Next Article in Journal
Heteroligand α-Diimine-Zn(II) Complexes with O,N,O′- and O,N,S-Donor Redox-Active Schiff Bases: Synthesis, Structure and Electrochemical Properties
Next Article in Special Issue
One-Dimensional Iodoantimonate(III) and Iodobismuthate(III) Supramolecular Hybrids with Diiodine: Structural Features, Stability and Optical Properties
Previous Article in Journal
Development of Ion Character Property Relationship (IC-PR) for Removal of 13-Metal Ions by Employing a Novel Green Adsorbent Aerva javanica
Previous Article in Special Issue
The Influence of the Halide in the Crystal Structures of 1-(2,3,5,6-Tetrafluoro-4-pyridyl)-3-benzylimidazolium Halides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Energy Electronegativity and Chemical Bonding

by
Stepan S. Batsanov
National Research Institute for Physical-Technical Measurements, 141570 Mendeleevo, Russia
Molecules 2022, 27(23), 8215; https://doi.org/10.3390/molecules27238215
Submission received: 25 October 2022 / Revised: 17 November 2022 / Accepted: 21 November 2022 / Published: 25 November 2022
(This article belongs to the Special Issue Covalent and Noncovalent Interactions in Crystal Chemistry)

Abstract

:
Historical development of the concept of electronegativity (EN) and its significance and prospects for physical and structural chemistry are discussed. The current cutting-edge results are reviewed: new methods of determining the ENs of atoms in solid metals and of bond polarities and effective atomic charges in molecules and crystals. The ENs of nanosized elements are calculated for the first time, enabling us to understand their unusual reactivity, particularly the fixation of N2 by nanodiamond. Bond polarities in fluorides are also determined for the first time, taking into account the peculiarities of the fluorine atom’s electronic structure and its electron affinity.

1. Introduction

Of some 108 substances known today, only a few hundred (i.e., elementary solids and molecules) are homo-nuclear, all the rest containing polar bonds. To describe the polarity, Avogadro (1809) and Berzelius (1811) listed elements according their ‘negative’ or ‘positive’ electrochemical characters. The development of this approach led to a conclusion that the bond polarity is caused by a shift of valence electrons towards the atom that attracts them more strongly, i.e., that with the higher EN [1,2]. Thus, by comparing the ENs of the bonded atoms, we find which of them has a negative charge and how large it is. The full history of the EN concept is described in detail by Jensen [3,4] and Sproul [5]. Remarkably, the popularity of this concept has not decreased in 200 years–on the contrary, in recent years it has begun to be used to solve new problems of materials science, structural chemistry and high-pressure physics and chemistry.
Fajans [6,7] created the theory of polarization to study chemical bonding from the ionic viewpoint, but this approach cannot quantify the chemical bond since the radii (and hence the energy of polarization) of ‘pure’ ions cannot be experimentally found [8], whereas the properties of covalent bonds can be measured accurately, yielding the basis for a quantitative scale of ENs.
The modern age of the EN concept began in 1932 when Pauling [9] found that
χM − χX = a [EMX − ½ (EMM + EXX)]½ = a ΔEMX½
where χ is the EN, E is the bond energy, and a = 0.208 or 0.102 if the energy is expressed in kcal/mol or kJ/mol, respectively (Pauling then expressed E in eV, for which a = 1—this is the now well-forgotten origin of his unit of EN!).Thus, Pauling showed that the thermal effect of a reaction is a function of charges on the atoms. This work initiated numerous studies deriving EN from, or relating it to, a wide variety of physical properties of materials. Here, we will limit ourselves only to the energy-based system of ENs and its relation to the effective charges of atoms in molecules and crystals. Section 2.1, Section 2.2 and Section 2.3 provide a concise review of the development of the EN concept from 1932 to the present day, as applied to free atoms, atoms in molecules and in solids. Section 2.4 contains the author’s new research on applying the EN concept to explain peculiar properties of solid nanoparticles, rationalizing their high reactivity and deriving (for the first time) a specific EN system for nanomaterials. Section 3.1, Section 3.2 and Section 3.3 review the applications of EN to the problem of effective atomic charges and bond polarities in molecules and solids (Section 3.3). Section 3.2 also contains new calculations for compounds and bonds containing fluorine, a stumbling-block for the EN approach due to its peculiar electronic configuration. Section 3.4 and Section 3.5 review the application of EN to coordination compounds of transition metals (including some counter-intuitive, but experimentally validated, predictions) and to the high-pressure behavior of substances. The concluding Section 4 briefly surveys various critiques—scientific, philosophical and ideological—of the EN concept.

2. Electronegativity

2.1. Electronegativity of Free Atoms

Mulliken was the first to express the ENs of atoms through their ionization energies [10,11]:
χ = ½ (Iv + Av)
where Iv is the first ionization potential and Av is the electron affinity for the valence states of the atoms. Between Mulliken’s and Pauling’s values of the EN, there is a straightforward relation: χP = 0.36 χM. Unfortunately, there is serious ambiguity in specifying the valence state of some atoms; e.g., for N and O one has to choose from 7 and 8 possible states, respectively, with ionization energies and electron affinities covering ranges of 10 or more eV [12]. Pearson [13] used the ground-state ionization energy and electron affinity of the atom (Io and Ao) to calculate the EN by Equation (2). Although his ENs generally show the expected trends in the Periodic Table, there are some strikingly unrealistic values. These violations disappear if the average (rather than the first) ionization potentials for the ground state are used. Obviously, if all outer valence electrons of an atom are involved in the chemical bonding, they become equivalent and their energy must be characterized by the average potential:
I ¯ = 1 v i = 1 v I i
Converting to the dimensionality of Pauling’s EN (square root of energy), we obtain [14]:
χ = c ( I ¯ ) 1 / 2
which gives good agreement with the standard system of ENs. The I ¯ were used for calculating ENs also in [15,16,17]. Table 1 lists the ENs from [14], improved in some cases by the use of up-to-date values of ionization potentials for polyvalent elements [18]. Mulliken’s approach was developed further by many authors (see reviews in [19,20]).

2.2. Electronegativity of Atoms in Molecules

Equation (1), or its modifications, were applied by many authors to calculate the ENs of elements for various valence states, using ever more extensive and precise experimental data (see a survey in [16]). However, it is really valid for low-polarity molecules only, for the bond energy depends on many factors besides the polarity, particularly on the bond distance, atomic valences and hybridization states. An elegant solution to these difficulties was found by Tantardini and Oganov [21] who proposed the equation:
E MX   = E MX cov   ( 1 + Δ χ MX 2 ) )
where E MX cov is the covalent component of the bond energy, and the EN difference is the measure of the relative (rather than absolute, as in Pauling’s approach) ionic contribution. Equation (5) allows us to obtain correct trends of EN in the Periodic Table and to eliminate some anomalies in both the absolute and the relative values of the ENs yielded by Equation (1), such as those of the alkali metals (Li 1.85, Na 1.96, K 1.99, Rb 2.00, Cs 1.93). Tantardini–Oganov’s ENs give a reasonable prediction of bond polarity and improve the description of the thermal effects of chemical reactions. However, these ENs differ so much from the generally accepted values that it is difficult to use them in previously found functional dependencies.
Therefore, Batsanov [22] improved Equation (1) purely empirically while preserving its philosophy. For MXn type compounds, this was achieved by replacing the constant a = 0.102 with a variable parameter c = 0.1 + 0.015(n* + dv) where n* is the ratio of the principal quantum numbers of the metal and nonmetal, d is the bond length (in Å), and v is the valence of M. This correction takes into account that bond polarity is enhanced by an increase in bond length (reducing the overlap of the valence orbitals) and of n* but lowered when the valence increases. Table 2 lists the average ENs calculated by this procedure using, for non-metals, χ = 3.9 (F), 3.1 (Cl), 2.9 (Br) and 2.6 (I).

2.3. Electronegativity of Atoms in Solids

Most EN systems are based on the properties of molecules which retain their structures in all aggregate states, but in inorganic solids with covalent, metallic or ionic bonds the situation drastically differs owing to the large cooperative effect. Despite this obvious difference, conventional ‘molecular’ ENs are still routinely applied to ionic crystals, metallic alloys or coordination compounds. Thus, the authors of [23,24,25,26] derived the ENs of elements from the physical properties of solids, while using the ‘molecular’ ENs of C, N, O as reference points, although Duffy [27,28] had shown that the ENs of halogen, oxygen and chalcogen atoms also differ somewhat in the solid state.
The idea of specific solid-state ENs for crystals was developed by Batsanov [29]. The ‘chemical scale’ of Pettifor [30,31] showed clear separation of different structures of binary compounds (MX) in two-dimensional maps in the coordinates χM and χX, and although he used Pauling’s ENs for Be through F, curiously enough, for most elements his values were close to the solid-state ENs of Batsanov [29]. The latter also adapted Equation (2) for the solid state by replacing the ionization potentials of free atoms with the work functions (Φ) as their analogs for the solid state [32] and the electron affinities of isolated atoms with those of atoms in solids (As), measured by secondary ion mass spectrometry [33]:
χs = ½ (Φ + As)
Usually, Mulliken’s ENs are converted to Pauling’s scale (χo) with the help of a constant factor [10], but more correctly, this conversion should be element-specific rather than uniform, by using Equation (7):
χ s χ o = Φ + A s I o + A o
where χo is the thermochemical EN in Pauling’s scale, referring to the lowest ionization state. The results of calculations of χ s by Equations (6) and (7) are listed in Table 3.
The thermochemical ENs of atoms in crystalline compounds MXn (Table 4) can be calculated by Equation (1), replacing ΔEMX with the enthalpy of the formation of compounds in the solid state (Qs):
Δχs = a (Qs/n) ½
where a = 0.1 + cn* and c = 0.01 for the Mn+ cations with a noble-gas electron configuration, or 0.02 otherwise [32].

