Abstract
Periodicity of elements is the basis of teaching and understanding inorganic chemistry. This review exemplifies simple rules and counting procedures as heuristic algorithms yielding often-dimensionless quantities that, as such or together with auxiliary parameters, allow us to predict not only the stoichiometry and bonding of compounds, but also some of their properties or reactions.
1 Introduction
Our understanding of chemistry follows the periodic system of repeat pattern controlled by the principal quantum number n (1, 2, 3, 4 …) of the periods. The azimuthal quantum number ℓ (of integer values from 0 up to n−1) controls orbital types (s, p, d, f …) in a given period, and the magnetic quantum number m ℓ (−ℓ to +ℓ) tells apart orbitals within each type (as one s, three p, five d, seven f, etc.). The orbital-energy sequence for isolated atom proceeds in order of increasing n+ℓ, if equal, then of n. Down the group of a given orbital type, the similarity evolves via increased and symmetry-dependent shielding of valence orbitals. Similarity along each period concerns gradual changes in properties of a step that diminishes with increasing ℓ, affected by slight irregularities due to high-symmetry s-orbital shielding and by half-filling stabilizations.
Chemistry is about bonding and reactions. Simple quantities that characterize bonding are essential for communication in chemistry and for simple predictions of properties and reactions. The heuristic approach to chemistry involves simple electron-counting rules and algorithms that may use simple parameters such as element electronegativity or standard potential for redox half-reactions. A review of some rules and heuristic-algorithm examples follows.
2 Rules
The basic rules in chemistry concern stable electron configurations. Completely filled electron shells possess increased stability: the 1s shell, the highest ns-shells, the sp-shell, and the dsp-shell. In addition, a half-filled d or f shell may behave as a stabilized configuration. The important formulations are the octet rule, the 18-plet rule, and the 8−N rule, which in some sense predict bonding. The counting requires the chemist to remember (or see) the periodic system of elements.
2.1 Octet rule and 18-plet rule
The concept of the rule of eight appeared simultaneously in papers by Lewis [1] and Kossel [2], and the term “octet” was introduced shortly after in the paper by Langmuir [3]. Inorganic chemistry textbook [4] p. 40 defines it: “An atom obeys the octet rule when it gains, loses or shares electrons to give an outer shell containing eight electrons (an octet) with a configuration ns 2 np 6.”
A current Wikipedia wording [5] goes beyond octet: “During the formation of a chemical bond, atoms combine together by gaining, losing or sharing electrons in such a way that they acquire nearest noble-gas configuration.” For bonded sp-elements, it implies octet (or doublet for ionic compounds of Li and Be), only the “nearest” is an adjective too strong, as boron tends to achieve octet and not the doublet. For d-elements acquiring and sharing electrons upon forming chemical bonds, we specify the noble-gas configuration as the nearest that follows; the 18-plet rule. Its history starts with Langmuir [6], as noted in Ref. [7] that deals with chemistry aspects of this rule. This somewhat weaker rule stabilizes some compounds of transition metals via electron-pair donation from ligands and by formation of metal–metal bonds. Neither is the octet rule without exceptions. So called “expansion of octet” applies to a heavier p-element atom bonded to several highly electronegative atoms.
To apply the octet rule on a simple compound, one distributes all available valence electrons upon drawing the Lewis formula of the molecule or ion, most conveniently with electron pairs as dashes. Often it is unambiguous. At times not. For N2O of 16 valence electrons, the formula
2.1.1 The 8−N rule
The “eight minus N” rule has been recognized on the structures of the p-block nonmetallic elements [8]:
Fluorine forms a molecule bonded by 1 two-electron bond.
Oxygen is bonded by 2 two-electron bonds, and so is sulfur.
Nitrogen is bonded by 3 two-electron bonds, and so is phosphorus.
Carbon is bonded by 4 two-electron bonds, and so is silicon.
Formulation for compounds in [9] p. 1027: “An electronegative sp atom of N valence electrons tends to form 8 − N, but no more than four, two-electron bonds with atoms of equal or lower electronegativity.”
It is important to recognize that the 8−N rule is enforced by the sp-atom of higher electronegativity. In SF2, SF4, and SF6, it is fluorine, and all bonds have approximately the same length as a single bond [10]. In the series BF, CO, and N2, the triple bond suggested by the octet rule occurs in N2 only, while O and F adjust the bond order towards 2 and 1, respectively [11]. Our
3 Algorithms to predict bonding
Bonding prediction for a given composition usually takes form of a simple algorithm utilizing rules to calculate the number N BO of bonding orbitals and N NBO of essentially nonbonding orbitals (lone pairs) at the atom. Their sum shows whether an octet or 18-plet is achieved.
3.1 Generalized 8−N rule
The two approaches described below are in fact combinations of the octet rule and of the 8−N rule.
3.1.1 “Ionic” VECA approach
For an sp-block non-metallic binary compound C c A a of a more electropositive “cation C” and an electronegative “anion A”, the algorithm of generalized 8−N rule originates in the Zintl concept [14] formalized by Parthé [15, 16], with history reviewed by Nesper [17]. The algorithm is:
Take C c A a binary compound and calculate the valence-electron total per the “anion” A, VEC A.
Electrons in excess of 8 remain at the electropositive “cation” C as C–C bonds or lone pairs.
VEC A = 8 means stable configuration at both atoms.
Electrons short of 8 are gained by forming A–A bonds like elements do:
1e by forming dumbbells.
2e by infinite chain, or double bond when A is small.
3e by tetrahedron, or by 3-connected net or sheet.
4e by a 4-connected net or by graphene-like sheet when A is small.
Examples: In NaTl of VEC
A = 4, the Tl atom gains the 4 missing electrons by forming four bonds in an anionic diamond-type network of the
For KI3, the underlying octet rule improves the picture of bonding: KI3 has VEC
A = 22/3 = 7 1/3, and 2/3 of an electron per iodine is to be obtained in bonds. The three iodine atoms miss 2 electrons, which they obtain by forming two 1-electron bonds (
When the electronegativity difference decreases, exceptions arise that combine both alternatives (of the above algorithm point 2 and 4), such as in the GeAs2 structure [19] where Ge keeps two electrons, VEC
A = 6 (instead of expected 14/2 = 7), and two
3.1.2 “Covalent” VEC approach
For the p-block non-metallic binaries, counting plain VEC, the valence-electron count per atom [15], may help. The 8−N rule then gives the average number of bonding orbitals at the atom in the generally sp 3 molecule as N BO = 8 − VEC. It yields many options. For molecules of the smallest atoms from top of the periodic table, multiple bonds can be encountered, like the N BO = 8/3 average per atom in N2O of Lewis formulas discussed above. This multiple-bond possibility should be considered in addition to the following single-bond based outcomes:
N BO = 1 sugests a single-bonded molecule of the two element atoms,
for 1 < N BO < 2, a molecule of a finite chain with or without branches is formed,
N BO = 2 suggests an infinite chain or a ring for a 1:1 binary compound,
for 2 < N BO < 3, cluster-like molecules or chains with intermittent connections into layers or other complicated shapes appear.
N BO = 3 suggests a 1:1 binary compound forming a binodal sheet or a tetrahedral molecule.
for 3 < N BO < 4, various networks, layers with interconnected infinite tubes, etc.
N BO = 4 for 1:1 binary suggests a binodal 4,4-connected network (sphalerite- or wurtzite-type).
An octet means that N
BO + N
NBO = 4. An integer VEC typically means two elements in 1:1 ratio that are either from the same group (such as the single-bonded BrF and As2P2) or from next-neighbor groups. In the latter case, the VEC value might apply for both as an average. An example is GeS of VEC = (4 + 6)/2 = 5, where the Ge and S atoms average to form 8 − VEC = 3 bonds each, upon achieving octet in a corrugated sheet of the black-phosphorus type. When the electron count does not have an integer average, both atoms tend to fulfill own 8−N rule and octet, in a large variety of bonding arrangements. An example from group 5 and 6 is As4S4 of VEC = 5.5 giving N
BO = 2.5, where all S form two bonds and all As three bonds in a cluster-like molecule shaped as if four of the six edges of the As elemental tetrahedron were replaced by
This 8−N covalent VEC approach can also be adopted to check polyatomic anions of one single p-element, some of them chains, some of them clusters. The chain example is Si2− of VEC = 6 giving N BO = 2 of an infinite anionic chain (alternatively a ring). A cluster example is As7 3− of 38 electrons, with VEC = 38/7 giving N BO = (56 − 38)/7 per atom, a number that seems no longer helpful. We may, however, realize that the anion charge appears on 3 atoms only. Having 6 electrons each, these 3 atoms form only 2, not 3, two-electron bonds. Can we predict the cluster shape?
3.2 The 8−N rule expanded towards electron-precise clusters of p-element atoms
An electron-precise cluster of p-atoms is bonded by two-electron bonds and fulfils both the 8−N rule and octet. We can verify the expected cluster bonding by the 8−N rule expressed per the cluster, not per individual atom. Considering the 8−N rule at 1 atom: N BO = 8 − VEC, where VEC is per cluster atom, we convert it to the expression valid for n cluster atoms along this line of thoughts:
8 becomes 8n.
