Bararite

Last updated
Bararite
Bararite.jpg
Pale pink yellowish crystals of bararite from Shamokin, Northumberland County, Pennsylvania, USA
General
Category Halide mineral
Formula
(repeating unit)
(NH4)2SiF6
IMA symbol Brr [1]
Strunz classification 3.CH.10
Crystal system Trigonal
Crystal class Hexagonal scalenohedral (3m)
H-M symbol: (3 2/m)
Space group P3m1
Unit cell a = 5.77 Å, c = 4.78 Å; Z = 1
Identification
ColorWhite to colorless
Crystal habit Tabular, sometimes elongated on {0001},
also appears in irregularly shaped or mammillary surfaces that comprise mainly cryptohalite
Twinning Interpenetration twins (paddlewheels/darts), axis parallel to {0001}
Cleavage [0001] perfect
Mohs scale hardness2.5
Luster Vitreous
Diaphaneity Transparent
Specific gravity 2.152 (synthetic)
Optical propertiesUniaxial (-)
Refractive index nω = 1.406 ± 0.001,
nε = 1.391 ± 0.003
Birefringence 0.015 ± 0.003
Solubility Dissolves in water
Other characteristicssalty taste
References [2] [3] [4] [5] [6] [7] [8]

Bararite is a natural form of ammonium fluorosilicate (also known as hexafluorosilicate or fluosilicate). It has chemical formula (NH4)2SiF6 and trigonal crystal structure. This mineral was once classified as part of cryptohalite. Bararite is named after the place where it was first described, Barari, Dhanbad, India. [3] It is found at the fumaroles of volcanoes (Vesuvius, Italy), over burning coal seams (Barari, India), and in burning piles of anthracite (Pennsylvania, U.S.). It is a sublimation product that forms with cryptohalite, sal ammoniac, and native sulfur. [4]

Contents

History

A. Scacchi first discovered cryptohalite in 1873. [3] It appeared in a volcanic sublimate from the Vesuvian eruption of 1850. In 1926, W.A.K. Christie reported his own chemical study. A microscope was used to pick out enough material for analysis. Distilling with sodium hydroxide (NaOH) produced ammonia (NH3). The anions of hexafluorosilicic acid (H2SiF6) precipitated as potassium fluorosilicate (K2SiF6). Barium sulfate (BaSO4) was thrown into the filtrate, and then calcium fluoride (CaF2). Christie found 20.43% (NH4)+ and 78.87% (SiF6)2−. [5]

Bararite is named after Barari, a locality in India. This was where the species was first completely described. Earlier, bararite was recognized as part of mixtures with cryptohalite. [3] However, it did not receive its own name until 1951. [3] [9] The East Indian Coal Company provided the sample that Christie used to evaluate bararite. [5]

Bararite has not received a quantitative chemical analysis in its natural form. [4] Christie received far too little for more than qualitative analysis through microchemistry. He utilized F. Emich's methods with capillary tube centrifuges. [5]

Structure

Bararite is the beta, trigonal (scalenohedral) form of ammonium hexafluorosilicate. Its symmetry is 32/m. [4] The space group is P3m1. The a-axes in the unit cell are 5.784 ± 0.005 Å (angstroms), and the c-axis is 4.796 ± 0.006 Å. The unit lattice is primitive. [6] [10] (Note: Data for the space group come from synthetic crystals.) Cryptohalite has the cubic (isometric) crystal structure and corresponds to the alpha form. Both minerals have the chemical formula (NH4)2SiF6. The halides of form AmBX6 fall into two groups: hieratite and malladrite. The hieratite group is isometric whereas the malladrite is hexagonal. [3]

The (SiF6)2− is octahedral—one fluorine atom at each vertex. [11] In bararite, the (NH4)+’s are trigonally coordinated. They all appear at sites of C3v (3m) symmetry. The (NH4)+ has 12 fluorine neighbors, which form four triangles. Three of these triangles are isosceles. These triangles themselves form a triangle—around the 3-fold axis containing the nitrogen atom. One triangle is equilateral. Its symmetry axis is the same axis that goes through the nitrogen atom. [12] (For structural diagrams, see link to unit cell [6] and downloadable articles [12] [13] in “References.”)