2.4. Electronegativity of Nanosized Elements

Experimental data and all theoretical models show that the cohesive energy of nanoparticles decreases linearly with their size (see the review in [34]). From the structural point of view, a decrease in the particle size reduces the average coordination number in it, as the fraction of surface atoms (with lower coordination numbers) relatively increases. Ultimately, in a 13-atom metal cluster, all atoms but one lie on the surface, and in an 8-atom fragment of a diamond-like structure—all atoms. The cohesive (atomization) energy (Ep) of a particle of arbitrary size is related to that of the bulk solid (Eb) as
E p = E b   N ¯ p N b
where N ¯ p is the average coordination number of atoms in the particle, and Nb is that in the bulk [35]. Methods of calculating the coordination numbers of atoms on the surface of cubic crystals are described in [36,37], and the results of these calculations are given in Table 5 for 5 nm-sized particles.
Of course, N ¯ p is only the average: the surface atoms are different from those in the interior of a particle, and while the ‘physical’ averaging does occur to some degree, its extent and character depends on the nature of the chemical bond in the particle. It is naïve indeed to imagine the surface bristling with broken, dangling bonds—such a particle will be too reactive to exist. The surface can be terminated by alien atoms or functional groups—e.g., nanodiamond (ND) particles are terminated by C-H bonds or by -OH, -COOH, etc., groups [38]—and thus become chemically different from the bulk. If such termination is absent or stripped away, the outer layer of ND reconstructs into a shell of sp2 carbon atoms, variously described as graphitic, fullerene- or onion-like [39], with an underlying thicker (up to 5 Å) layer of amorphous sp3 carbon [40]. In metals, the relaxation may take the form of stronger (and shorter) bonding between the first and the second outermost layers of atoms, compensated by longer distances between the second and the third, with such alternating distortions perpetrated for several layers inward [41].
A decrease in the cohesive energies in Cu, Au, Pt, C, Si, or Ge particles of such sizes by an average of ca. 60 kJ/mol will enhance their reactivity. Indeed, we found that complete evaporation of aqueous transparent Si or Ge colloids yields solid residues containing SiO2 or GeO2 microcrystals of the quartz type with refractive indices ne = 1.553, no = 1.544 and ne =1.735, no = 1.695, respectively. In the case of water colloids of carbon, we observed a mass loss of carbon equal to 0.02 wt % due to the formation of CO2, which either escapes or is consumed in the growth of fibers often formed from such colloids [42]. Thus, the nanosized grains of E = C, Si and Ge replace hydrogen in water under ambient conditions.
E + 2H2O = EO2 + 2H2
In fact, similar conclusions for Cu and Pt have been occasionally made earlier. Thus, it was shown [43] by DFT calculations that the cohesive energy of Cu atoms decreases linearly with their coordination numbers in particles and, therefore, copper nanoparticles have a lower equilibrium potential for dissolution in water. A full thermodynamic analysis [44,45] of size effect on the standard potential defining Pt/PtO equilibrium, has shown that under certain conditions PtO may form in water solutions.
The size effect in the cohesive energy of carbon was studied by measuring the oxidation heat of polycrystalline synthetic diamond (PD) with D = 146 μm, and carbon nanotubes (CNT) with D = 50–70 μm.
In addition, detonation nanodiamond (DND) with D = 5 nm [37] was also studied. The results, shown in Figure 1, indicate that the smaller the particle size is, the higher the combustion heat (viz. ΔHox = 29.7, 34.4 and 38.6 kJ/g for PD, CNT and DND, respectively), hence the cohesive energy is lower. Although CNT is sp2 carbon, while PD is sp3 carbon and DND mostly so, this difference is far too small to account for the observed effect: note that the enthalpy of the diamond to graphite phase transition is ca. 0.2 J/g.
From Table 5, it is evident that the difference between Eb and Ep for diamond (58 kJ/mol on average) is big enough even to change the sign of the thermal effect of a chemical reaction, as we observed in the case of fixation of the atmospheric nitrogen by the 5-nm sized detonation nanodiamond [42,46]. Over time, diluted aqueous colloids (suspensions) of DND gave rise to cottonwool-like fibers with the nitrogen content and nitrogen to carbon ratio higher than in the DND used, the only possible source of the excess nitrogen being the atmospheric, or specially injected, N2. One can envisage a three-stage process:
2C + N2 = C2N2
C2N2 + 4H2O = 2CO2 + 2NH3 + H2
C2N2 + 3O2 = 2CO2 + 2NO
The second and especially the third reaction are indeed consistent with the gaseous product detected by mass-spectroscopy [46]; they are also both exothermic, with ΔH = −47.2 and −297.5 kJ mol, respectively. The first stage, however, is highly unfavorable thermodynamically (ΔH = 308.9 kJ mol) under standard conditions and would require temperatures above 1400 °C to proceed. However, this refers to a bulk diamond with the atomization energy Ea = 716.7 kJ mol; the reaction would be exothermic for atomic carbon (ΔH = −1124.5 kJ mol) or C2 molecules (ΔH = −506.5 kJ mol). A 22% decrease in the Ea(C) would make this reaction exothermic. Admittedly, C2N2 was not detected, which can be due to its rapid hydrolysis or oxidation.
Obviously, as the particle size decreases further, N ¯ p will approach the coordination number of the surface atoms, Ns, and in this case, the surface energy, Es, will be
E s = E b   N s N b
Using the Es/Io ratio for each element as a conversion factor, one can calculate Is for the extremely small nanoparticles, and then with the help of Equation (4) find the ENs of nanophases of elements (Table 6).

3. Effective Charges of Atoms

3.1. Bond Polarity in Molecules

Effective charges on atoms can be neither directly measured nor theoretically defined in a unique manner, for the obvious reason that the electron shells of atoms in molecules or solids overlap, but nevertheless are very useful in chemistry. They can be deduced from observable parameters using a variety of models, from purely empirical and mechanistic to grounded in quantum chemistry, such as QTAIM [47]. Meister and Schwarz [48] surveyed 26 such models and found them to have the same underlying physics. It is the electron density (ED) distribution that determines all chemical and physical phenomena and may be measured, for instance, by high-resolution XRD experiments [49]. Thus, the task is reduced to finding the ED in a certain region of an atom and comparing it with that in a free atom.
Currently, the most popular is the evaluation of the effective charges of atoms in molecules from dipole moments, μ
μ = qd
where q is the atomic charge, and the dipole length d can be identified with the bond length. Then, the μ/d ratio should characterize the polarity of the chemical bond, p. This ratio was used by Fajans to estimate the bond polarities in HCl and HBr molecules [50], and later Pauling [51] used it to determine the ionicity (polarity) in any chemical bond as
i = 1 e X
where X = cΔχ2 (initially c = 0.25 was suggested, and later c = 0.18 [52]). Table 7 lists the dipole moments of diatomic molecules (from [18], unless specified) and their polarity of bonds calculated as p = μ/4.8 d, with μ in Debye units (D) and d in Å.
However, Equation (11) is a rather crude simplification. Coulson [55] concluded that the μ of a bond consists of several terms, caused by the EN difference (μi), the electron density difference (μρ), the hybridization of the atomic orbitals (μh), and the effect of non-bonding electrons (μe). Thus, the bond ionicity can be found only from μi, e.g., when other components of μ are known theoretically or empirically. Moreover, the assumption that the d in Equation (11) is the distance between the nuclei of atoms is also only the first approximation. Owing to the flexibility of the electron clouds around each nucleus, the effective charge center is not at the nucleus, and the accurate dipole moments should be equal to μ = qRc, where Rc is the separation between these charges [56]. Authors of [57] indicated that atomic charges in a highly ionic molecule can be computed only if the molecular dipole moment, the polarizabilities of free cations and neutral atoms of the anions, are all known. In [58], a new approach to calculating the local dipole moments and charges of atoms based on Bader’s AIM theory has been presented. Further development of this approach is based on the study of the infrared intensities of fundamental bands in gaseous molecules, which contain a wealth of information about electronic structure and its changes on molecular vibration. The authors of [59,60,61,62,63,64] showed that from these experimental data, one can obtain values of the bond ionicity in molecules (Table 8), which are qualitatively consistent with the results of calculations by Equation (12).