The valence-electron count becomes the total sum of electrons, g.
Each N BO in the «at» rule has counted for two atoms, now it doesn’t; divide by 2.
This formula is used in 1988 by Owen [23] and listed in Ref. [24] on p. 254. A simple example is P4 of 20 electrons. N BO = (32 − 20)/2 = 6 tetrahedron edges of this electron-precise cluster. For the above discussed As7 3− of 38 electrons, we get (56 − 38)/2 = 9 two-electron edges. Since this cluster must have its three As− merely two-bonded, we insert these As− as linkers into three adjoining edges of the As4 tetrahedron (to keep high symmetry). Also for the As4S4 cluster of two elements, the prediction (64 − 44)/2 = 10 bonds complies with the realgar [20] and pararealgar [21] structures.
3.3 The 18−N rule expanded for a cluster of d-element atoms
The derivation of the formula is now simple. Since the 18−N rule includes electrons donated by ligands, we count to g all electrons at the cluster and use 18n instead of 8n of the previous formula:
in Ref. [23], Ref. [24] p. 340 or [25] p. 139. Since n is any positive integer, this formula covers also a simple 18-plet complex of one single central atom. How to count the electron total g with the electrons donated to the cluster? The 18-plet electron counting is described in inorganic-chemistry textbooks, such as Ref. [4] p. 815. Let’s illustrate it on the nitroprusside anion [Fe(CN)5(NO)]2− by counting via neutral atoms. We evaluate how many electrons the ligands, taken as “neutral atoms”, provide to the central atom or the central cluster of atoms. The count for each ligand is a short heuristic algorithm:
For the cyanide ligand, we start with neutral uncharged •C≡N| that has to take 1e to form the |C≡N|− that donates 2e; hence 1 electron in total is donated.
The neutral • N=O| gives 1e to become |N≡O|+ that donates 2e to a bond along straight Fe, N, O line; hence 3 electrons are donated.
The count for [Fe(CN)5(NO)]2− is 8 electrons of Fe, 5 from cyanide ligands, 3 from nitrosyl, and 2 from the anion charge; an 18-plet at Fe.
For a simple molecular cluster, such as Os4(CO)14, where Os atoms have 32 valence electrons, and additional 28 are supplied by the neutral ligands CO, the formula yields N BO = (72 − 60)/2 = 6. These are edges of the Os4 tetrahedron bonded by single bonds. The cluster is electron-precise. However, the 18-plet rule is weaker than the octet, and is not necessarily always fulfilled in transition-metal compounds, in particular those of the early transition metals.
3.4 Wade rules for boranes (heuristic algorithm of several steps)
These rules [26] concern deltahedral boron clusters of at least 5 vertices. How to count electrons in the B6H6 2− anion? Well, how do we count C6H6? It has 30 electrons, hence 15 pairs. Of these, 6 must bond H and 6 must bond C. That leaves us with 3 pairs that comply with the Hückel rule [27] of odd number of itinerant π-electron pairs in aromatic flat rings. What is the counting algorithm for the B6H6 2− cluster?
Total valence electrons: 26
Total electron pairs: 13
Subtracting 1 pair per each B (to bond hydrogen) yields the number of skeletal pairs: 7
Of which 1 always is the radial bond, leaving n vertices for the underlying deltahedron shape: 6
If the cluster has n atoms, it is termed closo (no vertex missing in the deltahedron)
If 1 atom/vertex is missing, the cluster is termed nido.
If 2 atoms/vertices are missing, the cluster is termed arachno.
Let’s try B5H11 as another example: Valence electrons = 26. Electron pairs 13. Skeletal pairs 13 − 5 = 8. Underlying deltahedron vertices 8 − 1 = 7 of a pentagonal bipyramid. Actual boron vertices 5. Two vertices are missing, an arachno cluster.
We can even count the network of B12 clusters of the crystalline element itself. B12 has 36 valence electrons, 18 pairs per cluster. Now backwards: B12 has 12 vertices hence 13 skeletal pairs. Five pairs are thus left to bond the B12 clusters together into the 3D network. Of these 10 electrons per cluster, 6 are shared in 6 single bonds to another cluster. Remaining six borons of B12 contribute each with 2/3 of an electron (hence with 4 in total) to form six 2-electron 3-center bonds to other clusters.
3.5 Counting in networks of extended structures
The terms extended structure/solid [28] refer in a wider context to non-molecular solids, more narrowly to those containing 1D, 2D or 3D entities formed by directional bonds. In extended structures of binary compounds, the typical covalent single chain is the sp 3 chain. The variety of plain covalent atomic sheets is higher, but still rather limited. Layers of networks have a higher variety. If we include into 3D networks also ionic solids largely controlled by sphere packing, the variety increases substantially. And so on for compounds of three and more atoms. Are there any rules for, if not predict, then to verify or at least visualize such bonding patterns in crystalline solids?
One of them is the crystal-chemical formula with three balances behind it [29]. The chemical elements appear in the formula according to the symmetry of their local environments (sites) in the crystal structure. If an element occurs in two different bonding environments, it has two entries in the formula. Let’s take a binary compound of regular coordination polyhedra; TiO2 mineral rutile of oxygen octahedra around Ti and of titanium triangles around O. The crystal chemical formula Ti4+ 1 [6]O2− 2 [3] with each element at own crystallographic site having its oxidation state in superscript (as if ions) followed by the coordination number in square bracket.[1] The balances on Ti4+ 1 [6]O2− 2 [3] are:
Electroneutrality 1 × 4 = 2 × 2.
Connectivity 1 × 6 = 2 × 3. The number of Ti–O and O–Ti connections must be the same.
The bond-valence balance is 4/6 = 2/3 introduced as the “strength of the electrostatic valence bond” in the Pauling’s 2nd rule [30]. The quantity 2/3 per connection is called bond valence [31] or bond order. Three such bonds at oxygen sum to 2, six at titanium to 4. That’s the octet rule in ionic approximation; oxygen achieves the neon shell, titanium obtains the argon shell.
3.6 Niggli formula, connectivity matrix and bond graph
The parsimony rule of Pauling [30] (the 5th rule) states that “the number of essentially different kinds of constituents in a crystal tends to be small”. Let’s take an example of the B2O3 glass-former that only reluctantly [32] crystallizes into an open 3D network of BO3/2 triangles [33]. The latter formula per central atom of a coordination polyhedron is called Niggli formula and reads as “B O 3 two-connected”. Only brute force makes B2O3 to crystallize into a network of corner-sharing tetrahedra [34]. What’s the price? Two differently bonded oxygens. Trying to set up a BO1.5 Niggli formula for a BO4/n tetrahedron, we get no integer n solution. With two different O atoms, there are three alternatives: BO1/1+3/6 (1O one-connected, 3O six-connected), BO2/2+2/4 (2O two-connected, 2O four-connected), and BO1/2+3/3 (1O two-connected, 3O three-connected). Parsimony rule tells us which of them has a chance. We take 2 BO1/2+3/3 rewrite it as B2O(1)1O(2)2 with the two non-equivalent oxygen atoms and derive a simplified crystal-chemical formula B2 [1,3]O(1)[2]O(2)2 [3] from the connectivity balance for each boron connected to 1 O(1) and 3 O(2) in its square bracket. Complete with ionic charges, the B3+ 2 [1,3] O(1)2−[2]O(2)2− 2 [3] formula yields bond orders of these closest contacts. The O(1) atom has 2 bonds of order |−2|/2 = 1, and the O(2) atom has 3 bonds of order |−2|/3 = 2/3. The boron atom has 1 bond of bond order 1 and 3 bonds of bond order 2/3 with bond-valence sum 3 = 1 × 1 + 3 × ⅔. Drawing a bond graph [35, 36] in Fig. 1 helps visually. We distribute the atoms of the crystal-chemical formula, connect them by one line for each bonding interaction, and add bond order for each line.

Bond graph for B2O3 of boron-coordination tetrahedra with two types of O atoms as vertices. Bond-orders in blue.
The bond-valence balances can be formalized via a connectivity matrix [37] of “cation” rows versus “anion” columns for crystallographically different atoms.[2] For B2O3, the 1 × 2 matrix of connectivities in Table 1 is appended with the “total BVS” row and column that give the bond-valence sum for the cation total and for each of the two anion totals. These “total BVS” yield one equation for each such atom, but, because of the overall electroneutrality condition, one of these equations is a linear combination of the other two. When the matrix has a single column or row, we have enough equations to solve all unknown bond valences. Otherwise, the missing equation can be replaced by an estimate based on the parsimony principle in the “equal valence rule” [31] stating [35] that “the valence sum around any loop in the bond graph is zero”. It yields an auxiliary equation along a single-connection C–A–C–A– loop where the four unknown bond valences of signs +−+− sum to 0, such as the known ones do (⅔−⅔+1−1 = 0) in Fig. 1.
