Cation-anion radius ratio

Last updated

In condensed matter physics and inorganic chemistry, the cation-anion radius ratio can be used to predict the crystal structure of an ionic compound based on the relative size of its atoms. It is defined as the ratio of the ionic radius of the positively charged cation to the ionic radius of the negatively charged anion in a cation-anion compound. Anions are larger than cations. Large sized anions occupy lattice sites, while small sized cations are found in voids. In a given structure, the ratio of cation radius to anion radius is called the radius ratio. This is simply given by .

Contents

Ratio rule and stability

Critical Radius Ratio. This diagram is for octahedral interstices (coordination number six): 4 anions in the plane shown, 1 above the plane and 1 below. The stability limit is at rC/rA = 0.414 Criticalradiusratio.png
Critical Radius Ratio. This diagram is for octahedral interstices (coordination number six): 4 anions in the plane shown, 1 above the plane and 1 below. The stability limit is at rC/rA = 0.414

The radius ratio rule defines a critical radius ratio for different crystal structures, based on their coordination geometry. [1] The idea is that the anions and cations can be treated as incompressible spheres, meaning the crystal structure can be seen as a kind of unequal sphere packing. The allowed size of the cation for a given structure is determined by the critical radius ratio. [2] If the cation is too small, then it will attract the anions into each other and they will collide hence the compound will be unstable due to anion-anion repulsion; this occurs when the radius ratio drops below the critical radius ratio for that particular structure. At the stability limit the cation is touching all the anions and the anions are just touching at their edges. For radius ratios greater than the critical ratius ratio, the structure is expected to be stable.

The rule is not obeyed for all compounds. By one estimate, the crystal structure can only be guessed about 2/3 of the time. [3] Errors in prediction are partly due to the fact that real chemical compounds are not purely ionic, they display some covalent character. [1]

The table below gives the relation between critical radius ratio, , and coordination number, , which may be obtained from a simple geometrical proof. [4]

Critical Radius RatioCoordination numberType of voidCrystal structureExamples
0.15473 Trigonal-3D-balls.png
Trigonal planar
Kristallstruktur Bortrioxid.png
α-B2O3 structure
B2O3
0.22474 Tetrahedron-2-3D-balls.png
Tetrahedral
Sphalerite polyhedra.png
Zincblende structure
ZnS, CuCl
0.41426 Octahedron-2-3D-balls.png
Octahedral
NaCl octahedra in crystal.svg
Rock salt structure
NaCl, MgO
0.73208 CsCl crystal.png
Cubic
CsCl polyhedra.png
CsCl structure
CsCl, NH4Br

History

The radius ratio rule was first proposed by Gustav F. Hüttig in 1920. [5] [6] In 1926, Victor Goldschmidt [5] extended the use to ionic lattices. [7] [8] [9] In 1929, the rule was incorporated as the first of Pauling's rules for crystal structures. [10]

See also

Related Research Articles

Electronegativity, symbolized as χ, is the tendency for an atom of a given chemical element to attract shared electrons when forming a chemical bond. An atom's electronegativity is affected by both its atomic number and the distance at which its valence electrons reside from the charged nucleus. The higher the associated electronegativity, the more an atom or a substituent group attracts electrons. Electronegativity serves as a simple way to quantitatively estimate the bond energy, and the sign and magnitude of a bond's chemical polarity, which characterizes a bond along the continuous scale from covalent to ionic bonding. The loosely defined term electropositivity is the opposite of electronegativity: it characterizes an element's tendency to donate valence electrons.

<span class="mw-page-title-main">Ionic bonding</span> Chemical bonding involving attraction between ions

Ionic bonding is a type of chemical bonding that involves the electrostatic attraction between oppositely charged ions, or between two atoms with sharply different electronegativities, and is the primary interaction occurring in ionic compounds. It is one of the main types of bonding, along with covalent bonding and metallic bonding. Ions are atoms with an electrostatic charge. Atoms that gain electrons make negatively charged ions. Atoms that lose electrons make positively charged ions. This transfer of electrons is known as electrovalence in contrast to covalence. In the simplest case, the cation is a metal atom and the anion is a nonmetal atom, but these ions can be more complex, e.g. molecular ions like NH+
4
or SO2−
4
. In simpler words, an ionic bond results from the transfer of electrons from a metal to a non-metal to obtain a full valence shell for both atoms.

<span class="mw-page-title-main">Salt (chemistry)</span> Chemical compound involving ionic bonding

In chemistry, a salt or ionic compound is a chemical compound consisting of an ionic assembly of positively charged cations and negatively charged anions, which results in a neutral compound with no net electric charge. The constituent ions are held together by electrostatic forces termed ionic bonds.

In chemistry, the oxidation state, or oxidation number, is the hypothetical charge of an atom if all of its bonds to other atoms were fully ionic. It describes the degree of oxidation of an atom in a chemical compound. Conceptually, the oxidation state may be positive, negative or zero. While fully ionic bonds are not found in nature, many bonds exhibit strong ionicity, making oxidation state a useful predictor of charge.

