9295 f“""““’“ I.——.. error :I: 0.0004 A °$ 4.9235. """" --- ,.__.,.____,. 5‘3 “" 4.9273 4.9265« 4.9255 . ' f U ' .01 .02 .03 .04 .05 Percent V 5.4681. “‘1: 5.4657~ ........... 7 m I E . N bl /‘\ 5.4633 I 'I \x error :i: 0.002 5.4609 , Y , , , .01 .02 .(73 .04 .05 Percent V Figure 16-E. A diagram of axial length versus percent im- purity for AlPO4 crystals containing vanadium. a: s9 '. mm nun-rat I fl s-_" ‘71: - _‘6 A ‘3dfl‘£-‘ . . A 84 relative, but opposite, manner. The changes in axial length do not appear to follow directly the increase in impurity content of the crystal. It may be significant that the crystalline structure accom- modated the iron impurity at certain concentrations with comparatively large fluctuations of axial lengths while remaining relatively un- changed for intermediate concentrations. The chromium ion substituting in amounts Of the Same order as iron increases the 3 axis length sufficiently to exceed the limiting error (Figure 16-C). The pattern follows the one produced by iron although in this case the c axis variation is not sufficient to escape the error attached to the measurement. The cobalt and vanadium series present a significant varia- tion in the 3 axis length with added impurity (Figures 16 D—E). Both ions cause a decrease in a for impurity amounts of the order of 0.01 percent. In the vanadium series the alteration is more pro- nounced and the a axial length remains contracted. Lack of data outside the limits Of experimental error prevent 2 axis measurements from contributing to the Observations. Reasons for acceptance of impurities It is generally held that simple cations enter a crystalline structure from a melt as a result Of changes in temperature, valence, 85 and ionic size. In this investigation the temperature and valence were selected as experimental constants and the ionic radii allowed to vary. The original Goldschmidt (1926) postulation of crystals being built up by the packing Of spheres with the arrangement of ions in the crystal determined by the radius ratio (cation radius/ anion radius) has proved suitable for many simple inorganic com- pounds, however, there is a lack Of agreement on the Sizes of ions. The large amount of data on crystals yields internuclear distances and the problem of how to divide this distance into cationic and anionic components in any given crystal is inescapable when at- tempting tO determine ionic radii values. The earliest acceptable system of ionic radii was developed by Wasastjerna (1923), who based his data on 1.32 A for 0= radii and 1.33 A for F- radii. Goldschmidt’s modification of these early figures was founded on 1.40 A for O: and 1.36 A for F- and based upon extensive labora- tory data (Goldschmidt, 1929). Pauling (1927) developed a different set of values for ionic radii by a wave mechanical treatment of the factors involved which has proved extremely valuable. A late paper by Stern and Amis (1959) discusses the attempts of various investi- gators to establish more exactly the values of atomic and ionic radii. At the present time the difficulties of the problem provide a variation of 20—30 percent and have prevented these authors from 86 selecting any particular values as standard. The Goldschmidt radii are as acceptable to most workers as those of Pauling, Zachariasen (1931), Ahrens (1952), and others. Table 8 lists the empirical Gold- schmidt radii values for aluminum and the other trivalent cations used. On the basis of similar ionic radii, gallium should be most successful in substituting for aluminum in the crystal framework, as is the case. Iron has the largest ionic radius of any of the cations presented for substitution and Should be least successful in substi- tuting for aluminum. This is not true as iron ranks second to gal- lium in the ability to substitute