Connectivity matrix for B2O3 of corner-sharing tetrahedra with added “total BVS” equal to absolute value of the ionic charge total per crystal-chemical formula B3+ 2 [1,3] O(1)2−[2]O(2)2− 2 [3] and with bond valence v BO of the two different B–O bonds.
O2−(1) | 2 O2−(2) | Total BVS | ||
|
||||
2 B
3+
|
2 v
BO(1)
|
6 v
BO(2)
|
6 | |
Total BVS | 2 | 4 | ||
v BO = | 1 | 2/3 |
4 Algorithms for dimensionless quantities that characterize bonding
4.1 Bond order
Bond order is a quantity that gives the number of two-electron bonds connecting two atoms, integer or fractional. It has to be calculated for each bond by a suitable algorithm. In this case, several easy-to-grasp algorithms exist with simple starting parameters for the two bonded elements: For binary molecules, we draw the molecular-orbital (MO) scheme where the bond order = (bonding electrons − antibonding electrons)/2. A more general formulation is in the Gold Book. Alternatively, we draw a Lewis formula according to rules (8−N rule applied by electronegativity, octet, etc.) while counting electrons to obtain bond orders as integers or simple fractions. Bond order can also be calculated as the so-called bond valence from the bond length between the two atoms. A heuristic approach is based [38] on their Allred–Rochow electronegativities, the difference of which correlates with the covalence-caused bond shortening versus the sum of the two “ionic” radii that are fit by least squares to many bond lengths of known bond order between these two atoms. The electronegativity and ionic radius for the two atoms are the parameters needed to calculate their single-bond length via equation given in Ref. [38]. That single-bond length is the only parameter in the empirical exponential relation between the bond valence (bond order) and bond length, which has a long history between [39, 40] of details recapitulated in [41].
As an example, the parameters tabulated in Ref. [38] yield the B–O single-bond length R BO = 1.38 Å. That is one of four boron bonds in the tetrahedral network of the high-pressure B2O3 crystal [34]. The length of the three bonds with v BO = 2/3 follows from the bond-valence formula in Ref. [38] with the empirical constant 0.37 Å; d BO = R BO − 0.37ln(v BO) = 1.53 Å. The experimental values are 1.37 Å and 1.51 Å [34]. An alternative bond-valence approach [40] lists directly the empirical bond-valence parameters R ij for oxides, fluorides, etc., of elements. The R BO = 1.371 Å in [40, 41] yields 1.52 Å for the 2/3 bond via the same equation.
4.2 Oxidation state
Oxidation state is a quantity that characterizes bonding of an atom in a compound or ion. Based on ionic approximation of bonds, oxidation state of atoms in a chemical formula is typically counted by summing known ionic charges towards electroneutrality or the given charge of the chemical formula. At times, it might be useful or pedagogical to check this by a formal process.
As a dimensionless quantity derived from unit-less counts, oxidation state needs a concept [42], a verbal explanatory definition and a heuristic algorithm to calculate its value [43]. The definition [43] states: “The oxidation state of an atom is the charge of this atom after ionic approximation of its heteronuclear bonds.” Besides the direct ionic approximation limited to simple formulas (Ref. [9] p. 1025), two general alternative algorithms [43], subject to the same caveat, give the oxidation state. One algorithm uses Lewis formula, the other works also on bond graphs [35] of extended solids:
Algorithm 1:
“Oxidation state equals the charge of an atom after its homonuclear bonds have been divided equally and heteronuclear bonds assigned to the bond partners according to Allen electronegativity, except when the electronegative atom is bonded reversibly as a Lewis-acid ligand, in which case it does not obtain that bond’s electrons”.[3]
Algorithm 2:
“Heteronuclear-bond orders are summed at the atom as positive if that atom is the electropositive partner in a particular bond, and negative if it is not; the atom’s formal charge (if any) is added to that sum, yielding the oxidation state”. The same caveat as above applies also here.
Both algorithms are heuristic and start by drawing the Lewis formula of a molecule or a bond graph for an extended structure of a crystal. Let’s apply them on simple non-trivial examples. In trisulfane, H2S3, the Algorithm 1 assigns the two end bonds in the Lewis formula
Fig. 2 illustrates a Mn2(CO)10 example of homonuclear bonding on a Lewis formula of all valence electrons. It is set up from five electron-pair donor |C≡O| ligands at each manganese acceptor that obtains 18-plet upon pairing its odd electron with the other Mn. This results into 5 + 1 = 6 bonding and 3 essentially nonbonding pairs at each Mn. The N BO = (18n − g)/2 = (36 − 34)/2 = 1 bond.
Algorithm 1 performed with a pencil on the formula assigns the ligand-bonding pairs back to CO and leaves each Mn with its 6 lone-pair electrons and 1 electron from splitting the homonuclear Mn–Mn bond. The Mn oxidation state is 0, being the Mn “charge” counted as the 7 valence electrons of elemental manganese minus those 6 + 1 electrons the Mn atom has after ionic extrapolation.[4]
Algorithm 2 sums heteronuclear bond orders at Mn with positive sign, giving +5 to be summed with the formal charge of Mn in the Lewis formula in Fig. 2: Cutting all Mn bonds in half yields 12 electrons, formal charge 7 − 12 = −5, which yields oxidation state +5−5 = 0. What compensates the formal negative charge on Mn in this formula? Each oxygen has a formal charge 1+ equal to 6 − 5, obtained likewise.
Mn2(CO)10 has oxidation state 0 at Mn, which complies with the thermal decomposition [44] being a non-redox reaction into the electron-pair donor CO and Mn acceptor: Mn2(CO)10 = 2 Mn + 10 CO. Some similar carbonyls even form along the inverse path. Chemistry is about reactions.
One may wonder: Is the formula of Fig. 2 adequate to obtain oxidation state, despite its bonding approximations? It ignores the π-backdonation or backbonding that increases the Mn–C single-bond order by about 0.3 upon decreasing by similar amount the already <3 bond order in CO. This turns out not to be a problem; any backdonation assigns back to Mn under the ionic extrapolation to obtain the oxidation state, due to the caveat of low-electronegativity donor Mn and reversibly bonded acceptor CO. Another approximation might be the single Mn–Mn bond for an interaction weakened by repulsion of ligands and of the non-bonding pairs. Yet two electrons, one from each Mn, still mediate that bond.
A similar complex, yet with heteronuclear bonds, is in Fig. 3. The neutral-atom counting of electrons at the central “cluster” considers C5H5 that takes from it 1 electron to become the C5H5 − ligand that (weakly) shares its 6 aromatic electrons with Ir. These 5 electrons sum with 9 electrons of Ir and 4 electrons from two CO ligands to an 18-plet. Five CO donate 10 electrons to the 6 of W that obtains the 2 electrons missing to 18-plet from Ir. Upon ionic approximation, that bond assigns to Ir of higher Allen electronegativity, hence oxidation state of Ir is +1 and of W zero, as illustrated in Fig. 3.
The 18-plet is not always reached, as in the Re2Cl8 2− anion of quadruple bond [45]. The Re2Cl8 2− formula yields the +3 Re oxidation state, since 2 Re3+ sum with 8 Cl− to the anion charge (the direct ionic approximation in Ref. [9] p. 1025). Will the oxidation-state algorithms provide the same value? Counting by neutral atoms means that each of 8 neutral Cl-atoms must obtain 1 electron in order to donate a pair, giving 1 electron to the Re2 cluster. These 8 donated electrons sum up with 2 of the anionic charge and 14 electrons of the two Re atoms. In total (8 + 2 + 14)/2 = 12 electron pairs are to be drawn at the cluster in the Lewis formula (Fig. 4). It means that Re has mere 16-plet. Algorithm 1 assigns all bonds to Cl and leaves each Re with half of the homonuclear-bond electrons, yielding oxidation state 7 − 4 = +3 on Re. Algorithm 2 sums Re–Cl bonds with + sign at Re and adds the Re formal charge, 4 − 1 = +3 on Re.
Another example is the GeS corrugated sheet [46] of black-phosphorus type with three-connected atoms (Fig. 5). As noted earlier, both atoms have octet of three pairs bonding and one non-bonding. Formal charges, after dividing all bonds in half, are +1 on S and −1 on Ge. Algorithm 1 assigns all bonds to S, yielding oxidation state 6 − 8 = −2, while Ge of 1 lone pair has oxidation state 4 − 2 = +2. Algorithm 2 assigns the three single bonds to Ge with + sign and to S with − sign and adds the respective −1 and + 1 formal charges to yield the +2 and −2 oxidation states expected by direct ionic approximation from the GeS formula but perhaps less so from the structure.
In contrast, the formula As4S4 of pararealgar [21] does not yield all oxidation states correct by the direct ionic approximation that worked above. While all atoms keep octet and the 8−N rule, there are 3 different As atoms (Fig. 6). One bonds 3 S and has oxidation state +3 by Algorithm 1, another two bond 2 S each and have oxidation state +2, and one bonds 1 S and has oxidation state +1. These values are not far from 2.91, 2.02, and 1.14 obtained by Algorithm 2 with bond orders calculated via the bond-valence approach of Ref. [38] from bond lengths determined in Ref. [21].