The silicon atoms of cryptohalite, α-(NH4)2SiF6 (alpha), have cubic close(st) packing (CCP). A third form (gamma, γ) of (NH4)2SiF6 uses hexagonal close(st) packing (HCP). Bararite, β-(NH4)2SiF6, utilizes hexagonal primitive (HP) packing. Layers with distorted octahedral gaps separate those with the anions. The (NH4)+ ions appear a little below and above the (SiF6)2−. In all three phases, 12 fluorine atoms neighbor the (NH4)+. Distances range from about 3.0 to 3.2 Å. [13] The (NH4)+ has no free rotation. It only librates (oscillates)—at least when vibrationally excited. [12]

As a salt, bararite is an ionic compound. The ions, of course, have ionic bonding. The atoms of polyatomic ions are held together covalently. The orientation of (NH4)+ is sustained by four trifurcated (three-branch) hydrogen bonds. These bonds point toward the triangles containing the 12 fluorine neighbors. Three H bonds are equivalent. The fourth bond, pointing toward the equilateral triangle, has a shorter distance. [12]

The intermolecular distances between fluorine atoms are smaller in bararite (3.19 and 3.37 Å) than cryptohalite. In cryptohalite, each anion is coordinated to 12 others. Bararite has (2+6)-fold coordination. The two Si-Si distances between layers (4.796 ± 0.006 Å) do not equal the six within a layer (5.784 ± 0.005 Å). Bararite is more compressible along the c-axis than the a-axis. [13]

Bararite has no known solution or exsolution, but it is always mixed with other substances (cryptohalite, sal ammoniac, and sulfur). [4] Due to thermal motion, atomic behavior of ammonium salts can be very hard to evaluate. [11] The anions, however, are ordered and have no unusual motion from heat. [6]

A third form of (NH4)2SiF6 was discovered in 2001 and identified with the 6mm symmetry (hexagonal). [13] In all three arrangements, the (SiF6)2− octahedra come in layers. In the cubic form (cryptohalite), these layers are perpendicular to [111]. [13] In the trigonal (bararite) and hexagonal (gamma, γ) forms, the layers are perpendicular to the c-axis. [13] (Note: Trigonal crystals are part of the hexagonal group. But not all hexagonal crystals are trigonal. [14] )

Although bararite was claimed to be metastable at room temperature, [11] it does not appear one polymorph has ever turned into another. [13] Still, bararite is fragile enough that grinding it for spectroscopy will produce a little cryptohalite. [12] Even so, ammonium fluorosilicate assumes a trigonal form at pressures of 0.2 to 0.3 giga-pascals (GPa). The reaction is irreversible. If this phase is not bararite, it is at least very closely related. [13]

The hydrogen bonding in (NH4)2SiF6 allows this salt to change phases in ways that normal salts cannot. Interactions between cations and anions are especially important in how ammonium salts change phase. [13]

Physical properties

Bararite forms tabular crystals. They are flattened, sometimes elongated, on {0001} (perpendicular to c). [3] Christie reported tiny, transparent crystals of bararite that looked like paddlewheels and darts. Each had four barbs at 90°. The crystals reached up to 1 mm long, the barbs up to 0.2 mm wide. They were interpenetration twins, the twin axis perpendicular to the c-axis. [5] Visually, cryptohalite crystals are almost impossible to discern from sal ammoniac (NH4Cl). [15] Inclusions of bararite in cryptohalite can be seen only with plane-polarized light. [16]

Bararite has perfect cleavage on the {0001} plane. The hardness is probably 2+12. [3] The anions (as already shown) are bonded much more strongly within layers than between layers. Also, ionic bonds are not the strongest bonds, and halides cannot normally scratch glass plates. [14]

Bararite has a measured density of 2.152 g/mL (synthetic)—but a calculated density of 2.144 g/mL. It tastes salty, and it dissolves in water. Its luster is vitreous (like glass). Bararite is white to colorless. [3] These properties are similar to halite (NaCl) [14] —which gave the halide group its name.

Whereas cryptohalite belongs to the isotropic optical class, bararite is uniaxial negative. [17] [3] At 1.391 ± 0.003, the refractive index through c is smaller than through a (1.406 ± 0.001). [5] The c-axis in bararite is shorter than the a-axes (see “Structure”). Furthermore, only this path lets light hit nothing but the same ion in the same orientation (all the layers have the same structure and orientation [13] ).