In chemistry, bond energy (BE) is one measure of the strength of a chemical bond. It is sometimes called the mean bond, bond enthalpy, average bond enthalpy, orbond strength. IUPAC defines bond energy as the average value of the gas-phase bond-dissociation energy for all bonds of the same type within the same chemical species.

Ionic radius, rion, is the radius of a monatomic ion in an ionic crystal structure. Although neither atoms nor ions have sharp boundaries, they are treated as if they were hard spheres with radii such that the sum of ionic radii of the cation and anion gives the distance between the ions in a crystal lattice. Ionic radii are typically given in units of either picometers (pm) or angstroms (Å), with 1 Å = 100 pm. Typical values range from 31 pm (0.3 Å) to over 200 pm (2 Å).

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

In chemistry, triiodide usually refers to the triiodide ion, I
3
. This anion, one of the polyhalogen ions, is composed of three iodine atoms. It is formed by combining aqueous solutions of iodide salts and iodine. Some salts of the anion have been isolated, including thallium(I) triiodide (Tl+[I3]) and ammonium triiodide ([NH4]+[I3]). Triiodide is observed to be a red colour in solution.

Pauling's rules are five rules published by Linus Pauling in 1929 for predicting and rationalizing the crystal structures of ionic compounds.

In chemistry, the lattice energy is the energy change upon formation of one mole of a crystalline ionic compound from its constituent ions, which are assumed to initially be in the gaseous state. It is a measure of the cohesive forces that bind ionic solids. The size of the lattice energy is connected to many other physical properties including solubility, hardness, and volatility. Since it generally cannot be measured directly, the lattice energy is usually deduced from experimental data via the Born–Haber cycle.

In crystallography, a Frenkel defect is a type of point defect in crystalline solids, named after its discoverer Yakov Frenkel. The defect forms when an atom or smaller ion leaves its place in the structure, creating a vacancy and becomes an interstitial by lodging in a nearby location. In elemental systems, they are primarily generated during particle irradiation, as their formation enthalpy is typically much higher than for other point defects, such as vacancies, and thus their equilibrium concentration according to the Boltzmann distribution is below the detection limit. In ionic crystals, which usually possess low coordination number or a considerable disparity in the sizes of the ions, this defect can be generated also spontaneously, where the smaller ion is dislocated. Similar to a Schottky defect the Frenkel defect is a stoichiometric defect. In ionic compounds, the vacancy and interstitial defect involved are oppositely charged and one might expect them to be located close to each other due to electrostatic attraction. However, this is not likely the case in real material due to smaller entropy of such a coupled defect, or because the two defects might collapse into each other. Also, because such coupled complex defects are stoichiometric, their concentration will be independent of chemical conditions.

The bond valencemethod or mean method is a popular method in coordination chemistry to estimate the oxidation states of atoms. It is derived from the bond valence model, which is a simple yet robust model for validating chemical structures with localized bonds or used to predict some of their properties. This model is a development of Pauling's rules.

<span class="mw-page-title-main">Dimanganese decacarbonyl</span> Chemical compound

Dimanganese decacarbonyl, which has the chemical formula Mn2(CO)10, is a binary bimetallic carbonyl complex centered around the first row transition metal manganese. The first reported synthesis of Mn2(CO)10 was in 1954 at Linde Air Products Company and was performed by Brimm, Lynch, and Sesny. Their hypothesis about, and synthesis of, dimanganese decacarbonyl was fundamentally guided by the previously known dirhenium decacarbonyl (Re2(CO)10), the heavy atom analogue of Mn2(CO)10. Since its first synthesis, Mn2(CO)10 has been use sparingly as a reagent in the synthesis of other chemical species, but has found the most use as a simple system on which to study fundamental chemical and physical phenomena, most notably, the metal-metal bond. Dimanganese decacarbonyl is also used as a classic example to reinforce fundamental topics in organometallic chemistry like d-electron count, the 18-electron rule, oxidation state, valency, and the isolobal analogy.

In chemistry, a Zintl phase is a product of a reaction between a group 1 or group 2 and main group metal or metalloid. It is characterized by intermediate metallic/ionic bonding. Zintl phases are a subgroup of brittle, high-melting intermetallic compounds that are diamagnetic or exhibit temperature-independent paramagnetism and are poor conductors or semiconductors.

Lithium superoxide is an unstable inorganic salt with formula LiO2. A radical compound, it can be produced at low temperature in matrix isolation experiments, or in certain nonpolar, non-protic solvents. Lithium superoxide is also a transient species during the reduction of oxygen in a lithium–air galvanic cell, and serves as a main constraint on possible solvents for such a battery. For this reason, it has been investigated thoroughly using a variety of methods, both theoretical and spectroscopic.