for aluminum and creates the great- est structural alteration Of any Of the five trace ions. Electronegativity is an ionic property considered to be a principle factor (Fyfe, 1951) in determining the structure of crystals. Electronegativity is a combination of ionization potential and electron affinity and extensive calculations were performed by Pauling and Haissinsky (Fyfe, 1951) to provide the coefficients listed in Table .20. As with ionic radii, recent work casts doubt on the accuracy of these calculations. Fineman and Daignault (1959) question Paul- ing’s relationships for bonds with a great deal of ionic character and present coefficients 0.6 to 0.7 higher than those Of Pauling and Haissinsky. Ringwood (1955) attributed a major significance to the 87 TABLE 20 ELECTRONEGATIVITY VALUES ASSIGNED THE CATIONS LINKING THE P04 TETRAHEDRA IN THE CRYSTAL STRUCTURE OF AlPO4 Element 115 £33261 Aluminum 1 .5 Gallium 1 . 6 Iron 1 .8 Chromium 1 . 6 Cobalt 1 .7 Vanadium 1 .35 Li‘s, .- “mute-v.1- .. I“ 88 role played by electronegativity in elements which display diadochy. He is of the opinion an element displaying higher electronegativity than the host forms a weaker bond and will not be accepted readily while an element of lower electronegativity than the host will be preferentially incorporated. Such a relationship is reported to ap- ply when the electronegativity difference between the host and the trace element is greater than 0.1 (approximately). Using this criteria, vanadium should be accepted by the crys- talline aluminum phosphate in the largest concentration, and in turn provide the greatest change in lattice parameters while iron will be least acceptable and provide the smallest alteration in the crystal structure. Experimental evidence indicates this is not true. Ionization potential in its own right may be considered to play a fundamental part in the ability of an ion to substitute in a structure. Ionization potential is a measure Of the energy required to move a single electron from a free atom or ion to infinity, or conversely, as a measure of the energy lost by a recombination of the resulting ion and the electron. Ahrens and Morris (1956) em- ployed the ionization potential of the cation as an approximate com- parative measure of its power of attraction for the most loosely bound electrons of the neighboring anions when considering the 89 formation of crystals from different cations of the same charge and ionic radii with a given anionic constituent. According to their ar- gument, the greater the ionization potential of the cation, the greater the deviation from ionic bond character, the weaker the bond, and the less desirable the cation to the structure. Recent values for ioniza- tion potentials assembled by Basolo, Pearson, and others (1958) are listed in Table 21. Following the ideas presented by Ahrens and Morris, vanadium Should be preferred by the crystalline structure over the gallium due to the vanadium’s lower ionization potential. This is not found ex- perimentally to be true. The chemical nature of the trace ions entering a crystalline structure affects the bonding and is necessarily a prime factor in establishing any substitutional pattern. The electron configurations of the metallic elements involved in this study are listed in Table 22 (Rankama and Sahama, 1950, p. 796). It is evident that the 8-electron outer shell Of trivalent aluminum has reached a degree Of stability. The 18-electron cation of gallium also contains the maximum number of electrons in its M shell. This similarity plus the physical prop- erties previously listed permit gallium to proxy for aluminum in a predictable manner. X-ray evidence shows a reduction in the _a_ axis. length with addition Of gallium (Figure 16-A), and, as the amount of 90 TABLE 21 IONIzATION