The Mn2(CO)10 molecule and Lewis formula of all valence electrons to derive the needed quantities.

Lewis formula of (CO)2(η 5-C5H5)Ir–W(CO)5 with all electrons at the iridium–tungsten cluster. Formal charge in black, oxidation state in red, the cyclopentadienyl (Cp) shares 3 aromatic electron pairs formed with 1 electron from Ir.

Lewis formula of the Re2Cl8 2− anion. Formal charge in black, oxidation state in red.

Corrugated sheet of GeS reminiscent of black phosphorus.

The As4S4 molecule in pararealgar.
5 Prediction of properties
5.1 How to alloy a ferromagnet?
Iron, cobalt, and nickel are ferromagnetic. Their orbitals of d-parentage are stabilized by full filling in one spin direction while the s orbital is occupied partially in both spin directions. The other d-orbital spin has fewer electrons, and the excess of the majority spin is behind the ferromagnetism. Alloys of metals around this range may become ferromagnetic when the average per atom of their outer-shell d + s + p electrons is close to that of Fe, Co, or Ni.
A simple generalized picture is the Pauling’s [47] nearly equilateral triangle for saturated magnetic moments per atom (in Bohr-magneton units μB, about equal to the effect of 1 unpaired electron) as a function of valence-electron count per atom in binary alloys from Cr to Ni. The triangle sides rise from zero close to the chromium count of 6 valence electrons (all used in bonds or itinerant[5]) and from zero some ½ electron above the Ni valence-electron count 10. The triangle vertex is at saturated magnetic moment of about 2 µB close to the valence-electron count of Fe. Co has one unpaired electron less. For the saturated ferromagnetic moment μ sat of alloys along the Cr–Fe edge of the triangle, a simple per-atom rule μ sat/µB = N − 6 is known as the Slater–Pauling rule [47, 48], where N is the average valence-electron count. It works for many binary and ternary intermetallics, as can be seen from the Pauling-triangle updates in Fig. 1 of Ref. [49], Fig. 12 of Ref. [50], and Fig. 24 of Ref. [51].
Ferromagnetism of an intermetallic compound and its saturated magnetic moment depend not only on elemental composition but also on bonding. Crystal structures and molecular orbitals extended into bands become important. The general rule may need adjustments for some structure types. Heusler alloys are a group of intermetallic compounds that are often ferromagnetic:
Compositions XYZ denote half-Heusler alloys. Those relevant for ferromagnetism contain a late transition metal X, an early transition metal Y, and a p-metalloid Z. The late transition metal X has a cube of 8 nearest YZ neighbors that each have a tetrahedron of nearest neighbors X. One choice of the unit-cell origin starts with the cubic closest packing of Z with octahedral voids occupied by the early transition metal Y as in a NaCl-type cell. The late transition metal X centers four unit-cell octants, the tetrahedron of which yields the F
Full Heusler alloys can be regular or inverse. Regular Heusler alloys have composition X2YZ that stems from the above XYZ structure by the late transition metal X completing the tetrahedron of X into a cube that leads to the Fm
Keeping the symbol X for late and Y for early transition metal, inverse Heusler alloys have formula Y2XZ, where the X tetrahedron is completed by Y into an XY cube keeping the F
When Y = Sc, Ti in Y2XZ, (at times also V), only one d-orbital based MO set is doubly occupied, the Fermi level is lowest. The counting is then as with the half-Heusler alloys: The μ sat total/µB = N total − 18.
When X = Cu or Zn in Y2XZ, the entire additional d-orbital MO set would be doubly occupied (not just the three t 2 orbitals), the Fermi level is highest. There are N total − 8(sp 3) − 10(d 5) − 10(d 5) unpaired electrons. The μ sat total/µB = N total − 28 rule is valid [54, 55] for these predicted compounds.
Apart from the above two points, inverse Heusler alloys have μ sat total/µB = N total − 24.
The algorithms would be:
For a d-metal based half-Heusler alloy XYZ with Z = p-group atom, calculate the total number of valence electrons N total with valence electrons for the X and Y d-group metal, and only sp-electrons for the Z atom.
Estimate the total ferromagnetic moment μ sat total in Bohr magnetons as N total − 18.
For N total = 18, the alloy is not ferromagnetic.
For a transition-metal full-Heusler alloy with a p-group atom, calculate the total number of valence electrons N total with d-electrons for transition metals and sp-electrons for the p-group atom.
If the formula has Sc2 or Ti2 (rarely V2), estimate μ sat total in Bohr magnetons as N total − 18.
If the formula has Cu1 or Zn1 (possibly also Ag1, Au1, Cd1, Hg1; not tested) estimate μ sat total in Bohr magnetons as N total − 28.
Otherwise, estimate μ sat total in Bohr magnetons as N total − 24.
These estimated values often agree with quantum-chemical calculations, yet less often with the published magnetic measurements. One of the reasons is that the actual crystal structure deviates from the expected ideal, often by a minor disorder or intermixing. In particular for half-Heusler alloys, their ab-initio calculated values [56] are often contradicted by experiment, such as for CoMnSb of an ordered superstructure [57] and saturated moment close to 4 µB [57], [58], [59] instead of 3 µB [49] obtained by calculation for the ideal structure. Table 2 lists examples of half-Heusler alloys with valence-electron counts that agree with measurements on samples of checked and determined crystal structure.
Predictions of saturated ferromagnetic moments in half-Heusler alloys compared with the value measured in an experiment (exp.) or computed by first-principle calculations (calc.). N = valence-electron count, PM = paramagnetic.
Formula | N total | μ sat total/μB | exp. | Ref. | calc. | Ref. | cr. str. |
---|---|---|---|---|---|---|---|
NiMnSb | 22 | 22−18 = 4 | 4.04 | [60] | 3.96 | [61] | [60] |
NiTiSn | 18 | 18−18 = 0 | PM | [62] | [62] | ||
PtTiGe | 18 | 18−18 = 0 | PM | [63] | [63] |
Examples of predicted and observed saturated moments of regular full-Heusler alloys are in Table 3. Data for inverse Heusler alloys appear much less frequently in the literature (Table 4), and the magnetic moments are often just calculated. For majority of cases, the estimated saturated ferromagnetic moments per atom follow the Slater–Pauling rule N average−6, which means they coincide with the left edge of the Pauling’s triangle as in Ref. [51]. The two valence-electron count extremes towards the top and bottom of Table 4 for inverse Heusler alloys would form lines parallel with that triangle edge.
Predictions of saturated ferromagnetic moments in regular Heusler alloys compared with the value measured in an experiment (exp.) or computed by first-principle calculations (calc.). N = valence-electron count, PM = paramagnetic.
Formula | N aver. | N total | μ sat total/μB | exp. | Ref. | calc. | Ref. | cr. str. |
---|---|---|---|---|---|---|---|---|
Co2FeSi | 7.50 | 30 | 30−24 = 6 | 5.97 | [64] | 5.27 | [53] | [64] |
Co2MnSi | 7.25 | 29 | 29−24 = 5 | 5.07(5) | [65] | 4.94 | [53] | [65] |
Co2MnAl | 7.00 | 28 | 28−24 = 4 | 4.01(5) | [65] | 3.97 | [53] | [65] |
Co2CrGa | 6.75 | 27 | 27−24 = 3 | 3.01 | [66] | 3.01 | [66] | [66] |
Co2VGa | 6.50 | 26 | 26−24 = 2 | 1.95(6) | [67] | [67] | ||
Fe2CrAl | 6.25 | 25 | 25−24 = 1 | 1.67 1.6 | [68, 69] | 0.91 | [53] | [68, 69] |
Fe2VAl | 6.00 | 24 | 24−24 = 0 | PM | [58] |
Predictions of saturated ferromagnetic moments in inverse Heusler alloys compared with the value measured in an experiment (exp.) or computed by first-principle calculations (calc.). N = valence-electron count, PM = paramagnetic.