3.2. Equalization of Electronegativity of Atoms in Molecules

In this context, the idea expressed by Pauling [65] that in stable compounds ‘each atom has only a small electric charge, approximating zero’ is very important. Since the positive charge increases the EN of a metal atom (M), but the negative charge of nonmetal (X) reduces it, the ENs of atoms in an MX molecule must converge. Sanderson [66,67] developed this idea and postulated the full equalization of the atomic ENs in the molecule. Using, as a reference, the bond ionicity in the NaF molecule (0.75), Sanderson was able to calculate the effective charges of atoms in other molecules.
In the work of [68], it has been found that the energy of ionization M → MN+, expressed as a fraction of removing all valence electrons, shows a simple parabolic dependence on the degree of ionization; the ‘fractional charge’ i = n /N:
I ¯ = n I / N I = i 2
This dependence is a continuous and smooth function, despite the fact that the ionization of each isolated atom is a discrete process. As has been pointed out [69], the energy functions E(N) for atoms-in-molecules are differentiable with respect to N, although N has only integer values for isolated atoms.
Accordingly, the electron affinity
A ¯ = 1 i 2
The intersection point of the I (i) and A (i) curves means that at M + X, the contact of the ionization energy EI = i2I and the electron affinity EA = (1 − i2)A are equal (Figure 2), hence:
i = A A + I 1 / 2
The values of i for halides MXn can be calculated under the assumption that all outer electrons in a metal atom are equivalent, involved in bonding, and, upon hybridization, become averaged in their properties. As a result, the average ionization potential I ¯ n of all valence electrons turns out to be much more informative than I1 [14,15]. Use of the average ionization potential I ¯ n in Equation (15) allows us to calculate i in the MXn molecules with a different n, i.e., finding i for different valences of the metal atoms.
An important reservation concerns applying Equation (15) to fluorides. According to the correlation found between the electron affinity and atomic sizes or ionization potentials of heavier halogens, fluorine must have a higher A value than the observed one. The difference can be explained by the destabilizing effect (in F) of a high concentration of negative charge in a small volume [70,71,72]. In fact, other Period 2 elements show a similar trend: the electron affinities of O, N, C, B are lower than those of S, P, Si, Al, respectively. Bonding of a fluorine atom to any other increases the volume available for its electrons and, hence, relieves the inter-electron repulsion. Therefore, a fluorine atom in compounds is better characterized by the ‘extrapolated’ value of A = 4.49 eV [70]. Table 9 presents the values of the bond polarities (iB) in MX molecules, calculated by Equation (15) using this A (F) and the standard values of electron affinities of other halogens and ionization potentials of metals from [18].
In Table 9, Batsanov’s values of iB are compared with those calculated by Equations (12) and (16), i.e., according to Pauling, iP, or Sanderson (iS):
i S = S MX S M Δ S
where S is the EN in Sanderson’s scale, SMX = S M S X and ΔS = SM+SM [66]. As is seen from Table 9, there is only a qualitative agreement between these methods, but both the Pauling’s and the Sandersen’s equations depend on the accepted reference points, while Equation (15) uses only the experimentally established dependence of the ionization potentials on the charges of atoms, and the values of these characteristics themselves are measured with great accuracy. Generally speaking, there are 26 different methods (!) for estimating the effective charges of atoms in compounds calculated by various models of quantum chemistry [73]; however, in this review, we limit these to empirical approaches only using the EN concept.

3.3. Effective Charges of Atoms in Solid Compounds

The first experimental estimates of the effective charges of atoms in crystalline compounds were based on the work of Born [74], who suggested that in inorganic crystals, the electric field acting on each ion is equal to the applied external field related to the dielectric permittivity, ε, and the refractive index, n, with the frequency of the transverse lattice vibrations (ωt):
ε = n 2 + ( 4 π e 2 / ω t 2 m ¯ V )
where m ¯ is the reduced mass of the vibrating particle and ωt is a frequency of the transverse oscillations.
A comparison of the experimentally measured and calculated values of ε showed that, usually, the computed ε are less than the experimental ones. Szigeti [75,76,77,78], taking into account that the field strength in the dielectric is less than the external field due to the polarization of the specimen, deduced the equation:
ε = n 2 + 4 π e 2 ω o 2 m ¯ V ( n 2 + 2 ) 2 9
The values of ε calculated by Equation (18) are always higher than those measured experimentally. This led him to the idea to replace the formal charge of the ion by the real effective charge e*. Thus, Szigeti’s formula allows us to calculate the e* in crystals, using the experimental values of n and ε, which are higher than those in molecules owing to the excess of coordination numbers over the valence of atoms in crystals.
For the calculation of i in the MX crystals in Equation (15) it is necessary to replace Io with the work function, Φ, which is the ionization potential for a solid, and then
i s = A A + Φ 1 / 2
Table 10 shows that our values of the i of atoms calculated by Equation (19) are close to the experimental data of Szigeti’s e* [79].
The greater effective charges in halides CuX and AgX are caused by forming the additional bonds Cu, Ag →X of (n − 1)d-electrons in these metals, and Tl→X of np-electrons with vacant nd-orbitals in halogens [80,81].

3.4. Coordination Charges of Atoms in Chemical Compounds

Above, the ‘effective charge ’notion is interpreted as a synonym of ‘bond ionicity’ (in fractions of an electron). In reality, the term ‘charge’ means either an excess or a deficit of electrons in the given region of the atom, as compared to its free state. Therefore, it is important to determine the region of the atomic space to which the charge being determined relates. Batsanov [82] had proposed to divide the effective charges determined by different methods into two groups—proper (qo) and coordination (qc) charges. If the first term characterizes the situation inside the closed orbitals of an atom, then the second type is a result of the electron interaction of the given atom with its neighbors in the first coordination sphere.
This qo defines the Coulomb energy, is responsible for the IR absorption, causes the atomic polarization bands, and affects the binding energy of the internal electrons in the atom. However, what matters for redox reactions is the electron density in the interatomic space, i.e., the qc. Therefore, Suchet [83], instead of using the terms ‘proper’ and ‘coordination’ as they refer to the charges of atoms, used the terms ‘physical’ and ‘chemical’.
The notion ‘charge on the atom’ and different methods of its determination have been discussed by Catlow and Stoneham [84]. From the above it follows that
qo = vi, qc = v ± (1 − i) Nc
where v is a valence and Nc is the coordination number of the atom in the crystal structure. According to Equation (20), in crystalline compounds with low bond polarity, qc can become negative if Nc > Z (see details in [18]). Interestingly, in 1948, Pauling [65] calculated the qc according to Equation (20) without specifying this notion. Table 11 compares that of the experimental qc (determined by X-ray spectroscopy [85,86,87,88]) with the qc calculated by the EN method in several crystalline compounds.
As follows from Table 11, when q ¯ c = 1 ± 0.3, which is well in agreement with Pauling’s principle of electro-neutrality [52]: in a stable compound, the qc must be close to ±1e. As an illustration of this principle, we experimentally list (by the X-ray method) the measured distribution of charges in the Co and Cr complexes [89]:
[Co(NH3)6]2+  qN = −0.62
[Cr(CN)6]2− qN = −0.54
qH = + 0.36 qC = + 0.22
qCo = −0.49 qCo = −0.38
Coordination charges of high-EN elements (Cu, Au, Hg, Tl, Pt) in the low valence state in some crystalline compounds, according to the EN method, must be < 0 [90,91,92,93,94,95,96]; later, ESCA studies confirmed this conclusion with the example of platinum and gold compounds [97,98,99,100,101,102,103].
Consider, as an example, two models of distribution charges in compounds of univalent gold and divalent platinum: either Au + X and Pt + q X 2   q (classical chemistry), or Au X + and Pt q X 2 + q (concept of EN). In fact, the observed oxidation reactions occur according to the schemes:
AuI + Cl2 → AuICl2 and PtI2 + Cl2 → PtI2Cl2
which confirm the EN predictions. All possible mixed tetra-halides of Pt and tri-halides of Au, di-halides of Cu [95] and Hg [92], and chalcogenohalides of Tl [93] were synthesized by the oxidation of metals owing to their negative charges.
Interestingly, the mixed halides of PtIV have different properties depending on the order; in the case of halogens, these were added as follows: PtX2 + Y2 or PtY2 + X2. The structure of PtX2 has a motif of squares with shared vertices, and additional halogens complete these to form octahedra in PtX2Y2, while in PtY2X2, the atoms of Y are in squares, but additional atoms of X are in the vertical coordinate. Such compounds were named the square-coordinate isomers [91]. TlSeBr also showed different properties depending on the route of synthesis: TlBr + Se → Se = TlIII–Br or 2Tl + Se2Br2 → 2TlI–Se–Br. The given structural formulae are confirmed by IR-spectroscopy and DSC study; these compounds were named the valence isomers [93].
Studying literatures showed that similar reactions were observed by other researchers in Fe, Sb, Cr, Re, W, and U compounds with a variety of ligands, such as halogens, chalcogens, SCN, N3, NO3, CO3, SO4 and methyl. Thus, an oxidation reaction of the polyvalent metals in solid compounds is the chemical method that allows for the estimation of the sign of their coordination charges.