Bararite has about a 6% greater density than cryptohalite. [13] As discussed before, its structure is more packed. This substance can be produced easily from aqueous solution, [3] but only below 5 °C (41 °F) will pure bararite form. [3] [18] Above 13 °C (55 °F), almost pure cryptohalite emerges. [3] [5] Bararite sublimes without leaving residue. [3]

Geologic occurrence

In nature, bararite appears with cryptohalite, sal ammoniac, and native sulfur. [5] [15] It is found over a burning coal seam in Barari, India, [5] and as a sublimation product in Vesuvius, Italy, at fumaroles (opening in or near a volcano where hot sulfurous gases come out). [4] [19] It also is found in the United States, in Pennsylvania. It appears in burning piles of anthracite (highest grade of coal)—again as a sublimation product. [15]

Christie found translucent arborescent (treelike) crystals, with vitreous luster. He found white, opaque lumps that were a mixture of (NH4)2SiF6 with SiO2. They were irregularly shaped but usually had a mammillary surface (several convex surfaces smoothly rounded). These hold primarily cryptohalite but also some bararite. [5] In Pennsylvania, bararite normally comes as tiny inclusions in cryptohalite crystals. [15] [16] It appears that first, bararite forms through direct sublimation. Afterward, it quickly changes to cryptohalite. [16]

In Barari, burning-coal gases go through a dike (igneous intrusion) of mica and peridotite. The sulfur dioxide must attack apatite in the dike, which produces hydrofluoric acid that attacks the abundant silicates. Silicon fluoride is formed. Ammonia also comes from burning coal. From there, ammonium fluorosilicate can form. A slight excess of ammonia could lead to the white lumps of silica and cryptohalite. Bararite and cryptohalite in their pure forms, for the most part, grow out of these nodules. Recrystallization from the rain is probably responsible. [5]

Fluorosilicate minerals are thermodynamically unstable in soil. [20] Still, intense heat promotes the formation of (NH4)2SiF6 to some degree—as seen in some experiments by Rehim. But this compound will break up at 320 to 335 °C. [21] Both burning coal [5] [15] and volcanoes are important sources of SO2 and SiF4. [22]

Chemical properties and uses

Fluorosilicic acid and its salts are poisonous. [23] Ammonium fluorosilicate, however, is very rare in nature [15] and apparently much easier to synthesize. [3]

Related Research Articles

<span class="mw-page-title-main">Silicate</span> Any polyatomic anion containing silicon and oxygen

In chemistry, a silicate is any member of a family of polyatomic anions consisting of silicon and oxygen, usually with the general formula [SiO(4-2x)−
4−x
]
n
, where 0 ≤ x < 2. The family includes orthosilicate SiO4−4, metasilicate SiO2−3, and pyrosilicate Si2O6−7. The name is also used for any salt of such anions, such as sodium metasilicate; or any ester containing the corresponding chemical group, such as tetramethyl orthosilicate. The name "silicate" is sometimes extended to any anions containing silicon, even if they do not fit the general formula or contain other atoms besides oxygen; such as hexafluorosilicate [SiF6]2−.Most commonly, silicates are encountered as silicate minerals.

<span class="mw-page-title-main">Ammonium chloride</span> Chemical compound

Ammonium chloride is an inorganic chemical compound with the chemical formula NH4Cl, also written as [NH4]Cl. It is an ammonium salt of hydrogen chloride. It consists of ammonium cations [NH4]+ and chloride anions Cl. It is a white crystalline salt that is highly soluble in water. Solutions of ammonium chloride are mildly acidic. In its naturally occurring mineralogic form, it is known as salammoniac. The mineral is commonly formed on burning coal dumps from condensation of coal-derived gases. It is also found around some types of volcanic vents. It is mainly used as fertilizer and a flavouring agent in some types of liquorice. It is a product of the reaction of hydrochloric acid and ammonia.

<span class="mw-page-title-main">Salammoniac</span> Halide mineral

Salammoniac, also sal ammoniac or salmiac, is a rare naturally occurring mineral composed of ammonium chloride, NH4Cl. It forms colorless, white, or yellow-brown crystals in the isometric-hexoctahedral class. It has very poor cleavage and is brittle to conchoidal fracture. It is quite soft, with a Mohs hardness of 1.5 to 2, and it has a low specific gravity of 1.5. It is water-soluble. Salammoniac is also the archaic name for the chemical compound ammonium chloride.

<span class="mw-page-title-main">Epitaxy</span> Crystal growth process relative to the substrate

Epitaxy refers to a type of crystal growth or material deposition in which new crystalline layers are formed with one or more well-defined orientations with respect to the crystalline seed layer. The deposited crystalline film is called an epitaxial film or epitaxial layer. The relative orientation(s) of the epitaxial layer to the seed layer is defined in terms of the orientation of the crystal lattice of each material. For most epitaxial growths, the new layer is usually crystalline and each crystallographic domain of the overlayer must have a well-defined orientation relative to the substrate crystal structure. Epitaxy can involve single-crystal structures, although grain-to-grain epitaxy has been observed in granular films. For most technological applications, single-domain epitaxy, which is the growth of an overlayer crystal with one well-defined orientation with respect to the substrate crystal, is preferred. Epitaxy can also play an important role while growing superlattice structures.

