Abstract
The electron configurations of Ca, Zn and the nine transition elements M in between (and their heavier homologs) are reviewed on the basis of density functional theory and experimental facts. The d-s orbital energy and population patterns are systematically diverse. (i) The dominant valence electron configuration of most free neutral atoms M0 of groups g = 2–12 is 3d g−2 4s 2 (textbook rule), or 3d g−14s 1. (ii) Formal M q+ cations in chemical compounds have the dominant configuration 3d g−q 4s 0 (basic concept of transition metal chemistry). (iii) M0 atoms in metallic phases [M∞] of hcp, ccp(fcc) and bcc structures have intermediate populations near 3d g−1 4s 1 (lower d populations for Ca (ca. ½) and Zn (ca. 10)). Including the 4p valence orbitals, the dominant metallic configuration is 3d g−δ 4(sp) δ with δ ≈ 1.4 (±0.2) throughout (except for Zn). (iv) The 3d,4s population of atomic clusters M m varies for increasing m smoothly from single-atomic 3d g−24s 2 toward metallic 3d g−14s 1. – The textbook rule for the one-electron energies, i.e., ns < (n−1)d, holds ‘in a broader sense’ for the s block, but in general not for the d block, and never for the p block. It is more important to teach realistic atomic orbital (AO) populations such as the ones given above.
1 Atomic configurations and orbital energies, a historical review of concepts
1.1 Free atoms
A century ago, Bohr initiated the description of atomic states in terms of atomic electron configurations [1, 2]. After the invention of quantum mechanics in 1925, the consistent theory of atomic structure emerged on this sound basis, and was reviewed in 1935 [3]. Model theories of periodic solids and of individual molecules were set off, too [4], [5], [6], [7]. A configuration is the specification of spin orbital functions occupied by single electrons, the anti-symmetrized product of which often is a reasonable approximation for the true electronic matter-field, represented by a many-electron wavefunction. The single-electron orbitals are usually assumed to be of the ‘canonical’ (orthogonal and delocal) type, approximating many single-electronic ionization and excitation processes. Thereby, these orbitals can be connected to observable data [8]. This is an important aspect, since the orbital model is the best approach toward a deeper understanding of chemical bonding and reactions [9], [10], [11], [12], [13], [14], [15], [16], [17].
After World War II, Moore published her reviews of the theoretical analyses of a century of atomic vacuum Vis–UV spectroscopy [18, 19]. At the beginning of each section, she displayed the dominant electron configuration, from which the observed atomic orbital (AO) and spin-coupled S L J π (p) ground level is approximately derived (quantum numbers for total spin S with multiplicity 2S+1, total spatial angular momentum L, total angular momentum J, parity π, and parentage p). In the following years, in particular after the first centenary of the Periodic Table in 1969, textbooks of general chemistry (e.g., refs. [20], [21], [22], [23]) began selling the dominant electron configurations of chemically unbound, angular-momentum-coupled ground states of the neutral atoms in physical vacuum as paradigmatic for “the chemical elements”, i.e., the bound, partially charged, spatially perturbed atoms confined in chemical compounds. Combined with an oversimplified Aufbau rule (i.e., neglecting the two-electron repulsion energies), the energetic order of AOs is often assumed as independent of the element number Z, the atomic charge q and the bonding state, following always the so-called (n+ℓ,n) rule, see e.g., refs. [24], [25], [26], [27], [28]. A section of the (n+ℓ,n) AO energy pattern (namely the one relevant for the 3d elements in period n = 4) is displayed as rule (1).
the (n + ℓ, n) textbook rule 1, fitting: the group 2 atoms and their chemistry
But in reality, the (n+ℓ,n) energy rule, i.e., ns < (n−1)d < np, simulates the observable one-electron energy level pattern only for group 2 valence shells (i.e., neither for groups 0–1 nor for groups 3–18, nor for the inner atomic core (X-ray) levels, see e.g., refs. [18, 19, 29], [30], [31], [32]). For the alkali metal atoms of group 1, spectroscopic data shows that even ns < np < (n−1)d ≷ (n+1)s, while for groups 12–18 (n−1)d < ns (rule 3). Already in 1923, Bohr had documented the strong variation of d and f orbital levels versus the more regular pair of s and p levels [2].
Bohr’s century-old diagrams of single-electron ionization energies ε (from the smallest to the largest ones in terms of √|ε| values, in a single graphic) versus the nuclear charge or element numbers (up to 92) show a rather smooth behavior for the pairs of s and p (p½, p 3/2) levels, while the d and the f levels vary in a stepwise manner, crossing with the s,p pairs up and down. After WWII, the deductions from the spectroscopic observations were supported by many quantum-mechanical atomic structure calculations at the single-configuration density-functional or Hartree–Fock (HF) or Dirac–Fock (e.g., refs. [29], [30], [31], [32]) and the various multi-configurational levels. In particular, the involved variation of the order of the AO energies in the small-energy valence range was eventually unraveled and is depicted here in Figure 1 for the first four periods of elements. The insight, notably that the rules (1)–(3) are specifically relevant for the s, d, and p elements, respectively, entered only into a few chemistry textbooks (e.g., refs. [33], [34], [35], [36], [37], [38], [39]).