Formula | N aver. | N total | μ sat total/μB | exp. | Ref. | calc. | Ref. | cr. str. |
---|---|---|---|---|---|---|---|---|
Mn2ZnSi | 7.50 | 30 | 30−28 = +2 | +2 | [70] | |||
Mn2CuSi | 7.25 | 29 | 29−28 = +1 | +1 | [71] | |||
Cr2ZnSi | 7.00 | 28 | 28−28 = 0 | 0 | [72] | |||
Mn2CoSb | 7.00 | 28 | 28−24 = +4 | +3.92 | [73] | +3.99 | [73] | [73] |
Mn2CoGe | 6.75 | 27 | 27−24 = +3 | +2.99 | [73] | +2.99 | [73] | [73] |
Mn2CoAl | 6.50 | 26 | 26−24 = +2 | +1.95 | [73] | +2 | [73] | [73] |
Cr2CoSi | 6.25 | 25 | 25−24 = +1 | +1.06 | [54] | |||
Cr2CoAl | 6 | 24 | 24−24 = 0 | PM | [74] | +0.01 | [54] | [74] |
Cr2FeAl | 5.75 | 23 | 23−24 = −1 | −1.01 | [54] | |||
V2CoAl | 5.50 | 22 | 22−24 = −2 | −2.00 | [54] | |||
Cr2CrAl | 5.25 | 21 | 21−24 = −3 | −2.93 | [54] | |||
Ti2NiAl | 5.25 | 21 | 21−18 = +3 | +3.00 | [54] | |||
Ti2CoAl | 5.00 | 20 | 20−18 = +2 | +2.00 | [54] | |||
Ti2FeAl | 4.75 | 19 | 19−18 = +1 | +0.5 | [75] | +1.00 | [54] | [75] |
Ti2MnAl | 4.50 | 18 | 18−18 = 0 | PM | [75] | 0 | [54] | [75] |
Ti2VSi | 4.25 | 17 | 17−18 = −1 | −1.00 | [54] | |||
Sc2VAs | 4.00 | 16 | 16−18 = −2 | −2.06 | [54] | |||
Sc2VSi | 3.75 | 15 | 15−18 = −3 | −3.00 | [54] |
5.2 Suggest a high-T c superconductor
High-T c superconductivity in a periodic network solid is associated with a central-atom behind a largely nonbonding HOMO that is half-occupied and separated in energy by suitable ligand field (approximates the narrow nearly half-filled band advocated by physicists [76]). A diluted hole pair can be imagined if this orbital is slightly less than half filled, and diluted electron pair if slightly more than half filled. These become the superconducting pairs, perhaps by condensation [77] of the Bose–Einstein type.
An example is YBa2Cu3O7−x of two Cu-coordination square pyramids joined by square-planar copper coordination that accommodates the x oxygen vacancies per unit cell. Already the YBa2Cu3O6.5 composition of average Cu2+ (d 9) is superconducting because the square-pyramidal coordination of 5 oxygens keeps Cu>2+ with a single nearly half-filled orbital d x 2 −y 2 [78], while the x-vacant square-planar coordination has Cu<2+ on average. Removing the x-vacancies stabilizes the superconductivity. Under aliovalent substitutions, the critical temperature correlates with the bond-valence sum at Cu in the coordination pyramids where the hole pairs form, not with an integer oxygen content or disappearing distortions [79, 80]. Another example of hole-doped d 9 configuration is the Nd0.8Sr0.2NiO2 superconductor [81]. An example of electron doping is Nd2−x Ce x CuO4 superconducting below 24 K when x = 0.15 [82].
A similar case is the nearly half-occupied relativistic-stabilized 6s 2 orbital. It leads to superconductivity up to 13 K in BaPb1−x Bi x O3 [83], increased to 30 K in Ba0.6K0.4BiO3 [84]. In the Tl–Ba–Cu–O system, zero resistance at 81 K [85] was raised to 125 K [86] and 127 K [87]. HgBa2CuO4+δ becomes superconductor below 94 K [88], and HgBa2Ca2Cu3O8+δ [89] holds the record of zero resistivity at 134 K.
In 2001, it was discovered that MgB2 is a superconductor below 39 K [90]. Taken as a Zintl phase of VEC A = 4, its boron forms 8−4 = 4 bonds in graphene-type sheets separated by sheets of Mg2+ that is placed above and below each hexagon. However, the charge transfer from the cation is not entirely perfect; MgB2 is black, B–B distances are longer than expected, and calculations suggest an s→p charge transfer with superconductivity driven by s-band holes [91]. Similar cases are potassium [92], rubidium [93], or cesium [94] alkali-metal fullerides A3C60 with cations A filling all octahedral and tetrahedral holes in the cubic closest packing of the anion balls [95]. Cs3C60 is superconducting up to 33 K.
In 2006, superconductivity below 4 K in a LaOFeP layered network was reported [96]. The critical temperature increased to 26 K [97] upon electron-doping the

Bonds in the LiFeAs unit-cell repeat unit of the infinite iron cluster (in green) with magenta bonds to As ligands (large white) that are charged by lithium cations (small white). Cluster counting (see text) suggests ideally one lone pair per Fe.
The idea of nearly half-filled orbital/band, with implicit use behind hundreds of tried superconductor compositions, lends some heuristic quality to the rules considered in this section. An old empirical rule for superconductivity is worth mentioning; the 1955 Matthias rule [105] that there is a dip in d-metal alloy superconductivity around 6 valence electrons per atom (that are all unpaired in an isolated atom) with maxima around d 5 and d 7. The rule covers a wide range of d-metal alloys that may also contain pre- and post-transitional elements and form a wide range of structure types including Heusler alloys [106] as well as high-entropy alloys [107]. Some studies [108] see common denominators behind this rule and the high-T c superconductivity in theories that go beyond heuristic approach.
6 Prediction of reactions
6.1 Redox reactions in aqueous solutions
Chemists balance redox equations by counting exchanged electrons in a simple algorithm: The per-formula decrease in oxidation state of the oxidant-element atoms (the electron total they accept upon reduction) becomes the reaction coefficient in front of the reductant. The inverse for electrons the reductant formula gives away. Ionic charges are then balanced via a protic acid or base if needed. How to find useful reductant for a given oxidant, and vice versa? By considering oxidation and reduction separately. Every redox reaction is a sum of two half-reactions, one for oxidation upon losing electrons and one for reduction upon accepting them. In 1938, Wendell Latimer published “The Oxidation States of the Elements and their Potentials in Aqueous Solutions” [42]. Since then, we have tables of standard electrochemical reduction potentials E° for half-reactions of inorganic compounds or ions in their 1M (1 mol/L) aqueous solutions.
The redox-prediction algorithm is simple: Take the two half-reactions, invert the oxidation half-reaction so that it too refers to reduction, and look up the two standard reduction potentials. The half-reaction with higher standard reduction potential provides the oxidant to be reduced. The half-reaction with lower standard reduction potential provides the reductant to be oxidized. The oxidant and reductant then react with each other.
Reduction potentials concern aqueous solutions and tell us something about the chemistry of the element. Even the solid-state redox behavior may roughly follow these aqueous redox estimates. Standard reduction potentials are typically listed for pH 0 (
Let’s illustrate it on the HClO and Cl2 pair. At pH 0, their half-reaction equation includes H+ ions that balance the electron charge, HClO + e− + H+ = ½Cl2 + H2O. Its E° under varying pH beyond the standard pH 0 is expressed from the simplified 298 K Nernst equation while keeping the unity standard pressure of Cl2:
The amount concentration (molar) of H+ ions equals 10−pH, and the concentration of HClO molecules changes with pH as well. How it changes, we find out via the equilibrium constant of the protolysis reaction:
Since [HClO] + [ClO−] = 1 in the standard 1M solution, the unknown [ClO−] = 1 − [HClO] is plugged into the K a expression, yielding the [HClO] = [H+]/(K a + [H+]) needed in our adopted form of Nernst equation:
where we take [H+] = 10−pH. Fig. 8 shows the plot, calculated with
![Fig. 8:
Standard potential E° (in V) for reduction of HClO or ClO− to Cl2 as a function of pH ([HClO]+[ClO−] = 1 mol/L, p
Cl2
= 1 bar). Notice the reaction equation suggesting the double slope in alkaline solutions.](/document/doi/10.1515/pac-2022-1118/asset/graphic/j_pac-2022-1118_fig_008.jpg)
Standard potential E° (in V) for reduction of HClO or ClO− to Cl2 as a function of pH ([HClO]+[ClO−] = 1 mol/L, p Cl2 = 1 bar). Notice the reaction equation suggesting the double slope in alkaline solutions.
Let’s now bubble Cl2 gas into water of varied pH (by a redox-prone acid/base). One product will certainly be chloride ion. The reduction half-reaction Cl2 + 2e− = 2Cl− does not involve water ions; hence its standard potential
6.2 Products of carbide hydrolysis
Carbides contain C1 or C2 or C3 carbon groups. Some carbides hydrolyze in water or protic acids to the corresponding hydrocarbon; some form hydrocarbon mixtures. Salt-like carbides of electropositive sp-metals contain C4− or C2 2− or C3 4− anions and hydrolyze into the respective weak acid CH4 or C2H2 or C3H4. Some carbides of d- or f-transition metals are hydrolyzable, such as Sc2OC and Mn7C3 of single carbon atoms, or YC2 and other rare-earth dicarbides of C2 groups. Whereas Sc2OC and its rare-earth analogues hydrolyze to methane [111], Mn7C3 yields a hydrocarbon array of concentration decreasing with the chain length [112]. So does YC2, except that odd-numbered chains are missing.
To predict the outcome, we identify the carbon groups of the carbide and then use standard reduction potentials to predict the metal ion stable in the acidic aqueous environment. Balancing the primary hydrolysis-reaction gives the answer:
Sc2OC would form Sc3+ as Sc2OC + 6 H+ = 2 Sc3+ + CH4 + H2O.