3.5. Change of Chemical Bonding under Pressure

The effect of high pressure on the bond ionicity in crystalline compounds is a very interesting but insufficiently studied phenomenon. For a long time, the only source of experimental information about this effect was the study of pressure dependences upon the optical and polarization properties of crystals, and their explanation by Szigeti’s equation. These studies show that the e* of atoms in crystals under high pressures usually decreases, but in AgI, TlI, HgTe, AlSb, GaN, InAs, PbF2 and SiO2 it increases with pressure (see references in [18]). Difficulties in measuring the optical properties of crystals under high pressures, increasing anharmonicy of vibrations, and the deformation of IR absorption bands, reduces the precision of determining the effective charges of atoms in compressed crystal. Therefore, it is desirable to find an independent method to solve this problem. Such a method can be the EN technique, see [104,105].
Let reaction M + X → MX have the thermal effect Q at ambient thermodynamic conditions, and the compression works of the initial reagents and the final product under pressure denoted as Wc. Obviously, if Wc (mixture) > Wc (compound) > 0, then ΔWc should be subtracted from the standard heat effect to yield the Q corresponding to high pressures, and vice versa. Usually ΔWc > 0, hence under pressure, the Q and Δχ (bond ionicity) decrease. The results of such an approach qualitatively agrees with the experiment, except for alkali hydrides, where the calculations predict the e* to fall to 0, already at several tens of GPa, while in fact, no change of electronic structure occurred up to 100 GPa and even higher [106,107]. This contradiction can be resolved by taking into account that the compression work only partly goes into changing the chemical bonds. Thus, a comparison of the Wc of mixtures and compounds allows us to define a change in Q (hence of the ENs of atoms and bond polarity) on variation of the pressure. The reduction of Q and the bond polarity under pressure is observed in crystals of AB-type, viz. Group 1—Group 17, Group 2—Group 16, Group 13—Group 15 compounds. Table 12 shows that the ∂e*/∂P for these crystals decreases by 10−4 to 10−3 GPa−1. Szigeti’s method predicts the same signs and similar absolute values, ∂e*/∂P = 3.3 × 10−4 GPa−1. Remarkably, the increase in Q(P) in CuX, AgX, and TlX under high pressures indicates an increase in polarity; the corresponding data using Szigeti’s method are not available.
Such behavior of substances under pressure is caused by the different compressibility of cations and anions. If the anion is softer than the cation (e.g., in halides of Cu, Ag and Tl), then it will be compressed more strongly and the ∂e*/∂P > 0, i.e., the bond polarity, will increase with pressure. The electronegativity of the anion, being inversely related to the atomic size, will increase and, hence, so will Δχ, if χX > χM. In the case of a softer cation (e.g., in alkali halides), the χM on compression will increase more strongly than χX, and the ionicity will decrease.