POTENTIALS ASSIGNED THE CATIONS LINKING THE P04 TETRAHEDRA IN THE CRYSTAL STRUCTURE OF AlPO4 Total Ionization Element Potential (at 0°K) Al (III) 1228 kcal Ga (111) 1310 Fe (III) 1261 Cr (III) 1259 CO (III) 1365 V (III) 1094 TABLE 22 91 ELECTRON CONFIGURATIONS ASSIGNED THE CATIONS LINKING THE P04 TETRAHEDRA IN THE CRYSTAL STRUCTURE OF AlPO4 Ele- K L M N ment (1s) (2s) (2p) (as) (33) (3d) (4s) (4p) (4d) (40 Al 2 2 6 2 1 Ga 2 2 6 2 6 10 2 1 Fe 2 2 6 2 6 6 2 Cr 2 2 6 2 6 5 1 Co 2 2 6 2 6 7 2 V 2 2 6 2 6 3 2 92 substituted gallium increases, the axial contraction is more pro- nounced. Pauling (1929) noticed this type Of deformation and wrote: Deforming action of an eighteen-shell ion is much greater than the deforming action of an eight-shell ion of the same size and valence. This is due to larger effective nuclear charge. Iron, chromium, cobalt, and vanadium are transition metals and exist with their M Shells containing less than a stable 18-electron con- figuration. AS these atoms oxidize to the trivalent state it is as- sumed the N shell loses its electrons first and then the 5 d-Orbitals of the M shell may alter. The exact manner Of alteration and any simple prediction for the relative stability Of a resulting ion is a matter Of conjecture (Gould, 1955, p. 21). The iron and chromium originally were put in the crystalliz- ing solution as hydrated ferric and chromic phosphate. Spectro- scopic evidence indicates only a few tenths of a percent Of either the Fe (III) or the Cr (III) ions succeeded in proxying for aluminum. Keeping in mind the specific crystallizing temperature, pressure, and concentration, this failure to substitute may be due to either the in- ability to become free, mobile ions or to the energy demands of the lattice Sites. The latter possibility appears the more feasible. The insolubility of the cobaltic oxide in addition to its elec- tronic configuration may be a prime factor for its lack of success in substitution. From the appearance of the reaction, the cobalt 93 metal added to enhance the cation concentration formed the cobaltous ion and was not suited (smaller charge, larger size) to the diadochic process. The vanadium metal was soluble and the presence Of large numbers Of V (111) ions was indicated by the intensely green crystal- lizing solution. However, to the same degree as cobalt these cations appeared unable to occupy the structural aluminum sites, and spectro- scopic analysis showed less than 0.05 percent concentration of either ion under conditions which allowed gallium to substitute up to 18 per- cent. The presence Of transition metals in any compound character- istically produces a strong coloration, and Pitzer and Hildebrand (1941) point out that if the color of a compound differs from that of its constituent ions in solution it may be an index of deviation from pure ionic character. The AlPO4 crystals containing cobalt and va- nadium Show only faint shades of color——blue, in the case of cobalt, and, surprisingly, a tinge of lavender for the vandadium impurity. The hydrated Co (11) ion is known to be pink, and the anhydrous form is blue. The V (IV) ion is known to be blue, while the highly unstable V (II) ion is violet. It is probably the V (II) ion entering the structure instead of the common V (111) ion which existed in the green mother-liquor. The trivalent vanadium ion is recognized to be a comparatively strong reducing agent (Rankama and Sahama, 1950, 94 p. 596). However, this unexpected substitution of V (II) is postulated on the evidence of a faint tinge of color, and until more concrete data are Obtained it remains a highly conjectural phenomenon. Inability to establish ionic bonds could be a reason for iron, chromium, cobalt, and vanadium affecting the crystalline structure differently from gallium. Such a