Mn7C3 would form Mn2+ as Mn7C3 + 14 H+ = 7 Mn2+ + 3 CH4 + H2.
YC2 would form Y3+ as YC2 + 3 H+ = Y3+ + C2H2 + ½ H2.
The algorithm is simple: Those carbides that yield hydrocarbon series in the real reaction, release also hydrogen. Correspondingly, Mn7C3 is silvery and YC2 golden metallic.
How does hydrogen polymerize the carbon groups? In the reaction on the solid surface, one hydrated H+ ion protonates the carbon group towards the corresponding hydrocarbon while the metal reduces another H+ into the “nascent hydrogen” •H that interferes with the protonation in the vicinity by initializing polymerization of hydrocarbon radicals in n steps [112] into hydrocarbon chains of C1n or C2n . Hence, if the carbide composition suggests formation of hydrogen upon hydrolysis, an array of hydrocarbons forms as if by polymerizing the structural carbon groups present in the solid [113].
7 Conclusions
Heuristic algorithms are important because they facilitate understanding of chemistry via simple quantitative concepts. Every reader would surely come with additional quantity-based heuristic algorithms in chemistry. The purpose of this review was to present examples of such approaches in inorganic chemistry and to illustrate their usefulness in teaching and understanding it. One then also better appreciates solutions of more complicated tasks by ab-initio computations, while remembering that also these use input parameters the applicability of which needs to be considered.
References
[1] G. N. Lewis. J. Am. Chem. Soc. 38, 762 (1916), https://doi.org/10.1021/ja02261a002.Search in Google Scholar
[2] W. Kossel. Ann. Phys. 354, 229 (1916), https://doi.org/10.1002/andp.19163540302.Search in Google Scholar
[3] I. Langmuir. J. Am. Chem. Soc. 41, 868 (1919), https://doi.org/10.1021/ja02227a002.Search in Google Scholar
[4] C. E. Housecroft, A. G. Sharpe. Inorganic Chemistry, Pearson Education Limited, Harlow, England, 3rd ed. (2008).Search in Google Scholar
[5] Octet rule – Wikipedia, https://en.wikipedia.org/wiki/Octet_rule (accessed Apr 27, 2023).Search in Google Scholar
[6] I. Langmuir. Science 54, 59 (1921), https://doi.org/10.1126/science.54.1386.59.Search in Google Scholar PubMed
[7] P. Pyykkö. J. Organomet. Chem. 691, 4336 (2006), https://doi.org/10.1016/j.jorganchem.2006.01.064.Search in Google Scholar
[8] W. Hume-Rothery VIII. Philos. Mag. 9, 65 (1930), https://doi.org/10.1080/14786443008564982.Search in Google Scholar
[9] P. Karen, P. McArdle, J. Takats. Pure Appl. Chem. 86, 1017 (2014), https://doi.org/10.1515/pac-2013-0505.Search in Google Scholar
[10] D. E. Woon, T. H. DunningJr. J. Phys. Chem. A 113, 7915 (2009), https://doi.org/10.1021/jp901949b.Search in Google Scholar PubMed
[11] R. J. Martinie, J. J. Bultema, M. N. V. Wal, B. J. Nirkhart, D. A. V. Griend, R. L. DeKock. J. Chem. Educ. 88, 1094 (2011), https://doi.org/10.1021/ed100758t.Search in Google Scholar
[12] L. Pauling. The Nature of the Chemical Bond, Cornell University Press, Ithaca, NY, 3rd ed. (1960).Search in Google Scholar
[13] B. R. Miller, M. Fink. J. Chem. Phys. 83, 939 (1985), https://doi.org/10.1063/1.449420.Search in Google Scholar
[14] E. Zintl. Angew. Chem. 52, 1 (1939), https://doi.org/10.1002/ange.19390520102.Search in Google Scholar
[15] E. Parthé. Acta Crystallogr., Sect. B 29, 2808 (1973), https://doi.org/10.1107/S0567740873007570.Search in Google Scholar
[16] E. Parthé. Elements of Inorganic Structural Chemistry: A Course on Selected Topics, K. Sutter Parthé, Geneva, Switzerland (1990).Search in Google Scholar
[17] R. Nesper. Angew. Chem., Int. Ed. 640, 2639 (2014), https://doi.org/10.1002/zaac.201400403.Search in Google Scholar
[18] E. Parthé. Acta Crystallogr., Sect. B 36, 1 (1980), https://doi.org/10.1107/S0567740880002312.Search in Google Scholar
[19] J. H. Bryden. Acta Crystallogr. 15, 167 (1962), https://doi.org/10.1107/S0365110X62000407.Search in Google Scholar
[20] D. J. E. Mullen, W. Nowacki. Z. Kristallogr. 136, 48 (1972), https://doi.org/10.1524/zkri.1972.136.1-2.48.Search in Google Scholar
[21] P. Bonazzi, S. Menchetti, G. Pratesi. Am. Mineral. 80, 400 (1995), https://doi.org/10.2138/am-1995-3-422.Search in Google Scholar
[22] T. Wadsten. Acta Chem. Scand. 21, 593 (1967), https://doi.org/10.3891/acta.chem.scand.21-0593.Search in Google Scholar
[23] S. M. Owen. Polyhedron 7, 253 (1988), https://doi.org/10.1016/S0277-5387(00)80467-2.Search in Google Scholar
[24] D. M. P. Mingos. Essential Trends in Inorganic Chemistry, Oxford University Press, New York (1998).10.1093/hesc/9780198559184.001.0001Search in Google Scholar
[25] U. Müller. Inorganic Structural Chemistry, John Wiley & Sons Ltd, Chichester, England (2006).Search in Google Scholar
[26] K. Wade. J. Chem. Soc. D 1971, 792 (1971), https://doi.org/10.1039/C29710000792.Search in Google Scholar
[27] E. Hückel. Z. Phys. 70, 204 (1931), https://doi.org/10.1007/BF01339530.Search in Google Scholar
[28] T. Hughbanks, R. Hoffmann. J. Am. Chem. Soc. 105, 1150 (1983), https://doi.org/10.1021/ja00343a014.Search in Google Scholar
[29] J. Lima de Faria, E. Hellner, F. Liebau, E. Makovický, E. Parthé. Acta Crystallogr., Sect. A 46, 1 (1990), https://doi.org/10.1107/S0108767389008834.Search in Google Scholar
[30] L. Pauling. J. Am. Chem. Soc. 51, 1010 (1929), https://doi.org/10.1021/ja01379a006.Search in Google Scholar
[31] I. D. Brown. Acta Crystallogr. Sect. B 33, 1305 (1977), https://doi.org/10.1107/S0567740877005998.Search in Google Scholar
[32] M. J. Aziz, E. Nygren, J. F. Hays, D. Turnbull. J. Appl. Phys. 57, 2233 (1985), https://doi.org/10.1063/1.334368.Search in Google Scholar
[33] G. E. Gurr, P. W. Montgomery, C. D. Knutson, B. T. Gorres. Acta Crystallogr., Sect. B 26, 906 (1970), https://doi.org/10.1107/S0567740870003369.Search in Google Scholar
[34] C. T. Prewitt, R. D. Shannon. Acta Crystallogr., Sect. B 24, 869 (1968), https://doi.org/10.1107/S0567740868003304.Search in Google Scholar
[35] I. D. Brown. Z. Kristallogr. 199, 255 (1992), https://doi.org/10.1524/zkri.1992.199.3-4.255.Search in Google Scholar
[36] I. D. Brown. Acta Crystallogr., Sect. B 53, 381 (1997), https://doi.org/10.1107/S0108768197002474.Search in Google Scholar
[37] M. O’Keeffe. Acta Crystallogr., Sect. A 46, 138 (1990), https://doi.org/10.1107/S0108767389011104.Search in Google Scholar
[38] M. O’Keeffe, N. E. Brese. J. Am. Chem. Soc. 113, 3226 (1991), https://doi.org/10.1021/ja00009a002.Search in Google Scholar
[39] L. Pauling. J. Am. Chem. Soc. 69, 542 (1947), https://doi.org/10.1021/ja01195a024.Search in Google Scholar
[40] I. D. Brown, D. Altermatt. Acta Crystallogr., Sect. B 41, 244 (1985), https://doi.org/10.1107/S0108768185002063.Search in Google Scholar
[41] N. E. Brese, M. O’Keeffe. Acta Crystallogr. Sect. B 47, 192 (1991), https://doi.org/10.1107/S0108768190011041.Search in Google Scholar