4. Concluding Remarks

The concept of EN, like other methods of theoretical chemistry, has its critics, which can be divided into two groups—some researchers criticize the particular approach to the calculation of EN and propose new, more successful ones (from their point of view); others see fundamental defects in the concept of EN and suggest it be completely excluded from chemistry.
The first criticism came from Fajans [108], who pointed out that in the series HC ≡ CCl → H2C=CHCl → H3C–CH2Cl, the charge of Cl changes the sign from +1 to 0 to −1, which contradicts the notion of a constant EN of the carbon atom. In fact, χ(C) depends on the state of the hybridization of carbon, being 2.5 for sp3, 2.9 for sp2 and 3.2 for sp, and since χ(Cl) = 3.0, this explains the reversion of its charge; this was noted in the author’s letter to Fajans. Hückel [109] criticized the EN dimension (energy½), as physically meaningless; however, to calculate bond ionicity, Pauling used the values Δχ2; we noted [110] a definite analogy between Δχ and the wave function ψ, bearing in mind that only |ψ|2 has physical significance. Iczkowski and Margrave [111] draw attention to the different dimensions of ENs of different methods. Syrkin [112] stated that the EN concept cannot explain modern data, ignoring the agreement of the XRS atomic charges in crystals with the results of the EN method [113]. Spiridonov and Tatevsky [114] noted that the EN concept, using the atom in the molecule approach, contradicts (?) the philosophy of quantum chemistry. Comments on this irrational criticism are given in works [82,115]. A summary of the acute discussion of the concept of EN, which took place in the Soviet Union in 1962, can be found in [116].
The arguments in favor of EN can be summarized thus. The fact that EN is defined through different observed properties, and so has a non-unique dimensionality, merely reflects the multi-faceted nature of the chemical bond. Indeed, this can be an asset rather than a liability, as EN can serve as a nodal point connecting the various physical characteristics of a substance, hence its wide usage in chemistry. A certain ‘fuzziness’ of the EN concept is really typical for chemistry cf. the notions of metallicity, acidity, reactivity, etc. Sixty years have passed since the stormy discussions and condemnations of the EN concept, but its development and application in various fields of science, such as structural chemistry, materials science, molecular spectroscopy, and various fields of physical and inorganic chemistry [117,118,119,120,121] does not stop; it was even suggested to use EN as the third coordinate in the Periodic Table [122].
Pauling’s concept of EN formed the quantitative dependence of the thermodynamics of substances on the chemical bonding within them. This foundation provided the successful development and application of the EN concept in chemistry.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The author thanks A.S. Batsanov (Durham University, UK) for helpful discussions.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Lewis, G.N. The atom and the molecule. J. Am. Chem. Soc. 1916, 38, 762–785. [Google Scholar] [CrossRef] [Green Version]
  2. Ingold, C.K. The principles of aromatic substitution, from the standpoint of the electronic theory of valency. Rec. Trav. Chim. 1929, 48, 797–812. [Google Scholar] [CrossRef]
  3. Jensen, W.B. Electronegativity from Avogadro to Pauling. I. Origins of the electronegativity concept. J. Chem. Educ. 1996, 73, 11–20. [Google Scholar] [CrossRef]
  4. Jensen, W.B. Electronegativity from Avogadro to Pauling: II. Late nineteenth- and early twentieth-century developments. J. Chem. Educ. 2003, 80, 279–287. [Google Scholar] [CrossRef]
  5. Sproul, G.D. Evaluation of electronegativity scales. ACS Omega 2020, 5, 11585–11594. [Google Scholar] [CrossRef]
  6. Fajans, K.; Berlin, T. Quantization of molecules, inter- and intra-molecular forces. Phys. Rev. 1943, 63, 309–312. [Google Scholar] [CrossRef]
  7. Fajans, K. Quantikel-Theorie der chemischen Bindung. Chimia 1959, 13, 349–366. [Google Scholar]
  8. Batsanov, S.S. On the relationship between the theory of polarization and the concept of electronegativity. Russ. J. Inorg. Chem. 1957, 2, 1482–1487. (In Russian) [Google Scholar]
  9. Pauling, L. The nature of the chemical bond IV. The energy of single bonds and the relative electronegativity of atoms. J. Am. Chem. Soc. 1932, 54, 3570–3582. [Google Scholar]
  10. Mulliken, R.S. A new electroaffinity scale; together with data on valence states and on valence ionization potentials and electron affinities. J. Chem. Phys. 1934, 2, 782–793. [Google Scholar] [CrossRef]
  11. Mulliken, R.S. Electronic structures of molecules. Electroaffinity, molecular orbitals and dipole moments. J. Chem. Phys. 1935, 3, 573–585. [Google Scholar] [CrossRef]
  12. Hinze, J.; Jaffé, H.H. Electronegativity. Orbital electronegativity of neutral atoms. J. Am. Chem. Soc. 1962, 84, 540–546. [Google Scholar] [CrossRef]
  13. Pearson, R.G. Absolute electronegativity and hardness: Application to inorganic chemistry. Inorg. Chem. 1988, 27, 734–740. [Google Scholar] [CrossRef]
  14. Martynov, A.I.; Batsanov, S.S. A new approach to the determination of the electronegativity of atoms. Russ. J. Inorg. Chem. 1980, 25, 1737–1739. [Google Scholar]
  15. Allen, A.C. Electronegativity is the average one-electron energy of the valence-shell electrons in ground-state free atoms. J. Am. Chem.Soc. 1989, 111, 9003–9014. [Google Scholar] [CrossRef]
  16. Politzer, P.; Murray, J.S. Electronegativity—A perspective. J. Mol. Model. 2018, 24, 214–221. [Google Scholar] [CrossRef]
  17. Rahm, M.; Zeng, T.; Hoffmann, R. Electronegativity seen as the ground-state average valence electron bonding energy. J. Am. Chem. Soc. 2019, 141, 342–351. [Google Scholar] [CrossRef]
  18. Batsanov, S.S.; Batsanov, A.S. Introduction to Structural Chemistry; Springer: Dordrecht, The Netherlands, 2012. [Google Scholar]
  19. Campero, A.; Díaz Ponce, J.A. Averaged scale in electronegativity joined to physicochemical perturbations. Consequences of periodicity. ACS Omega 2020, 5, 25520–25542. [Google Scholar] [CrossRef]
  20. Sproul, G. Cardinal electronegativity values correlate with physicochemical properties. Inorg. Chem. 2021, 60, 10970–10978. [Google Scholar] [CrossRef]
  21. Tantardini, C.; Oganov, A.R. Thermochemical electronegativities of the elements. Nat. Commun. 2021, 12, 2087. [Google Scholar] [CrossRef]
  22. Batsanov, S.S. Thermochemical electronegativities of metals. Russ. J. Phys. Chem. 2000, 74, 267–270. [Google Scholar]
  23. Stevenson, D.P. Heat of chemisorption of hydrogen in metals. J. Chem. Phys. 1955, 23, 203. [Google Scholar] [CrossRef]
  24. Phillips, J.C. Ionicity of the chemical bond in crystals. Rev. Mod. Phys. 1970, 42, 317–356. [Google Scholar] [CrossRef]
  25. Trasatti, S. Electronegativity, work function, and heat of adsorption of hydrogen on metals. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1972, 68, 229–236. [Google Scholar] [CrossRef]
  26. Frese, K.W., Jr. Simple method for estimating levels of solids. J. Vac. Sci. Technol. 1979, 16, 1042–1044. [Google Scholar] [CrossRef]
  27. Duffy, J.A. Variable electronegativity of oxygen in binary oxides: Possible relevance to molten fluorides. J. Chem. Phys. 1977, 67, 2930–2931. [Google Scholar] [CrossRef]
  28. Duffy, J.A. Chemical bonding in the oxides of the elements: A new appraisal. J. Solid State Chem. 1986, 62, 145–157. [Google Scholar] [CrossRef]
  29. Batsanov, S.S. System of electronegativities and effective atomic charge in crystalline compounds. Russ. J. Inorg. Chem. 1975, 20, 1437–1440. [Google Scholar]
  30. Pettifor, D.G. A chemical scale for crystal-structure maps. Solid State Commun. 1984, 5, 31–34. [Google Scholar] [CrossRef]