<span class="mw-page-title-main">Ammonium fluoride</span> Chemical compound

Ammonium fluoride is the inorganic compound with the formula NH4F. It crystallizes as small colourless prisms, having a sharp saline taste, and is highly soluble in water. Like all fluoride salts, it is moderately toxic in both acute and chronic overdose.

<span class="mw-page-title-main">Chalcogenide</span>

A chalcogenide is a chemical compound consisting of at least one chalcogen anion and at least one more electropositive element. Although all group 16 elements of the periodic table are defined as chalcogens, the term chalcogenide is more commonly reserved for sulfides, selenides, tellurides, and polonides, rather than oxides. Many metal ores exist as chalcogenides. Photoconductive chalcogenide glasses are used in xerography. Some pigments and catalysts are also based on chalcogenides. The metal dichalcogenide MoS2 is a common solid lubricant.

<span class="mw-page-title-main">Hexafluorosilicic acid</span> Octahedric silicon compound

Hexafluorosilicic acid is an inorganic compound with the chemical formula H
2
SiF
6
. Aqueous solutions of hexafluorosilicic acid consist of salts of the cation and hexafluorosilicate anion. These salts and their aqueous solutions are colorless.

<span class="mw-page-title-main">Ammonium bifluoride</span> Chemical compound

Ammonium bifluoride is the inorganic compound with the formula [NH4][HF2] or [NH4]F·HF. It is produced from ammonia and hydrogen fluoride. This colourless salt is a glass-etchant and an intermediate in a once-contemplated route to hydrofluoric acid.

Silicon compounds are compounds containing the element silicon (Si). As a carbon group element, silicon often forms compounds in the +4 oxidation state, though many unusual compounds have been discovered that differ from expectations based on its valence electrons, including the silicides and some silanes. Metal silicides, silicon halides, and similar inorganic compounds can be prepared by directly reacting elemental silicon or silicon dioxide with stable metals or with halogens. Silanes, compounds of silicon and hydrogen, are often used as strong reducing agents, and can be prepared from aluminum–silicon alloys and hydrochloric acid.

<span class="mw-page-title-main">Tin(IV) fluoride</span> Chemical compound

Tin(IV) fluoride is a chemical compound of tin and fluorine with the chemical formula SnF4 and is a white solid with a melting point above 700 °C.

<span class="mw-page-title-main">Halide mineral</span> Minerals with a dominant fluoride, chloride, bromide, or iodide anion

Halide minerals are those minerals with a dominant halide anion. Complex halide minerals may also have polyatomic anions.

<span class="mw-page-title-main">Hexagonal crystal family</span> Union of crystal groups with related structures and lattices

In crystallography, the hexagonal crystal family is one of the 6 crystal families, which includes two crystal systems and two lattice systems. While commonly confused, the trigonal crystal system and the rhombohedral lattice system are not equivalent. In particular, there are crystals that have trigonal symmetry but belong to the hexagonal lattice.

Ammonium fluorosilicate (also known as ammonium hexafluorosilicate, ammonium fluosilicate or ammonium silicofluoride) has the formula (NH4)2SiF6. It is a toxic chemical, like all salts of fluorosilicic acid. It is made of white crystals, which have at least three polymorphs and appears in nature as rare minerals cryptohalite or bararite.

In chemistry, molecular oxohalides (oxyhalides) are a group of chemical compounds in which both oxygen and halogen atoms are attached to another chemical element A in a single molecule. They have the general formula AOmXn, where X is a halogen. Known oxohalides have fluorine (F), chlorine (Cl), bromine (Br), and/or iodine (I) in their molecules. The element A may be a main group element, a transition element, a rare earth element or an actinide. The term oxohalide, or oxyhalide, may also refer to minerals and other crystalline substances with the same overall chemical formula, but having an ionic structure.

Farneseite is a mineral from the cancrinite sodalite group with 14 layer stacking. It is a complex silicate mineral with formula (Na,Ca,K)56(Al6Si6O24)7(SO4)12·6H2O. It was named after a location in Farnese, Lazio, Italy. It is a member of the cancrinite-sodalite group, approved in 2004 as a new mineral species. The group is characterized by the number of stacking layers making up each member, with farneseite being one of newest minerals in the group with a 14 layer stacking structure. It is a clear transparent mineral and has a hexagonal crystal system with crystal class of 6/m and space group of P63/m. The specimens discovered in Farnese were in a pyroclastic rock from the Làtera Cauldera region.