[42] W. M. Latimer. The Oxidation States of the Elements and Their Potentials in Aqueous Solutions, Prentice Hall, Inc., New York, NY (1938).Search in Google Scholar
[43] P. Karen, P. McArdle, J. Takats. Pure Appl. Chem. 88, 831 (2016), https://doi.org/10.1515/pac-2015-1204.Search in Google Scholar
[44] X. Qin, H. Sun, F. Zaera. J. Vac. Sci. Technol. A 30, 01A112 (2021), https://doi.org/10.1116/1.3658373.Search in Google Scholar
[45] F. A. Cotton, C. B. Harris. Inorg. Chem. 4, 330 (1965), https://doi.org/10.1021/ic50025a015.Search in Google Scholar
[46] H. Wiedemeier, H. G. von Schnering. Z. Anorg. Allg. Chem. 431, 299 (1977), https://doi.org/10.1002/zaac.19774310134.Search in Google Scholar
[47] L. Pauling. Phys. Rev. 54, 899 (1938), https://doi.org/10.1103/PhysRev.54.899.Search in Google Scholar
[48] J. C. Slater. Phys. Rev. 49, 931 (1936), https://doi.org/10.1103/PhysRev.49.931.Search in Google Scholar
[49] J. Kübler. Phys. B+C 127, 257 (1984), https://doi.org/10.1016/S0378-4363(84)80039-X.Search in Google Scholar
[50] C. Felser, G. H. Fecher, B. Balke. Angew. Chem., Int. Ed. 46, 668 (2007), https://doi.org/10.1002/anie.200601815.Search in Google Scholar PubMed
[51] T. Graf, C. Felser, S. S. P. Parkin. Prog. Solid State Chem. 39, 1 (2011), https://doi.org/10.1016/j.progsolidstchem.2011.02.001.Search in Google Scholar
[52] H. C. Kandpal, C. Felser, R. Seshadri. J. Phys. D: Appl. Phys. 39, 776 (2006), https://doi.org/10.1088/0022-3727/39/5/S02.Search in Google Scholar
[53] I. Galanakis, P. H. Dederichs, N. Papanikolaou. Phys. Rev. B 66, 174429 (2002), https://doi.org/10.1103/PhysRevB.66.174429.Search in Google Scholar
[54] S. Skaftouros, K. Özdogan, E. Şaşıoğlu, I. Galanakis. Phys. Rev. B 87, 024420 (2013), https://doi.org/10.1103/PhysRevB.87.024420.Search in Google Scholar
[55] H. Luo, F. Meng, H. Liu, J. Li, E. Liu, G. Wu, X. Zhu, C. Jiang. J. Magn. Magn. Mater. 324, 2127 (2012), https://doi.org/10.1016/j.jmmm.2012.02.026.Search in Google Scholar
[56] I. Galanakis. Phys. Rev. B 71, 012413 (2005), https://doi.org/10.1103/PhysRevB.71.012413.Search in Google Scholar
[57] V. Ksenofontov, G. Melnyk, M. Wojcik, S. Wurmehl, K. Kroth, S. Reiman, P. Blaha, C. Felser. Phys. Rev. B 74, 134426 (2006), https://doi.org/10.1103/PhysRevB.74.134426.Search in Google Scholar
[58] K. H. J. Buschow, P. G. van Engen, R. Jongebreur. J. Magn. Magn. Mater. 38, 1 (1983), https://doi.org/10.1016/0304-8853(83)90097-5.Search in Google Scholar
[59] H. Nakamura, M. Y. P. Akbar, T. Tomita, A. A. Nugroho, S. Nakatsuji. JPS Conf. Proc. 29, 013004 (2020), https://doi.org/10.7566/JPSCP.29.013004.Search in Google Scholar
[60] P. J. Webster, R. M. Mankikar. J. Magn. Magn. Mater. 42, 300 (1984), https://doi.org/10.1016/0304-8853(84)90113-6.Search in Google Scholar
[61] I. Galanakis, P. H. Dederichs, N. Papanikolaou. Phys. Rev. B 66, 134428 (2002), https://doi.org/10.1103/PhysRevB.66.134428.Search in Google Scholar
[62] J. Pierre, R. V. Skolozdra, Y. K. Gorelenko, M. Kouacou. J. Magn. Magn. Mater. 134, 95 (1994), https://doi.org/10.1016/0304-8853(94)90078-7.Search in Google Scholar
[63] S.-V. Ackerbauer, A. Senyshyn, H. Borrmann, U. Burkhardt, A. Ormeci, H. Rosner, W. Schnelle, M. Gamża, R. Gumeniuk, R. Ramlau, E. Bischoff, J. C. Schuster, F. Weitzer, A. Leithe-Jasper, L. H. Tjeng, Yu. Grin. Chem. - Eur. J. 18, 6272 (2012), https://doi.org/10.1002/chem.201102401.Search in Google Scholar PubMed
[64] S. Wurmehl, G. H. Fecher, H. C. Kandpal, V. Ksenofontov, C. Felser, HJ Lin. Appl. Phys. Lett. 88, 032503 (2006), https://doi.org/10.1063/1.2166205.Search in Google Scholar
[65] P. J. Webster. J. Phys. Chem. Solids 32, 1221 (1971), https://doi.org/10.1016/S0022-3697(71)80180-4.Search in Google Scholar
[66] R. Y. Umetsu, K. Kobayashi, R. Kainuma, A. Fujita, K. Fukamichi. Appl. Phys. Lett. 85, 2011 (2004), https://doi.org/10.1063/1.1790029.Search in Google Scholar
[67] K. R. A. Ziebeck, P. J. Webster. J. Phys. Chem. Solids 35, 1 (1974), https://doi.org/10.1016/0022-3697(74)90002-X.Search in Google Scholar
[68] K. H. J. Buschow, P. G. van Engen. J. Magn. Magn. Mater. 25, 90 (1981), https://doi.org/10.1016/0304-8853(81)90151-7.Search in Google Scholar
[69] R. Y. Umetsu, N. Morimoto, M. Nagasako, R. Kainuma, T. Kanomata. J. Alloys Compd. 528, 34 (2012), https://doi.org/10.1016/j.jallcom.2012.02.163.Search in Google Scholar
[70] S. Kervan, N. Kervan. Curr. Appl. Phys. 13, 80 (2013), https://doi.org/10.1016/j.cap.2012.06.018.Search in Google Scholar
[71] S. V. Faleev, Y. Ferrante, J. Jeong, M. G. Samant, B. Jones, S. S. P. Parkin. Phys. Rev. B 95, 045140 (2017), https://doi.org/10.1103/PhysRevB.95.045140.Search in Google Scholar
[72] T. Yang, J. Youa, L Hao, R. Khenatab, Z.-Y. Wang, X. Wang. J. Magn. Magn. Mater. 498, 166188 (2020), https://doi.org/10.1016/j.jmmm.2019.166188.Search in Google Scholar
[73] G. D. Liu, X. F. Dai, H. Y. Liu, J. L. Chen, Y. X. Li, G. Xiao, G. H. Wu. Phys. Rev. B 77, 014424 (2008), https://doi.org/10.1103/PhysRevB.77.014424.Search in Google Scholar
[74] A. Datta, M. Modak, P. Paul, M. M. Patidar, T. V. C. Rao, V. Ganesan, S. Banerjee. J. Magn. Magn. Mater. 521, 167522 (2021), https://doi.org/10.1016/j.jmmm.2020.167522.Search in Google Scholar
[75] J. Goraus, J. Czerniewski. J. Magn. Magn. Mater. 498, 166106 (2020), https://doi.org/10.1016/j.jmmm.2019.166106.Search in Google Scholar
[76] E. Krüger. J. Supercond. Novel Magn. 23, 213 (2010), https://doi.org/10.1007/s10948-009-0518-1.Search in Google Scholar
[77] T. Hashimoto, Y. Ota, A. Tsuzuki, T. Nagashima, A Fukushima, S. Kasahara, Y. Matsuda, K. Matsura, Y. Mizukami, T. Shibauchi, S. Shin, K. Okazaki. Sci. Adv. 6, eabb9052 (2020), https://doi.org/10.1126/sciadv.abb9052.Search in Google Scholar PubMed PubMed Central
[78] A. J. Freeman, J. Yu. Phys. B+C 270, 50 (1988), https://doi.org/10.1016/0378-4363(88)90105-2.Search in Google Scholar
[79] P. Karen, H. Fjellvåg, A. Kjekshus. J. Solid State Chem. 97, 257 (1992), https://doi.org/10.1016/0022-4596(92)90033-R.Search in Google Scholar
[80] P. Karen, A. Kjekshus, A. F. Andresen. Acta Chem. Scand. 46, 1059 (1992), https://doi.org/10.3891/acta.chem.scand.46-1059.Search in Google Scholar
[81] D. Li, K. Lee, K. B. Y. Wang, M. Osada, S. Crossley, H. R. Lee, Y. Cui, Y. Hikita, H. Y. Hwang. Nature 572, 624 (2019), https://doi.org/10.1038/s41586-019-1496-5.Search in Google Scholar PubMed
[82] Y. Tokura, H. Takagi, S. Uchida. Nature 337, 345 (1989), https://doi.org/10.1038/337345a0.Search in Google Scholar
[83] A. W. Sleight, J. L. Gillson, P. E. Bierstedt. Solid State Commun. 17, 27 (1975), https://doi.org/10.1016/0038-1098(75)90327-0.Search in Google Scholar
[84] R. J. Cava, B. Batlogg, J. J. Krajewski, R. Farrow, L. W. RuppJr., A. E. White, K. Short, W. F. Peck, T. Kometani. Nature 332, 814 (1988), https://doi.org/10.1038/332814a0.Search in Google Scholar