  31. Pettifor, D.G. Structure maps revisited. J. Phys. Condens. Matter 2003, 15, V13–V16. [Google Scholar] [CrossRef]
  32. Batsanov, S.S.; Batsanov, A.S. Solid-state electronegativity of atoms: New approaches. Acta Cryst. B 2021, 77, 495–505. [Google Scholar] [CrossRef]
  33. Wilson, R.G.; Stevie, F.A.; Magee, C.W. Secondary Ion Mass Spectrometry: A Practical Handbook for Depth Profiling and Bulk Impurity Analysis; Wiley-Interscience: New York, NY, USA, 1989. [Google Scholar]
  34. Yang, C.C.; Mai, Y.-W. Thermodynamics at the nanoscale: A new approach to the investigation of unique physicochemical properties of nanomaterials. Mater. Sci. Eng. R Rep. 2014, 79, 1–40. [Google Scholar]
  35. Shandiz, M.A.; Safaei, A.; Sanjabi, S.; Barber, Z.H. Modeling the cohesive energy and melting point of nanoparticles by their average coordination number. Solid State Commun. 2008, 145, 432–437. [Google Scholar] [CrossRef]
  36. Pirkkalainen, K.; Serimaa, R. Coordination number in ideal spherical nanocrystals. J. Appl. Cryst. 2009, 42, 442–447. [Google Scholar] [CrossRef]
  37. Batsanov, S.S.; Dan’kin, D.A. Size effect in cohesive energy of elements. Mater. Chem. Phys. 2017, 196, 245–248. [Google Scholar] [CrossRef]
  38. Kulakova, I.I. Surface chemistry of nanodiamonds. Phys. Solid State 2004, 46, 636–643. [Google Scholar] [CrossRef]
  39. Chang, S.L.Y.; Dwyer, C.; Osawa, E.; Barnard, A.S. Size dependent surface reconstruction in detonation nanodiamonds. Nanoscale Horiz. 2018, 3, 213–217. [Google Scholar] [CrossRef]
  40. Fang, X.W.; Mao, J.D.; Levin, E.M.; Schmidt-Rohr, K. Nonaromatic core-sell structure of nanodiamond from solid-state NMR spectroscopy. J. Am. Chem. Soc. 2009, 131, 1426–1435. [Google Scholar] [CrossRef] [PubMed]
  41. Kadas, K.; Nabi, Z.; Kwon, S.K.; Vitos, L.; Ahuja, R.; Johansson, B. Surface relaxation and surface stress of 4d metals. Surf. Sci. 2006, 600, 395–402. [Google Scholar] [CrossRef]
  42. Batsanov, S.S.; Guriev, E.D.L.; Gavrilkin, S.M.; Hamilton, K.A.; Lindsey, K.; Mendis, B.G.; Riggs, H.J.; Batsanov, A.S. On the nature of fibres grown from nanodiamond colloids. Mater. Chem. Phys. 2016, 173, 325–332. [Google Scholar]
  43. Taylor, C.D.; Neurock, M.; Scully, J.R. First-principles investigation of the fundamental corrosion properties of a model Cu38 nanoparticle and the (111), (113) surfaces. J. Electrochem. Soc. 2008, 155, C407–C414. [Google Scholar] [CrossRef]
  44. Tang, L.; Han, B.; Persson, K.; Friesen, C.; He, T.; Sieradzki, K.; Ceder, G. Electrochemical stability of nanometer-scale Pt particles in acidic environments. J. Am. Chem. Soc. 2010, 132, 596–600. [Google Scholar] [CrossRef]
  45. Tang, I.; Li, X.; Cammarata, R.C.; Friesen, C.; Sieradzki, K. Electrochemical stability of elemental metal nanoparticles. J. Am. Chem. Soc. 2010, 132, 11722–11726. [Google Scholar] [CrossRef] [PubMed]
  46. Batsanov, S.S.; Gavrilkin, S.M.; Shatalova, T.B.; Mendis, B.G.; Batsanov, A.S. Fixation of atmospheric nitrogen by nanodiamonds. New J. Chem. 2018, 42, 11160–11164. [Google Scholar] [CrossRef]
  47. Bader, R.W.F. Atoms in Molecules: A Quantum Theory; Oxford University Press: Oxford, UK, 1990. [Google Scholar]
  48. Meister, J.; Schwarz, W.H.E. Principal components of ionicity. J. Phys. Chem. 1994, 98, 8245–8252. [Google Scholar] [CrossRef]
  49. Coppens, P. X-ray Charge Densities and Chemical Bonding; Oxford University: Oxford, UK, 1997. [Google Scholar]
  50. Fajans, K. Deformation von Ionen und Molekeln auf Grund refraktometrischer Daten. Z. Phys. 1928, 50, 531–536. [Google Scholar] [CrossRef]
  51. Pauling, L. The Nature of the Chemical Bond, 3rd ed.; Cornell University Press: Ithaca, NY, USA, 1960. [Google Scholar]
  52. Pauling, L. Interatomic distances and bond character in the oxygen acids and related substances. J. Phys. Chem. 1952, 56, 361–365. [Google Scholar] [CrossRef]
  53. Thomas, J.M.; Walker, N.R.; Cooke, S.A.; Gerry, M.C.L. Microwave spectra and structures of KrAuF, KrAgF, and KrAgBr; 83Kr nuclear quadrupole coupling and the nature of noble gas-noble metal halide bonding. J. Am. Chem. Soc. 2004, 126, 1235–1246. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, R.; Steimle, T.C.; Cheng, L.; Stanton, J.F. The permanent electric dipole moment of gold chloride, AuCl. Mol. Phys. 2015, 113, 2073–2080. [Google Scholar] [CrossRef]
  55. Coulson, C.A. Valence; Clarendon Press: Oxford, UK, 1952. [Google Scholar]
  56. Hou, S.; Bernath, P.F. Relationship between dipole moments and harmonic vibrational frequencies in diatomic molecules. J. Phys. Chem. A 2015, 119, 1435–1438. [Google Scholar] [CrossRef] [PubMed]
  57. Hou, S.; Qureshi, A.H.; Wei, Z. Atomic charges in highly ionic diatomic molecules. ACS Omega 2018, 3, 17180–17187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Groß, L.; Herrmann, C. Local electric dipole moments: A generalized approach. J. Comput. Chem. 2016, 37, 2260–2265. [Google Scholar] [CrossRef] [PubMed]
  59. Gussoni, M.; Castiglioni, C.; Zerbi, G. Physical meaning of electrooptical parameters derived from infrared intensities. J. Phys. Chem. 1984, 88, 600–604. [Google Scholar] [CrossRef]
  60. Gussoni, M.; Castiglioni, C.; Ramos, M.N.; Zerbi, G. Ab initio counterpart of infrared atomic charges. Chem. Phys. Lett. 1987, 142, 515–518. [Google Scholar] [CrossRef]
  61. Haiduke, R.L.A.; Bruns, R.E. An atomic charge-charge flux-dipole flux atom-in-molecule decomposition for molecular dipole-moment derivatives and infrared fundamental intensities. J. Phys. Chem. A 2005, 109, 2680–2688. [Google Scholar] [CrossRef] [PubMed]
  62. Silva, A.F.; Richter, W.E.; Bassi, A.B.M.S.; Bruns, R.E. Dynamic atomic contributions to infrared intensities of fundamental bands. Phys. Chem. Chem. Phys. 2015, 17, 30378–30388. [Google Scholar] [CrossRef] [Green Version]
  63. Richter, W.E.; Duarte, L.J.; Silva, A.F.; Bruns, R.E. Review of experimental GAPT and infrared atomic charges in molecules. J. Braz. Chem. Soc. 2016, 27, 979–991. [Google Scholar] [CrossRef]
  64. Richter, W.E.; Duarte, L.J.; Bruns, R.E. Are “GAPT charges” really just charges? J. Chem. Inf. Model. 2021, 61, 3881–3890. [Google Scholar] [CrossRef]
  65. Pauling, L. The modern theory of valency. J. Chem. Soc. 1948, 1461–1467. [Google Scholar] [CrossRef]
  66. Sanderson, R.T. An interpretation of bond lengths and a classification of bonds. Science 1951, 114, 670–672. [Google Scholar] [CrossRef]
  67. Sanderson, R.T. Electronegativity and bond energy. J. Am. Chem. Soc. 1983, 105, 2259–2261. [Google Scholar] [CrossRef]
  68. Batsanov, S.S. Simple semi-empirical method for evaluating bond polarity in molecular and crystalline halides. J. Mol. Struct. 2010, 980, 225–229. [Google Scholar] [CrossRef]
  69. Fuentealba, P.; Parr, R.G. Higher-order derivatives in density-functional theory, especially the hardness derivative ∂η/∂N. J. Chem. Phys. 1991, 94, 5559–5564. [Google Scholar] [CrossRef]
  70. Politzer, P. Anomalous properties of fluorine. J. Am. Chem. Soc. 1969, 91, 6235–6237. [Google Scholar] [CrossRef]
  71. Batsanov, S.S. Structure and properties of fluorine, oxygen, and nitrogen atoms in covalent bonds. Russ. Chem. Bull. 1989, 38, 410–412. [Google Scholar] [CrossRef]
  72. Murray, J.S.; Seybold, P.G.; Politzer, P. The many faces of fluorine: Some noncovalent interactions of fluorine compounds. J. Chem. Thermodyn. 2021, 156, 106382. [Google Scholar] [CrossRef]
  73. Manz, T.A. Seven confluence principles: A case study of standardized statistical analysis for 26 methods that assign net atomic charges in molecules. RSC Adv. 2020, 10, 44121–44148. [Google Scholar] [CrossRef]
  74. Born, M. Über elektrostatische Gitterpotentiale. Z. Physik 1921, 7, 124–140. [Google Scholar] [CrossRef]
  75. Szigeti, B. Polarisability and dielectric constant of ionic crystals. Trans. Faraday Soc. 1949, 45, 155–166. [Google Scholar] [CrossRef]
  76. Szigeti, B. Compressibility and absorption frequency of ionic crystals. Proc. R. Soc. Lond. Ser. A 1950, 204, 51–62. [Google Scholar]