<span class="mw-page-title-main">Perovskite</span> Oxide mineral

Perovskite (pronunciation: ) is a calcium titanium oxide mineral composed of calcium titanate (chemical formula CaTiO3). Its name is also applied to the class of compounds which have the same type of crystal structure as CaTiO3, known as the perovskite structure, which has a general chemical formula A2+B4+(X2−)3. Many different cations can be embedded in this structure, allowing the development of diverse engineered materials.

<span class="mw-page-title-main">Conductivity (electrolytic)</span> Measure of the ability of a solution containing electrolytes to conduct electricity

Conductivity or specific conductance of an electrolyte solution is a measure of its ability to conduct electricity. The SI unit of conductivity is siemens per meter (S/m).

Nanofluidic circuitry is a nanotechnology aiming for control of fluids in nanometer scale. Due to the effect of an electrical double layer within the fluid channel, the behavior of nanofluid is observed to be significantly different compared with its microfluidic counterparts. Its typical characteristic dimensions fall within the range of 1–100 nm. At least one dimension of the structure is in nanoscopic scale. Phenomena of fluids in nano-scale structure are discovered to be of different properties in electrochemistry and fluid dynamics.

Pauling's principle of electroneutrality states that each atom in a stable substance has a charge close to zero. It was formulated by Linus Pauling in 1948 and later revised. The principle has been used to predict which of a set of molecular resonance structures would be the most significant, to explain the stability of inorganic complexes and to explain the existence of π-bonding in compounds and polyatomic anions containing silicon, phosphorus or sulfur bonded to oxygen; it is still invoked in the context of coordination complexes. However, modern computational techniques indicate many stable compounds have a greater charge distribution than the principle predicts.

<span class="mw-page-title-main">Tetramethylammonium auride</span> Chemical compound

Tetramethylammonium auride, (CH3)4NAu, is an ionic compound containing tetramethylammonium as cation and gold in a –1 oxidation state as anion. It is an example of a compound containing this rare ionic form of gold, and the first auride paired with a cation that is not a metal atom that has been synthesized.

An oxyhydride is a mixed anion compound containing both oxide O2− and hydride ions H. These compounds may be unexpected as the hydrogen and oxygen could be expected to react to form water. But if the metals making up the cations are electropositive enough, and the conditions are reducing enough, solid materials can be made that combine hydrogen and oxygen in the negative ion role.

References

  1. 1 2 Michmerhuizen, Anna; Rose, Karine; Annankra, Wentiirim; Vander Griend, Douglas A. (2017-08-09). "Radius Ratio Rule Rescue". Journal of Chemical Education. 94 (10). American Chemical Society (ACS): 1480–1485. doi:10.1021/acs.jchemed.6b00970. ISSN   0021-9584.
  2. Pauling, Linus (1960). Nature of the Chemical Bond (3rd ed.). Ithaca, New York: Cornell University Press. p. 544. ISBN   9780801403330.
  3. Nathan, Lawrence C. (1985). "Predictions of crystal structure based on radius ratio: How reliable are they?". Journal of Chemical Education. 62 (3). American Chemical Society (ACS): 215. doi:10.1021/ed062p215. ISSN   0021-9584.
  4. Toofan, Jahansooz (1994). "A Simple Expression between Critical Radius Ratio and Coordination Number". Journal of Chemical Education. 71 (2). American Chemical Society (ACS): 147. doi:10.1021/ed071p147. ISSN   0021-9584. (and Erratum 71(9): 749 doi : 10.1021/ed071p749), Following the erratum, equations should read and .
  5. 1 2 Jensen, William B. (2010-04-23). "The Origin of the Ionic-Radius Ratio Rules". Journal of Chemical Education. 87 (6). American Chemical Society (ACS): 587–588. doi:10.1021/ed100258f. ISSN   0021-9584.
  6. Hüttig, Gustav F. (1920-11-11). "Notiz zur Geometrie der Koordinationszahl". Zeitschrift für anorganische und allgemeine Chemie (in German). 114 (1). Wiley: 24–26. doi:10.1002/zaac.19201140103. ISSN   0863-1786.
  7. Goldschmidt, V.; Barth, T.; Lunde, G.; Zachariasen, W. (1926). Geochemische Verteilungsgesetze der Elemente. VII. Die Gesetze der Krystallochemie (in German). Oslo: Dybwad. pp. 112–117. OCLC   174577644.
  8. Goldschmidt, V. (1927). Geochemische Verteilungsgesetze der Elemente. VIII. Untersuchungen über Bau und Eigenschaften von Krystallen (in German). Oslo: Dybwad. pp. 14–17. OCLC   19831825.
  9. Goldschmidt, V. M. (1929). "Crystal structure and chemical constitution". Transactions of the Faraday Society. 25. Royal Society of Chemistry (RSC): 253. doi:10.1039/tf9292500253. ISSN   0014-7672.
  10. Pauling, Linus (1929). "The principles determining the structure of complex ionic crystals". Journal of the American Chemical Society. 51 (4). American Chemical Society (ACS): 1010–1026. doi:10.1021/ja01379a006. ISSN   0002-7863.