possibility leads again to the prob- lems of assessing the amount of ionic character Of a particular bond, defining the meaning of bonds intermediate between ionic and coval- ent, and justifying the concept Of orbital hybridization. It serves little purpose to attempt to answer a question as to the scheme of iso-positional substitution in aluminum phosphate by posing new prob- lems based on conflicting theories although the ultimate answer may lie contained therein. As previously stated, iron was the most successful of the transition metals in proxying for aluminum and provided a unique change in the structural pattern of the crystal. Whereas the gal- lium impurity contracted the _a_ axis by 0.003 A at the maximum sub- stitution of 18 percent, iron expanded the 3 axis by 0.006 A with the minimum substitution of 0.15 percent. This same percentage Of iron decreased the 9 axis a measurable amount. It is significant that (1) iron altered the axis in a manner Opposite from gallium, (2) caused a relatively large axial change with trace substitution, 95 (3) shifted the axial lengths for discrete percentages of impurity, and (4) allowed the lattice to remain relatively unchanged for inter- mediate values. One explanation for this may be a combination of electron configuration and ionic size. The 18-electron gallium ion when proxying for the 8-electron aluminum exerts a stronger Coulomb force on its nearest neighbor oxygens. The net result is a partial collapse of the structure as evidenced by a shrinkage of the 2 axis. The similarity in ionic size of gallium to aluminum permits fourfold coordination and fails to create a space demand. An iron cation in the place of the aluminum in the structure will exert a different at- traction on the surrounding oxygens. The electron structure of this transition metal may react to the electrostatic field by altering the valence configuration. If an electron should be shifted it would polarize in a way to minimize the Coulomb energy of the local sys- tem. An electron more strongly concentrated around one of the Fe-O bonds not only would satisfy the minimum energy requirement but might be responsible for a dimensional change in the structure. The iron ion is assigned an ionic radius near 15 percent larger than aluminum, which suggests sixfold coordination and enhances the space demands in the structure. Rankama and Sahama (1950, p. 659) discuss the diadochy of Fe (III) and Al (III) and observe that 96 considerable substitution comes about when the aluminum occurs as a cation outside the silicon-aluminum network in the aluminum sili- cates. This replacement does not occur when the aluminum replaces silicon within the oxygen tetrahedra of the aluminosilicates. Chromic compounds resemble ferric substances, although the Cr (III) ion is known to be less strongly oxidizing than the Fe (111) ion (Moeller, 1957, p. 880). In this study the chromium entered the framework of AlPO4 and altered the lattice parameters in a weak but similar manner to that of iron. The electron configurations of these ions and their abilities to establish crystalline bonds appear to be the major factors in their success in proxying for Al (III) in the crystal. The failure of cobalt and vanadium to substitute for aluminum is postulated as due to their chemical nature. Trivalent vanadium has been described as an unstable ion in need of additional electrons. Trivalent cobalt is reported always to have a coordination number of 6 (Rankama and Sahama, 1950, p. 683), and Moeller (1957, p. 203) of- fers it as an example of an ion having six bonds directed toward the corners of a regular octahedron. The present investigation indicates these two ions fail to enter the crystalline framework in more than trace amounts. Again it seems the “size” of the ion is a function 97 