  77. Szigeti, B. Higher -order terms in the dielectric constant of ionic crystals. Proc. R. Soc. Lond. Ser. A 1959, 252, 217–235. [Google Scholar]
  78. Szigeti, B. The magnitude of the anharmonic contribution to the static dielectric constant. Proc. R. Soc. Lond. Ser. A 1961, 261, 274–280. [Google Scholar]
  79. Batsanov, S.S. Dielectric methods of studying the chemical bond and the concept of electronegativity. Russ. Chem. Rev. 1982, 51, 684–697. [Google Scholar] [CrossRef]
  80. Berkowitz, J.; Batson, C.; Goodsman, G. Photoelectron spectroscopy of AgCl, AgBr, and AgI vapors. J. Chem. Phys. 1980, 72, 5829–5837. [Google Scholar] [CrossRef]
  81. Tossel, J.A.; Vaughan, D.J. Relationships between valence orbital binding energies and crystal structures in compounds of copper, silver, gold, zinc, cadmium, and mercury. Inorg. Chem. 1981, 20, 3333–3340. [Google Scholar] [CrossRef]
  82. Batsanov, S.S. The concept of electronegativity. Conclusions and prospects. Russ. Chem. Rev. 1968, 37, 332–351. [Google Scholar] [CrossRef]
  83. Suchet, J.P. Chemical Physics of Semiconductors; Van Nostrand: London, UK, 1965. [Google Scholar]
  84. Catlow, C.R.A.; Stoneham, A.M. Ionicity in solids. J. Phys. C Solid State Phys. 1983, 16, 4321–4338. [Google Scholar] [CrossRef]
  85. Barinskii, R.L. Determination of effective charges of atoms in complexes from the X-ray absorption spectra. J. Struct. Chem. 1960, 1, 183–190. [Google Scholar] [CrossRef]
  86. Vainshtein, E.E.; Kopelev, Y.F. X-ray spectral analysis of aromatic complexes of the transition elements. J. Struct. Chem. 1962, 3, 433–441. [Google Scholar] [CrossRef]
  87. Barinskii, R.L. On the paper by E.E. Vainstein and Yu. F. Kopelev “X-ray spectral analysis of aromatic complexes of the transition elements”. J. Struct. Chem. 1962, 3, 442–445. [Google Scholar] [CrossRef]
  88. Batsanov, S.S.; Ovsyannikova, I.A. Atomic charges in nickelecene and nickelecinium determined by means of X-ray spectra. Dokl. Akad. Nauk SSSR 1965, 165, 855–856. (In Russian) [Google Scholar]
  89. Iwata, M. X-ray determination of the electron distribution in crystals of [Co(NH3)6] [Cr(CN)6] at 80 K. Acta Crystallogr. Sect. B 1977, 33, 59–69. [Google Scholar] [CrossRef]
  90. Batsanov, S.S.; Ruchkin, E.D. Mixed halides of tetravalent platinum. Russ. J. Inorg. Chem. 1959, 4, 779–783. [Google Scholar]
  91. Batsanov, S.S.; Ruchkin, E.D. About isomers in mixed platinum halides. Russ. J. Inorg. Chem. 1965, 10, 1415–1418. [Google Scholar]
  92. Batsanov, S.S.; Podberezskaya, N.V.; Khripin, L.A. Mercury salts with mixed anions. Synthesis and properties of mixed halides. Russ. Chem. Bull. 1965, 14, 199–203. [Google Scholar] [CrossRef]
  93. Batsanov, S.S.; Rigin, V.I. Chalcogenohalides of thalium. Dokl. Akad. Nauk SSSR 1964, 158, 1355–1357. (In Russian) [Google Scholar]
  94. Batsanov, S.S.; Rigin, V.I. Isomerism of thallium selenobromides. Dokl. Akad. Nauk SSSR 1966, 167, 89–90. (In Russiasn) [Google Scholar]
  95. Batsanov, S.S.; Zalivina, E.N.; Derbeneva, S.S.; Borodaevsky, V.E. Synthesis and properties of copper bromo- and iodo-chlorides. Dokl. Akad. Nauk SSSR 1968, 181, 599–602. (In Russian) [Google Scholar]
  96. Batsanov, S.S.; Sokolova, M.N.; Ruchkin, E.D. Mixed halides of gold. Russ. Chem. Bull. 1971, 20, 1757–1759. [Google Scholar] [CrossRef]
  97. Pantelouris, A.; Kueper, G.; Hormes, J.; Feldmann, C.; Jansen, M. Anionic gold in Cs3AuO and Rb3AuO established by X-ray absorption spectroscopy. J. Am. Chem. Soc. 1995, 117, 11749–11753. [Google Scholar] [CrossRef]
  98. Feldmann, C.; Jansen, M. Zur Kristallchemischen Ahnlichkeit von aurid- und halogenid-ionen. Z. Anorg. Allgem. Chem. 1995, 621, 1907–1912. [Google Scholar] [CrossRef]
  99. Mudring, A.-V.; Jansen, M. Base-induced disproportionation of elemental gold. Angew. Chem. Int. Ed. 2000, 39, 3066–3067. [Google Scholar] [CrossRef]
  100. Karpov, A.; Nuss, J.; Wedig, U.; Jansen, M. Cs2Pt: A platinide(-II) exhibiting complete charge separation. Angew. Chem. Int. Ed. 2003, 42, 4818–4821. [Google Scholar] [CrossRef]
  101. Saltykov, V.; Nuss, J.; Konuma, M.; Jansen, M. Investigation of the quasi-binary system BaAu–BaPt. Z. Allgem. Anorg. Chem. 2009, 635, 70–75. [Google Scholar] [CrossRef]
  102. Smetana, V.; Mudring, A.-V. Cesium platinide hydride 4Cs2Pt·CsH: An intermetallic double salt featuring metal anions. Angew. Chem. Int. Ed. 2016, 55, 14838–14841. [Google Scholar] [CrossRef] [PubMed]
  103. Agnarelli, L.; Prots, Y.; Burkhardt, U.; Schmidt, M.; Koželj, P.; Leithe-Jasper, A.; Grin, Y. Mg3Pt2: Anionic chains in a Eu3Ga2-type structure. Inorg. Chem. 2021, 60, 13681–13690. [Google Scholar] [CrossRef]
  104. Batsanov, S.S. Pressure effect on the heat of formation of condensed substances. Russ. J. Phys. Chem. 1999, 73, 1–6. [Google Scholar]
  105. Batsanov, S.S. Chemical bonding evolution on compression of crystals. J. Struct. Chem. 2005, 46, 306–314. [Google Scholar] [CrossRef]
  106. Ghandehari, K.; Luo, H.; Ruoff, A.L.; Trail, S.S.; Disalvo, F.J. Crystal structure and band gap of rubidium hydride to 120 GPa. Mod. Phys. Lett. B 1995, 9, 1133–1140. [Google Scholar] [CrossRef]
  107. Ghandehari, K.; Luo, H.; Ruoff, A.L.; Trail, S.S.; Disalvo, F.J. New high pressure crystal structure and equation of state of cesium hydride to 253 GPa. Phys. Rev. Lett. 1995, 74, 2264–2267. [Google Scholar] [CrossRef]
  108. Fajans, K. General chemistry. By Linus Pauling. J. Phys. Chem. 1951, 55, 1107–1108. [Google Scholar] [CrossRef]
  109. Hückel, W. Die chemische Bindung. Kritische Betrachtung der Systematik, der Ausdrucksweisen und der formelmäßigen Darstellung. J. Prakt. Chem. 1957, 5, 105–174. [Google Scholar] [CrossRef]
  110. Batsanov, S.S. Comments on Hückel’s book. Russ. J. Phys. Chem. 1960, 34, 444–445. [Google Scholar]
  111. Iczkowski, R.P.; Margrave, J.L. Electronegativity. J. Am. Chem. Soc. 1961, 83, 3547–3551. [Google Scholar] [CrossRef]
  112. Syrkin, Y.K. Effective charges and electronegativity. Russ. Chem. Rev. 1962, 31, 197–207. [Google Scholar] [CrossRef]
  113. Batsanov, S.S. Comment on the article “Effective charges and electronegativites” of Syrkin. Russ. J. Phys. Chem. 1963, 37, 761-174. [Google Scholar]
  114. Spiridonov, V.P.; Tatevskii, V.M. On concept of electronegativity of atoms. Russ. J. Phys. Chem. 1963, 37, 522–526, 661–664, 848–849, 1070–1072, 1177–1179. [Google Scholar]
  115. Batsanov, S.S. Comment on articles of Spiridonov and Tatevskii criticizing concept of electronegativity. Russ. J. Phys. Chem. 1967, 41, 1298–1301. [Google Scholar]
  116. Zvolinskii, V.P. A conference on the concept of electronegativity. J. Struct. Chem. 1962, 3, 478–483. [Google Scholar] [CrossRef]
  117. Accorinti, H.L. Incompatible models in chemistry: The case of electronegativity. Found. Chem. 2019, 21, 71–81. [Google Scholar] [CrossRef]
  118. Tandon, H.; Chakraborty, T.; Suhag, V. A scale of atomic electronegativity in terms of atomic nucleophilicity index. Found. Chem. 2020, 22, 335–346. [Google Scholar] [CrossRef]
  119. von Szentpály, L. Theorems and rules connecting bond energy and bond order with electronegativity equalization and hardness maximization. Theor. Chem. Acc. 2020, 139, 54. [Google Scholar] [CrossRef]
  120. Nordholm, S. From electronegativity towards reactivity—Searching for a measure of atomic reactivity. Molecules 2021, 26, 3680. [Google Scholar] [CrossRef] [PubMed]
  121. Racioppi, S.; Rahm, M. In-situ electronegativity and the bridging of chemical bonding concepts. Chem. Eur. J. 2021, 27, 18156–18167. [Google Scholar] [CrossRef] [PubMed]
  122. Cherkasov, A.R.; Galkin, V.I.; Zueva, E.M.; Cherkasov, R.A. The concept of electronegativity: The current state of the problem. Russ. Chem. Rev. 1998, 67, 375–392. [Google Scholar] [CrossRef]
Figure 1. DSC/TGA curves for the oxidation of carbon with different grain sizes.