Nickel compounds are chemical compounds containing the element nickel which is a member of the group 10 of the periodic table. Most compounds in the group have an oxidation state of +2. Nickel is classified as a transition metal with nickel(II) having much chemical behaviour in common with iron(II) and cobalt(II). Many salts of nickel(II) are isomorphous with salts of magnesium due to the ionic radii of the cations being almost the same. Nickel forms many coordination complexes. Nickel tetracarbonyl was the first pure metal carbonyl produced, and is unusual in its volatility. Metalloproteins containing nickel are found in biological systems.

<span class="mw-page-title-main">Thorium compounds</span> Chemical compounds

Many compounds of thorium are known: this is because thorium and uranium are the most stable and accessible actinides and are the only actinides that can be studied safely and legally in bulk in a normal laboratory. As such, they have the best-known chemistry of the actinides, along with that of plutonium, as the self-heating and radiation from them is not enough to cause radiolysis of chemical bonds as it is for the other actinides. While the later actinides from americium onwards are predominantly trivalent and behave more similarly to the corresponding lanthanides, as one would expect from periodic trends, the early actinides up to plutonium have relativistically destabilised and hence delocalised 5f and 6d electrons that participate in chemistry in a similar way to the early transition metals of group 3 through 8: thus, all their valence electrons can participate in chemical reactions, although this is not common for neptunium and plutonium.

Lucabindiite is a mineral discovered in 1998 from the La Fossa crater at Vulcano, the Aeolian islands off the coast of Italy. It has the chemical formula As4O6(Cl,Br) and is hexagonal. After months of collecting sublimates and encrustations, the researchers discovered lucabindiite which was found on the surface of pyroclastic breccia. The mineral is named after Luca Bindi, who was a professor of mineralogy and former head of the Division of Mineralogy of the Natural History Museum of the University of Florence.

<span class="mw-page-title-main">Aluminium compounds</span>

Aluminium (British and IUPAC spellings) or aluminum (North American spelling) combines characteristics of pre- and post-transition metals. Since it has few available electrons for metallic bonding, like its heavier group 13 congeners, it has the characteristic physical properties of a post-transition metal, with longer-than-expected interatomic distances. Furthermore, as Al3+ is a small and highly charged cation, it is strongly polarizing and aluminium compounds tend towards covalency; this behaviour is similar to that of beryllium (Be2+), an example of a diagonal relationship. However, unlike all other post-transition metals, the underlying core under aluminium's valence shell is that of the preceding noble gas, whereas for gallium and indium it is that of the preceding noble gas plus a filled d-subshell, and for thallium and nihonium it is that of the preceding noble gas plus filled d- and f-subshells. Hence, aluminium does not suffer the effects of incomplete shielding of valence electrons by inner electrons from the nucleus that its heavier congeners do. Aluminium's electropositive behavior, high affinity for oxygen, and highly negative standard electrode potential are all more similar to those of scandium, yttrium, lanthanum, and actinium, which have ds2 configurations of three valence electrons outside a noble gas core: aluminium is the most electropositive metal in its group. Aluminium also bears minor similarities to the metalloid boron in the same group; AlX3 compounds are valence isoelectronic to BX3 compounds (they have the same valence electronic structure), and both behave as Lewis acids and readily form adducts. Additionally, one of the main motifs of boron chemistry is regular icosahedral structures, and aluminium forms an important part of many icosahedral quasicrystal alloys, including the Al–Zn–Mg class.

Protactinium compounds are compounds containing the element protactinium. These compounds usually have protactinium in the +5 oxidation state, although these compounds can also exist in the +2, +3 and +4 oxidation states.