of its environment and the chemical nature of the ion is the funda- mental factor in establishing a substitutional pattern. Welcher and Hahn (1958, p. 25) point out the relative stability of the ionic state is often best correlated with nuclear charge and size of the atom, but the electronic structure determines primarily what oxidation state a given element will exhibit. In keeping with the question of bonding between ions, Corwin (1958), in studying reactions under controlled conditions between salts of alkaline earth metals and silica, found that the rate of reaction and the formation of crystalline structure are correlated with the initial pH of the solution and the relative dimensions of the ions involved. Since the X-ray diffraction peaks studied in this investigation retain their identity with added impurity, it is assumed that the for- eign ions are substituting for aluminum and no new crystal structure is being formed. Concentrations of gallium approaching 18 percent in the aluminum phosphate crystal tend to split the diffraction peaks and create the impression that a super lattice of gallium phosphate may be forming. The crystals bearing iron near the maximum of 0.60 percent evidence some broadening of the diffraction peaks, but the data are considered insufficient to support any contention of an ordered substitutional pattern. 98 In summary, it does not yet appear possible to distinguish clearly between the space occupied by an atom within a crystalline structure and the influence that it exerts. Until the present uncer- tainty concerning atomic and molecular constants, accepted as known a decade ago, is reduced, a concise explanation of structural alter- ation in crystalline substance caused by iso-positional substitution of foreign ions is impossible. Conclusions 1. The tetrahedral structures A104 and P04 are replicas of the SiO4 (Si 8104), may be produced under ordinary laboratory conditions. tetrahedral unit, and crystalline AlPO , similar to 8102 2. Impurities of Ga, Fe, Cr, Co, and V may be introduced into crystallizing AlPO4 without destroying its gross morphology or lattice structure. With the exception of gallium, less than 1.0 per- cent impurity enters the crystal. 3. The principal influence on the ability of an impurity to substitute iso-positionally in a given structure (temperature and charge constant) is not the similarity of ionic radii. 4. The electron configuration of an ion attempting to proxy for another in a given structure is the prime factor in its occupying a given structural site. SUGGESTIONS FOR FURTHER WORK 99 SUGGESTIONS FOR FURTHER WORK 1. Continue increased substitution of gallium into the struc— ture with the purpose of obtaining data concerning the binary system AlPO4 -GaPO4. 2. Retrace the substitutional pattern of the iron impurity with emphasis on concentrations above 0.61 percent. 3. Maintaining temperature, concentration, and radii constant, investigate the substitutional behavior of ions with different charge, e.g., Cr (III) radius 0.64 A and Ti (IV) radius 0.64 A. 4. Holding the experimental conditions constant, add equal numbers of two or more foreign ions to the crystallizing solution and determine the correlation of ionic radii, ionic potential, electron configuration, and other parameters of the ions with their success in competing for structural sites. 5. Determine variation in physical properties of the alumi- num phosphate crystal (e.g., directional hardness) with the kind and. amount of foreign ions substituted. 6. Investigate the effect of the pH of the reacting solution on the crystalline product. 