Figure 1. DSC/TGA curves for the oxidation of carbon with different grain sizes.
Molecules 27 08215 g001
Figure 2. Schematic presentation of I and A as functions of i; the curves correspond to Equation (15).
Figure 2. Schematic presentation of I and A as functions of i; the curves correspond to Equation (15).
Molecules 27 08215 g002
Table 1. Ionization electronegativity of free atoms.
Table 1. Ionization electronegativity of free atoms.
LiBeBCNOF
0.901.451.902.372.853.323.78
NaMgAlSiPSCl
0.891.311.641.982.322.652.98
KCaScTiVCrMnFeCoNi
0.811.171.501.25 b1.60 c1.33 b1.32 b1.35 b1.38 b1.40 b
1.861.92 d1.66 c1.70 c1.67 c1.72 c1.76 c
2.221.98 d2.03 d
CuZnGaGeAsSeBr
1.481.641.832.091.70 c2.612.88
1.66 b 2.27
RbSrYZrNbMoTcRuRhPd
0.801.131.411.23 b1.53 c1.92 d1.88 d1.35 b1.39 b1. 45 b
1.712.022.33 1.93 d1.94 d2.01 d
AgCdInSnSbTeI
1.571.651.801.29 b1.60 c2.452.68
2.002.24
CsBaLaHfTaWReOsIrPt
0.771.081.351.28 b1.54 c1.83 d1.85 d1.39 b1.40 b1.45 b
1.711.992.282.521.86 d1.89 d1.95 d
AuHgTlPbBiPoThU
1.781.790.96 a1.31 b1.58 c2.291.59 d1.68 d
1.93 c 1.892.072.13
valence a v = 1, b v = 2, c v = 3, d v = 4.
Table 2. Thermochemical electronegativity of metals in molecular halides.
Table 2. Thermochemical electronegativity of metals in molecular halides.
M (I)χM (II)χM (II)χM (III)χM (IV)χ
Li0.98Cu2.04Hg1.95Sc1.34Ti1.77
Na0.93Be1.43Sn1.42Y1.25Zr1.63
K0.72Mg1.38Pb1.55La1.12Hf1.56
Rb0.67Ca1.00Cr1.53B1.96C2.60
Cs0.53Sr0.95Mn1.35Al1.64Si2.05
Cu1.46Ba0.82Fe1.58In1.73Ge2.15
Ag1.61Zn1.60Co1.64Ga1.73Sn2.06
Au1.84Cd1.70Ni1.71V1.80Pb2.17
Tl1.06 Cr1.92W2.19
Table 3. Ionization electronegativity ( χ s ) of atoms in solid metals.
Table 3. Ionization electronegativity ( χ s ) of atoms in solid metals.
LiBeBC
0.470.651.171.29
NaMgAlSi
0.410.621.031.18
KCaScTiVCrMnFeCoNi
0.380. 520.700.981.031.030.890.980.971.16
CuZnGaGeAsSe
0.930.641.141.281.181.46
RbSrYZrNbMoTcRuRhPd
0.350.500.650.971.051.171.011.321.341.30
AgCdInSnSbTeI
1.020.761.111.010.981.341.78
CsBaLaHfTaWReOsIrPt
0.280.460.660.800.831.141.221.201.281.32
AuHgTlPbBiThU
1.230.820.781.071.140.890.97
Table 4. Thermochemical electronegativity ( χ s ) of metals in solid compounds.
Table 4. Thermochemical electronegativity ( χ s ) of metals in solid compounds.
M (I) χ s M (II) χ s M (II) χ s M (III) χ s M (IV) χ s
Li0.55Cu1.33Sn1.09Sc0.69Ti0.68
Na0.48Be0.98Pb1.04Y0.72Zr0.77
K0.37Mg0.68Cr0.96La0.58Hf0.75
Rb0.35Ca0.44Mn0.73Al1.00C2.60
Cs0.32Sr0.41Fe1.18In1.73Si2.05
Cu1.25Ba0.42Co1.07Ga1.19Ge2.15
Ag1.30Zn0.85Ni1.24V0.88Sn0.81
Au2.01Cd0.91Pd1.55Nb0.66Pb1.53
Tl0.85Hg1.44Pt1.64Sb1.27Th0.82
Bi1.22U0.95
Cr1.01W1.14
Table 5. Bulk and average coordination numbers of particles, cohesive energies (kJ/mol) in bulk and nanosized cubic elements.
Table 5. Bulk and average coordination numbers of particles, cohesive energies (kJ/mol) in bulk and nanosized cubic elements.
MNb N ¯ EbEpMNb N ¯ EbEp
Ac129.828406332Mn1210.42283246
Ag1210.33285245Mo86.863659565
Al1210.34331285Na86.45010786.7
Au1210.33368317Nb86.849733627
Ba86.186179138Ni1210.56430378
C43.629717650Pb89.976195162
Ca129.717178144Pd1210.41377327
Co1210.55427375Pt1210.40566490
Cr86.958397346Ra86.140159122
Cs85.78176.555.3Rb85.93980.960.0
Cu1210.69337301Rh1210.44556434
Fe86.964415362Si43.435450386
Ge43.412372317Sr126.205164127
Ir1210.43609529Ta86.807782671
K86.07789.067.6Th129.921602498
La129.839431353V86.907515445
Li86.732159134W86.856851729
Table 6. Electronegativities of nanosized elements.
Table 6. Electronegativities of nanosized elements.
MNsEsEb/IoIsχnanoMNsEsEb/IoIsχnano
Ac7.1812430.36526.8961.02MnII7.1811690.25456.9021.02
Ag7.1811700.38974.5340.83MoIV4.5233720.301212.821.40
Al7.1811980.193110.631.27Na4.52360. 80.21682.9050.66
Au7.1812200.41365.5200.92NbIII4.5234140.49418.6921.15
Ba4.5231010.24404.2980.81NiIII7.1812570.2192112.21.36
C23580.200718.501.68PbII4.5231100.18026.3490.98
Ca7.1811060.20495.3820.90PdII7.1812250.28118.3101.12
CoIII7.1812550.226911.661.33PtII7.1813380.42618.2331.12
CrIII4.5232250.228010.211.25Ra4.52389.90.21364.3610.81
Cs4.52343.20.20362.2010.58Rb4.52345.70.20072.3620.60
Cu7.1812020.45264.6230.84RhIII7.1813330.305411.291.31
FeIII4.5232350.235910.321.25Si22250.180912.891.40
Ge21860.158612.151.36Sr7.18198.10.20325.0030.87
IrIII7.1813640.351010.761.28TaIII4.5234420.51928.8251.16
K4.52350.30.21252.4540.61ThIV7.1813600.38189.7771.22
La7.1812580.37297.1681.04VIII4.5232910.31629.5511.20
Li4.52390.10.30623.0480.68WIV4.5234810.486410.251.25
Table 7. Dipole moments (in D) and bond polarities in MX molecules.
Table 7. Dipole moments (in D) and bond polarities in MX molecules.
MFClBrI
μpμpμpμp
H1.8260.411.1080.180.8270.120.4480.06
Li6.3270.847.1290.737.2260.697.4280.65
Na8.1560.889.0020.799.1180.769.2360.71
K8.5920.8210.270.8010.630.7810.820.74
Rb8.5460.7810.510.7810.860.7711.480.75
Cs7.8840.7010.390.7310.820.7311.690.73
Cu5.770.695.2 a0.53
Ag6.220.656.080.555.620.494.550.37
Au4.320.473.69 b0.35
Tl4.2280.424.5430.384.490.364.610.34
a [53], b [54].
Table 8. IRS-bond ionicity (e*/v) in molecules.
Table 8. IRS-bond ionicity (e*/v) in molecules.
MoleculeiMoleculeiMoleculeiMoleculeiMoleculei
HF0.382OH20.236CH40.028BF30.506LiF0.861
HCl0.184NH30.034SiH40.226BCl30.249LiCl0.760
HBr0.114NF30.385GeH40.216CF40.512NaF0.889
HI0.040PH30.119SnH40.254CCl40.261NaCl0.809
LiH0.654PF30.580 CO20.268KCl0.830
Table 9. Bond ionicity in MX halides.
Table 9. Bond ionicity in MX halides.
MXiBiPiSMXiBiPiS
LiF0.670.780.75CuF0.610.660.37
LiCl0.630.550.67CuCl0.560.380.28
LiBr0.620.480.62CuBr0.550.310.24
LiI0.600.380.54CuI0.530.210.15
NaF0.680.800.80AgF0.610.610.41
NaCl0.640.570.71AgCl0.570.330.33
NaBr0.630.500.67AgBr0.550.260.28
NaI0.610.400.58AgI0.540.160.20
KF0.710.840.85AuF0.570.53
KCl0.670.640.76AuCl0.530.25
KBr0.660.570.72AuBr0.520.18
KI0.640.470.64AuI0.500.13
CsF0.730.870.97TlF0.650.760.64
CsCl0.690.690.89TlCl0.610.530.55
CsBr0.680.640.84TlBr0.590.460.51
CsI0.660.540.76TlI0.560.350.43
Table 10. Effective charges of atoms in MX crystals.
Table 10. Effective charges of atoms in MX crystals.
MFClBrI
ise*ise*ise*ise*
Li0.810.810.780.770.760.740.730.54
Na0.810.830.780.780.760.750.740.74
K0.820.920.790.810.770.770.750.75
Rb0.820.970.800.840.770.800.750.77
Cs0.840.960.820.850.800.820.780.78
Cu0.71 0.680.980.650.960.620.91
Ag0.710.890.680.710.650.670.570.61
Tl0.74 0.710.880.680.840.650.83
Table 11. Coordination effective charges of metals in compounds.
Table 11. Coordination effective charges of metals in compounds.
MetalCompoundsqc (exp)qc (cal)MetalCompoundsqc (exp)qc (cal)
CrCrSO4·7H2O1.91.8CoCo(NO3)31.20.6
Cr(NO3)31.21.3 Co(C5H5)20.40.7
K2CrO40.10.5 Co(C5H5)2Cl1.00.9
Cr(C6H6)21.31.4NiNi(C5H5)20.70.6
MnMn(NO3)2·4H2O1.81.8 Ni(C5H5)2Cl1.00.8
K3Mn(CN)60.90.6OsOsO20.80.7
Mn(C5H5)21.51.3 K2OsCl60.80.5
Fe(NH4)2Fe(SO4)2·6H2O1.91.7 K2OsO40.80.7
K3Fe(CN)61.00.4 K2OsNCl50.70.8
Fe(C5H5)20.60.7 KOsO3N1.00.9
Table 12. Change of atomic charges (−∂e*/∂P, 10−4 GPa−1) in crystals MX under P = 10 GPa.
Table 12. Change of atomic charges (−∂e*/∂P, 10−4 GPa−1) in crystals MX under P = 10 GPa.
MIFClBrIMIIOSSeTe
Li5.21.501.9Be0−1.61.63.7
Na7.02.41.54.0Mg0.9005.6
K11.75.95.44.1Ca3.72.23.37.5
Rb10.63.92.41.7Sr7.05.97.29.6
Cs10.46.78.37.6Ba16.516.518.024.6
Cu −8.4−4.76.9Zn0−2.204.4
Ag−3.7−5.6−3.4−6.6Cd0−2.805.9
Tl−12.6−12.5−11.6−10.9Hg−5.9−2.80.610.0
MIIINPAsSbSn−12.2−2.800
B0.69.74.7 Pb−7.8−3.704.7
Al0.65.30.60Mn−0.4−4.0 1.6
Ga2.29.42.52.8
In6.910.01.91.9
La 15.314.714.4
Th2.57.5
U03.4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Batsanov, S.S. Energy Electronegativity and Chemical Bonding. Molecules 2022, 27, 8215. https://doi.org/10.3390/molecules27238215

AMA Style

Batsanov SS. Energy Electronegativity and Chemical Bonding. Molecules. 2022; 27(23):8215. https://doi.org/10.3390/molecules27238215

Chicago/Turabian Style

Batsanov, Stepan S. 2022. "Energy Electronegativity and Chemical Bonding" Molecules 27, no. 23: 8215. https://doi.org/10.3390/molecules27238215

Article Metrics

Back to TopTop