References

  1. Warr, L.N. (2021). "IMA–CNMNC approved mineral symbols". Mineralogical Magazine. 85 (3): 291–320. Bibcode:2021MinM...85..291W. doi: 10.1180/mgm.2021.43 . S2CID   235729616.
  2. Mineralienatlas
  3. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Palache, C., Berman, H., and Frondel, C. (1951) Dana’s System of Mineralogy, Volume II: Halides, Nitrates, Borates, Carbonates, Sulfates, Phosphates, Arsenates, Tungstates, Molybdates, etc. John Wiley and Sons, Inc., New York, 7th edition.
  4. 1 2 3 4 5 6 Anthony, J.W., Bideaux, R.A., Bladh, K.W., and Nichols, M.C. (1997) Handbook of Mineralogy, Volume III: Halides, Hydroxides, Oxides. Mineral Data Publishing, Tucson.
  5. 1 2 3 4 5 6 7 8 9 10 11 12 Christie, W.A.K. (1926) An Occurrence of Cryptohalite (Ammonium Fluosilicate). Records of the Geological Survey of India, 59, 233.
  6. 1 2 3 4 Schlemper, E.O. and Hamilton, W.C. (1966) On the Structure of Trigonal Ammonium Fluorosilicate. Journal of Chemical Physics, 45, 408–409.
  7. Bararite on Mindat.org
  8. Bararite data on Webmineral
  9. Fleischer, M. (1952) "New Mineral Names". American Mineralogist, 37, 359–362.
  10. Anthony et al. (1997) and Palache et al. (1951) use outdated information for the crystal axes. The information in these handbooks is linked ultimately to two articles by Gossner and Krauss from 1934, in Zeitschrift für Kristallographie. The replacement source, Schlemper and Hamilton (1966), is cited not just by this article but also Boldyreva et al. (2007).
  11. 1 2 3 Schlemper, Elmer O. (1966). "Structure of Cubic Ammonium Fluosilicate: Neutron-Diffraction and Neutron-Inelastic-Scattering Studies". The Journal of Chemical Physics. 44 (6): 2499–2505. Bibcode:1966JChPh..44.2499S. doi:10.1063/1.1727071.
  12. 1 2 3 4 5 Oxton, I.A., Knop, O., and Falk, M. (1975) "Infrared Spectra of the Ammonium Ion in Crystals". II. The Ammonium Ion in Trigonal Environments, with a Consideration of Hydrogen Bonding. Canadian Journal of Chemistry, 53, 3394–3400.
  13. 1 2 3 4 5 6 7 8 9 10 11 Boldyreva, E.V.; Shakhtshneider, T.P; Sowa, H. & Ahsbas, H. (2007). "Effect of hydrostatic pressure up to 6 GPa on the crystal structures of ammonium and sodium hexafluorosilicates, (NH4)2SiF6 and Na2SiF6; a phase transition in (NH4)2SiF6 at 0.2–0.3 GPa". Zeitschrift für Kristallographie. 222 (1): 23–29. Bibcode:2007ZK....222...23B. doi:10.1524/zkri.2007.222.1.23. S2CID   97174719.
  14. 1 2 3 Klein, C. and Dutrow, B. (2008) The 23rd Edition of the Manual of Mineral Science. John Wiley & Sons, Hoboken, NJ.
  15. 1 2 3 4 5 6 Barnes, J. and Lapham, D. (1971) Rare Minerals Found in Pennsylvania. Pennsylvania Geology, 2, 5, 6–8.
  16. 1 2 3 Lapham, D.M., Barnes, J.H., Downey, W.F., Jr., and Finkelman, R.B. (1980) Bararite. Mineralogy associated with burning anthracite deposits of eastern Pennsylvania. Pennsylvania Geological Survey: Mineral Resource Report, 78, 45–47.
  17. To learn what makes a uniaxial crystal, visit Introduction to Uniaxial Minerals.
  18. Gossner, B. (1903) Ammoniumsiliciumfluorid. Zeitschrift für Kristallographie, 38, 147–148.
  19. Scacchi, A. (1874) "Appendice alle contribuzioni mineralogiche sull’ incendio vesuviano del 1872". Rendiconto dell’Accademia delle scienze fisiche e matematiche (sezione della Società reale di Napoli), 8, 179–180.
  20. Elrashidi, M.A. and Lindsay, W.L. (1986) Chemical Equilibria of Fluorine in Soils: A Theoretical Development. Soil Science: An Interdisciplinary Approach to Soil Research, 141, 274–280.
  21. Rehim, A. M. Abdel (1992). "Thermal study of synthesis of cryptohalite". Journal of Thermal Analysis. 38 (3): 475–486. doi:10.1007/BF01915512. S2CID   95930085.
  22. Mori, T., Sato, M., Shimoike, Y., and Notsu, K. (2002) "High SiF4 ratio detected in Satsuma-Iwojima volcano’s plume by remote FT-IR observation". Earth Planets Space, 54, 249–256.
  23. Wiberg, E., Wiberg, N., and Holleman, A.F. (2001) Inorganic chemistry. Academic Press, San Diego.