100 BIBLIOGRAPHY 101 BIBLIOGRAPHY Ahrens, L. H., “The Use of Ionization Potentials. Part 1: Ionic Radii of Elements,” Geochimi. et Cosmochimi. Acta, 2:155 (1952). Ahrens, L. H., and Morris, D. F. C., “Ionization Potentials and the Chemical Binding and Structure of Simple Inorganic Crystal- line Compounds I—II,” Jour. Inorg. Chem., 3:263 (1956). Azaroff, L. V., and Buerger, M. J., The Powder Method, McGraw- Hill, New York, New York, 1958, pp. 342. Basolo, F., and Pearson, R. G., Mechanisms of Inorganic Reactions, John Wiley & Sons, New York, New York, 1958, pp. 426. Bowen, N. L., “The Reaction Principle in Petrogensis,” Jour. Geology, 30:177 (1922). Bragg, W. L., “The Structure of Silicates,” Zeit. Krist., 74:23? (1930). Bragg, W. L., Atomic Structure of Minerals, Cornell Univ. Press, Ithaca, New York, 1937, pp. 292. Braspeur, Paul, “X-ray Examination of Anhydrous Ferric Phos- phate,” Compt. Rend., 202:761 (1936). Brill, R., and deBretteville, A., “On the Crystal Structure of AlPO4,” Am. Mineralogist, 33:750 (1948). Brill, R., and deBretteville, A., “On the Chemical Bond Type in AlPO4,” Acta Crystallographica, 8:567 (1955). Caglioti, V., “Sulla Struttura Del Fosfato Ferrico,” Accademia dei Lincei, Atti, 22:146 (1935). 102 103 Clarke, W., and Washington, H. S., “The Average Chemical Compo- sition of Igneous Rocks,” Proc. Natl. Acad. Sci. U.S., 8:108 (1922). Clarke, W., and Washington, H. S., “The Composition of the Earth’s Crust,” U.S. Geol. Survey, Profess. Paper 127 (1924), pp. 117. Corwin, J. F., “Hydrothermal Reactions under Super Critical Con- ditions,” Jour. Phys. Chemistry, 62:1086 (1958). Dachille, F., and Glasser, L. S. D., “High Pressure Forms of BPO4 and BAsO4; Quartz Analogues,” Acta Crystallograph- i_<_:_a, 12:820 (1959). Evans, R. C., Crystal Chemistry, Cambridge Univ. Press, Cambridge, England, 1952, pp. 388. Fineman, M. A., and Daignault, R., “Electronegativities Calculated Using a Modification of Pauling’s Equation,” Jour. Inorg. & Nucl. Chem., 10:205 (1959). Ford, W. E., Dana’s Textbook of Mineralogy, 4th. Ed., John Wiley & Sons, New York, New York, 1954, pp. 851. Fyfe, W. S, “Isomorphism and Bond Type,” Am. Mineralogist, 36:538 (1951). Goldschmidt, V. M., “Geochemische Verteilungsgesetze der Elemente, VIII,” Skrifter Norske Videnskaps-Akad. Oslo, 1, Mat-Naturv. .19." 1926, pp. 156. Goldschmidt, V. M., “Crystal Structure and Chemical Composition,” Trans. Faraday Soc., 25:253 (1929). Goldschmidt, V. M., “Kristallchemie,” Fortschritte der Mineralogie, Kristallographie und Petrographie, 15:974 (1931). Goldschmidt, V. M., “Kristallographie und Stereochemie Anorganischer Verbindugen,” Stereochemie, verlag von Frank Deuticke in Leipzig, 1932 (reprint). 104 Goldschmidt, V.M., “The Principles of Distribution of Chemical Ele- ments in Minerals and Rocks,” Jour. Chem. Soc. London, p. 655 (1937). Goldschmidt, V. M., Geochemistry, Clarendon Press, Oxford, England, 1954, pp. 730. Goldsmith, J., “Gallium and Germanium Substitution in Synthetic Feldspars,” Jour. Geology, 58:518 (1950). Gould, E. 8., Inorganic Reactions and Structure, Henry Holt 8: Co., New York, New York, 1955, pp. 470. Gruner, J. W., “Isostructural Relationship of AlPO4 and 818104,” Am. Mineralogist, 31:196 (1946). Hendricks, S. B., “Polymer Chemistry of Silicates, Borates, and Phosphates,” J. Wash. Acad. Sci., 34:241 (1944). Hill, J. D., 1959, Personal Communication. Hummel, F. A., “Properties of Some Substances Isostructural with Silica,” Am. Ceram. Soc. Jour., 32:321 (1949). Huttenlocher, H. F., “Kristallstruktur des Aluminiumorthophosphates AlPO4,” Zeit. Krist., 89—90:508 (1935). Kaiser, W., and Bond, W. L., “Nitrogen, A Major Impurity in Com- mon Type I Diamond,” Bull. Am. Phys. Soc., Ser. 11, No. 1 (1959), p. 27. Klug, H. P., and Alexander, L. E., X-ray Diffraction Procedures for Polycrystalline and Amorphous Materials, John Wiley & Sons, New York, New York, 1954, pp. 716. Manly, R. L., Jr., “The Differential Thermal Analysis of Certain Phosphates,” Am. Mineralogist, 35:108 (1950). Moeller, T., Inorgnic Chemistry, John Wiley & Sons, New York, New York, 1957, pp. 966. 