[85] Z. Z. Sheng, A. M. Hermann. Nature 332, 55 (1988), https://doi.org/10.1038/332055a0.Search in Google Scholar
[86] S. S. P. Parkin, V. Y. Lee, E. M. Engler, A. I. Nazzal, T. C. Huang, G. Gorman, R. Savoy, R. Beyers. Phys. Rev. Lett. 60, 2539 (1988), https://doi.org/10.1103/PhysRevLett.60.2539.Search in Google Scholar PubMed
[87] T. Kaneko, H. Yamauchi, S. Tanaka. Physica C 178, 377 (1991), https://doi.org/10.1016/0921-4534(91)90086-E.Search in Google Scholar
[88] S. N. Putilin, E. V. Antipov, O. Chmaissem, M. Marezio. Nature 362, 226 (1993), https://doi.org/10.1038/362226a0.Search in Google Scholar
[89] Z. J. Huang, R. L. Meng, X. D. Qiu, Y. Y. Sun, J. Kulik, Y. Y. Xue, C. W. Chu. Physica C 217, 1 (1993), https://doi.org/10.1016/0921-4534(93)90786-P.Search in Google Scholar
[90] J. Nagamatsu, N. Nakagawa, T. Muranaka, Y. Zenitani, J. Akimitsu. Nature 410, 63 (2001), https://doi.org/10.1038/35065039.Search in Google Scholar PubMed
[91] J. M. An, W. E. Pickett. Phys. Rev. Lett. 86, 4366 (2001), https://doi.org/10.1103/PhysRevLett.86.4366.Search in Google Scholar PubMed
[92] A. F. Hebard, M. J. Rosseinsky, R. C. Haddon, D. W. Murphy, S. H. Glarum, T. T. M. Palstra, A. P. Ramirez, A. R. Kortan. Nature 350, 600 (1991), https://doi.org/10.1038/350600a0.Search in Google Scholar
[93] M. J. Rosseinsky, A. P. Ramirez, S. H. Glarum, D. W. Murphy, R. C. Haddon, A. F. Hebard, T. T. M. Palstra, A. R. Kortan, S. M. Zahurak, A. V. Makhija. Phys. Rev. Lett. 66, 2830 (1991), https://doi.org/10.1103/PhysRevLett.66.2830.Search in Google Scholar PubMed
[94] K. Tanigaki, T. W. Ebbesen, S. Saito, J. Mizuki, J. S. Tsai, Y. Kubo, S. Kuroshima. Nature 352, 222 (1991), https://doi.org/10.1038/352222a0.Search in Google Scholar
[95] P. W. Stephens, L. Mihaly, P.-L. Lee, R. L. Whetten, S.-M. Huang, R. Kaner, F. Deiderich, K. Holczer. Nature 351, 632 (1991), https://doi.org/10.1038/351632a0.Search in Google Scholar
[96] Y. Kamihara, H. Hiramatsu, M. Hirano, R. Kawamura, H. Yanagi, T. Kamiya, H. Hosono. J. Am. Chem. Soc. 128, 10012 (2006), https://doi.org/10.1021/ja063355c.Search in Google Scholar PubMed
[97] Y. Kamihara, T. Watanabe, M. Hirano, H. Hosono. J. Am. Chem. Soc. 130, 3296 (2008), https://doi.org/10.1021/ja800073m.Search in Google Scholar PubMed
[98] M. Rotter, M. Tegel, D. Johrendt. Phys. Rev. Lett. 101, 107006 (2008), https://doi.org/10.1103/PhysRevLett.101.107006.Search in Google Scholar PubMed
[99] J. H. Tapp, Z. Tang, B. Lv, K. Sasmal, B. Lorenz, P. C. W. Chu, A. M. Guloy. Phys. Rev. B 78, 060505(R) (2008), https://doi.org/10.1103/PhysRevB.78.060505.Search in Google Scholar
[100] R. H. Duncan Lyngdoh, H. F. Schaefer, R. B. King. Chem. Rev. 118, 11626 (2018), https://doi.org/10.1021/acs.chemrev.8b00297.Search in Google Scholar PubMed
[101] K. Kuroki, H. Usui, S. Onari, R. Arita, H. Aoki. Phys. Rev. B 79, 224511 (2009), https://doi.org/10.1103/PhysRevB.79.224511.Search in Google Scholar
[102] A. Wang, A. Milosavljevic, A. M. Milinda Abeykoon, V. Ivanovski, Q. Du, A. Baum, E. Stavitski, Y. Liu, N. Lazarevic, K. Attenkofer, R. Hackl, Z. Popovic, C. Petrovic. Inorg. Chem. 61, 11036 (2022), https://doi.org/10.1021/acs.inorgchem.2c00568.Search in Google Scholar PubMed
[103] J. W. Lynn, P. Dai. Physica C 469, 469 (2009), https://doi.org/10.1016/j.physc.2009.03.046.Search in Google Scholar
[104] S. Avci, O. Chmaissem, D. Y. Chung, S. Rosenkranz, E. A. Goremychkin, J. P. Castellan, I. S. Todorov, J. A. Schlueter, H. Claus, A. Daoud-Aladine, D. D. Khalyavin, M. G. Kanatzidis, R. Osborn. Phys. Rev. B 85, 184507 (2012), https://doi.org/10.1103/PhysRevB.85.184507.Search in Google Scholar
[105] B. T. Matthias. Phys. Rev. 97, 74 (1955), https://doi.org/10.1103/PhysRev.97.74.Search in Google Scholar
[106] L. Kautzsch, F. Mende, G. H. Fecher, J. Winterlik, C. Felser. Materials 12, 2580 (2019), https://doi.org/10.3390/ma12162580.Search in Google Scholar PubMed PubMed Central
[107] J. Kitagawa, S. Hamamoto, N. Ishizu. Metals 10, 1078 (2020), https://doi.org/10.3390/met10081078.Search in Google Scholar
[108] E. Krüger, H. P. Strunk. Symmetry 7, 561 (2015), https://doi.org/10.3390/sym7020561.Search in Google Scholar
[109] P. W. Atkins, T. L. Overton, J. P. Rourke, M. T. Weller, F. A. Armstrong. Inorganic Chemistry, W. H. Freeman and Company, New York, 5th ed. (2010).Search in Google Scholar
[110] P. Delahay, M. Pourbaix, P. Van Rysselberghe. J. Chem. Educ. 27, 683 (1950), https://doi.org/10.1021/ed027p683.Search in Google Scholar
[111] B. Hájek, P. Karen, V. Brožek. J. Less Common Met. 98, 245 (1984), https://doi.org/10.1016/0022-5088(84)90297-2.Search in Google Scholar
[112] V. Brožek, B. Hájek, P. Karen, M. Matucha, L. Žilka. J. Radioanal. Chem. 80, 165 (1983), https://doi.org/10.1007/BF02517659.Search in Google Scholar
[113] B. Hájek, P. Karen, V. Brožek. Rev. Inorg. Chem. 8, 117 (1986), https://doi.org/10.1515/REVIC.1986.8.1-2.117.Search in Google Scholar
© 2023 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- In this issue
- Editorial
- Special issue in honour of Dr. Mary Lowe Good
- Special topic papers
- Metal ion-assisted supramolecular gelation
- Synthesis and optical properties of phosphorus doped ZnO: X-ray absorption, X-ray emission, and X-ray excited optical luminescence studies
- A facile preparation of graphene hydrogel-supported bimetallic RuM (M: Co, Ni, Cu) nanoparticles as catalysts in the hydrogen generation from ammonia borane
- How to get deeper insights into the optical properties of lanthanide systems: a computational protocol from ligand to complexes
- Heuristic algorithms for understanding chemistry via simple quantities
- A NiII–WV(CN)8 layer magnet showing metamagnetic behavior
- Transition metal complexes for electrochromic and electrofluorochromic devices
- Dispersion control by using a bulky surfactant medium in the LB films for the enhancement of linearly polarized luminescence of Eu complexes
Articles in the same Issue
- Frontmatter
- In this issue
- Editorial
- Special issue in honour of Dr. Mary Lowe Good
- Special topic papers
- Metal ion-assisted supramolecular gelation
- Synthesis and optical properties of phosphorus doped ZnO: X-ray absorption, X-ray emission, and X-ray excited optical luminescence studies
- A facile preparation of graphene hydrogel-supported bimetallic RuM (M: Co, Ni, Cu) nanoparticles as catalysts in the hydrogen generation from ammonia borane
- How to get deeper insights into the optical properties of lanthanide systems: a computational protocol from ligand to complexes
- Heuristic algorithms for understanding chemistry via simple quantities
- A NiII–WV(CN)8 layer magnet showing metamagnetic behavior
- Transition metal complexes for electrochromic and electrofluorochromic devices
- Dispersion control by using a bulky surfactant medium in the LB films for the enhancement of linearly polarized luminescence of Eu complexes