105 Mooney, Rose C. L., “The Crystal Structure of Aluminum Phos- phate and Gallium Phosphates, Low-Cristobalite Type,” Acta Crystallographica, 9:728 (1956). National Bureau of Standards Applied Mathematics Series 10, “Tables for Conversion of X-ray Diffraction Angles to Interplanar Spacing” (1950), p. 121. Norelco Reporter, “An Outline on the Practice of X-ray Spectrog- raphy,” 1:7 (1953). Palache, C., Berman, H., and Frondel, C., Dana’s System of Min— eralogy, Vol. 2, John Wiley & Sons, New York, New York, 1957, pp. 1124. Parrish, W., “X-ray Intensity Measurements with Counter Tubes,” Philips Tech. Review, 17:206 (1956). Parrish, W., Hamacher, E. A., and Lowitzch, K., “The Norelco X-ray Diffractometer,” Philips Tech. Review, 16:123 (1954). Pauling, L., “The Sizes of Ions and Structure of Ionic Crystals,” J. Am. Chem. Soc., 49:765 (1927). Pauling, L., “Structure of Complex Ionic Crystals,” J. Am. Chem. Soc., 51:1012 (1929). Pauling L. , The Nature of the Chemical Bond, 2nd Ed., Cornell Univ. Press, Ithaca, New York, 1940, pp. 450. Perloff, A., “Temperature Inversions of Anhydrous Gallium Ortho- phosphate,” J. Am. Ceramic Soc., 39:83 (1956). Pitzer, K. S., and Hildebrand, J. H., “Color and Bond Character,” J. Am. Chem. Soc., 63:2472 (1941). Ramberg, H., “Chemical Bonds and the Distribution of Cations in Silicates,” Jour. Geology, 60:331 (1952). Rankama, K., and Sahama, Th. G., Geochemistry, Univ. Chicago Press, Chicago, Illinois, 1950, pp. 912. 106 Ringwood, A. E., “The Principles Governing Trace Element Distri- bution During Magmatic Crystallization. Part I,” Geochimi. et Cosmochimi. Acta, 7:189 (1955). Shafer, E. C., and Roy, R., “Studies of Silica-Structure Phases: I, GaPO4, GaAsO4, and GaSbO4,” J. Am. Ceramic Soc., 39:330 (1956). Shafer, E. C., and Roy, R., “Studies of Silica Structure Phases, III: New Data on the System AlPO4,” Zeitschrift fur Physikalische Chemie, Neue Folge, 11:30 (1957). Shafer, E. C., Shafer, M. W., and Roy, R., “Studies of Silica Structure Phases 11: Data on FePO4, MnPO4, BPO4, AlVO4 and Others,” Zeit. Krist., 108:263 (1956). Shafer, M. W., and Roy, R., “Hydrothermal Study of Chromium Orthophosphate,” J. Am. Chem. Soc., 78:1087 (1956). Shaw, D. M., “The Camouflage Principle and Trace Element Distri- bution in Magmatic Minerals,” Jour. Geology, 61:141 (1953). Stanley, J. M., “Hydrothermal Synthesis of Large Aluminum Phos- phate Crystals,” Ind. and Eng. ChemistLy, 46:1684 (1954). Stern, K. H., and Amis, E. S., “Ionic Size,” Chem. Reviews, 59:1 (1959). Strada, M., “Crystalline Structure of Some Phosphates and Arsen- ates of Trivalent Metals: I, Aluminum Phosphate and Aluminum Arsenate,” Gazz. Chim. Ital., 64:653 (1934). Strunz, H., “Titanit and Tilasit. Uber die Verwandtschaft der Silikate mit den Phosphaten und Arsenaten,” Zeit. Krist., 96:7 (1937). Strunz, H., “AlPO4,” Zeit. Krist., 103:228 (1941). Swanson, H. E., and Fuyat, R. K., National Bureau of Standards Circular 539, 3:25 (1954). Van Wazer, J. R., “Phosphorus Chemistry,” Encyclopedia of Chem. Technology, 10:403 (1953). 107 Wasastjerna, J. A., “On the Radii of Ions,” Soc. Sci. Fenn. Comm. Phys. Math., 38:1 (1923). Welcher, F. J., and Hahn, R. B., Semimicro Qualitative Analysis, D. Van Nostrand Company, Inc., Princeton, New Jersey, 1958, pp. 497. Wells, A. F., Structural Inorganic Chemistry, 2nd Ed., Clarendon Press, Oidord, England, 1950, pp. 727. Westbrook, J. H., “Temperature Dependence of Strength and Brittle- ness of Some Quartz Structures,” Am. Ceram. Soc. Jour., 41:433 (1958). Zachariasen, W. H., “A Set of Empirical Crystal Radii for Ions with Inert Gas Configuration,” Zeit. Krist., 80:137 (1931). Zinn, J., 1959, Personal Communication. H 'll‘lu . .v 3. .1 R \ at: a? was...»