![Figure 1:
Free atomic configuration-averaged 2nd ionization energies |ε| (A+ → A2+) plotted as √|ε| (in √eV) versus element numbers Z, for A = 2He through 36Kr; derived from the NIST data [19]. A+ yields representative orbitals for chemically bonded atoms (usually slightly contracted in comparison to those of the free neutral atoms). |ε| corresponds to the negative of the one-electron Kohn–Sham (KS) atomic-orbital energy. For increasing Z, the pairs of ns and np energy levels (in blue) descent together rather smoothly, while the d (red; and f, lilac) levels vary in steps. With regard to the energy, for the lightest elements: 3d « 4s (hydrogenic rule 3); from 12Mg to 20Ca: 4s < 3d (the (n+ℓ,n) textbook rule 1); for Ar and K: 3d ≈ 5s; for 3d elements 21Sc to 29Cu: 3d ≲ 4s (the ‘true’ transition element rule 2), and from Zn onward: 3d « 4s (for the core shells of heavy elements again the hydrogenic rule 3).](/document/doi/10.1515/znb-2021-0139/asset/graphic/j_znb-2021-0139_fig_001.jpg)
Free atomic configuration-averaged 2nd ionization energies |ε| (A+ → A2+) plotted as √|ε| (in √eV) versus element numbers Z, for A = 2He through 36Kr; derived from the NIST data [19]. A+ yields representative orbitals for chemically bonded atoms (usually slightly contracted in comparison to those of the free neutral atoms). |ε| corresponds to the negative of the one-electron Kohn–Sham (KS) atomic-orbital energy. For increasing Z, the pairs of ns and np energy levels (in blue) descent together rather smoothly, while the d (red; and f, lilac) levels vary in steps. With regard to the energy, for the lightest elements: 3d « 4s (hydrogenic rule 3); from 12Mg to 20Ca: 4s < 3d (the (n+ℓ,n) textbook rule 1); for Ar and K: 3d ≈ 5s; for 3d elements 21Sc to 29Cu: 3d ≲ 4s (the ‘true’ transition element rule 2), and from Zn onward: 3d « 4s (for the core shells of heavy elements again the hydrogenic rule 3).
1.2 Bonded atoms
One origin of these problems is that the dominant atomic electron configuration may be different for the four situations: (i) atoms delocally bound in metallic substances at high coordination numbers (CNs), (ii) atoms with two-center bonds in common chemical coordination compounds at intermediate CN, (iii) atoms in small molecules with low CN (e.g., diatomics), or (iv) unbound atoms in vacuum. Further, chemists (and physicists) often apply the same word ‘element’ to three different objects and concepts: (1) the free atoms of element number Z, (2) all the various pure allotropes and phases of bound atoms Z, (3) the ‘abstract’ elements of atoms Z in various chemical compounds, conserved in chemical transmutations, and sometimes even for (4) a sole nucleus less or more dressed by electronic shells. IUPAC’s Gold book [40] explicitly defines the ‘chemical element’ as either a blend of the 1st and 3rd concept, “A species of atoms Z”, bound or unbound, or as the 2nd concept, “A pure chemical substance Z”.
When chemistry of the elements flourished again after WWII, a successful comprehensive model for d-metal complexes was developed (ligand field theory), based on rule (2) [33], [34], [35], [36]. However, the main drawback of the (n+ℓ,n) rule (1) is not the ‘improvable’ (n−1)d > ns energy pattern but the missing indication of the chemically most important large orbital energy gaps. In rules (2) and (3), the symbol “≪” indicates those large gaps (i.e., large in comparison to chemical bond energies). The large hydrogen-like (n−1)sp ≪ n(sp) gap distinguishes the closed (n−1)(sp)8 noble gas shells, i.e., chemically inert-core shells from the chemically active n(sp) g valence shells of elements in group g. This gap is the physical reason and origin of the chemical periodicity of the post-noble gas elements under ambient chemical conditions [41], [42], [43]. A large gap of type (n−1)spd < n(sp) also develops after the d 10-shell filling (in periods n ≥ 4), from group 12 onward, causing the double periodicity of main and transition groups.
Actually, there are many examples for different dominant electron configurations of an ‘element’ in different environments (compounds), corresponding to the different concepts of an ‘element’. For instance, the dominant configuration of the free Ni atom in the ground state is 3d 8 4s 2 (just 2½ kJ mol−1 below a 3d 9 4s 1 state), of Ni in the metallic phase it is 3d 9 4s 1 ; in many small (diatomic) molecules it is also 3d 9 4s 1 , but in the tetra-carbonyl nickel Ni(CO)4 complex is 3d 10 4s 0 , and for oxidized NiII in common nickel salts such as [Ni(H2O)6]Cl2 it is 3d 8 4s 0 [44], [45], [46], [47], [48], [49], [50], [51], [52], [53]. In his early plots of AO energies versus the element numbers, Bohr was not sure at which element number the valence (n–1)d level would fall below the valence ns level [1, 2]. Löwdin concluded his famous article on the Periodic Table’s centenary of 1969 [54] with the ‘most important’ question, at which degree of oxidation the electron configuration of an element would change. Yet, discourses on the Periodic Table usually cite another note of Löwdin, namely that a generally valid (n+ℓ,n) rule has not been derived from the Schrödinger equation. Remarkably this fault is not assigned to the inadequacy of the (n+ℓ,n) rule, but often to the incompleteness of the Schrödinger equation in the chemical realm. It is often not considered that the total state energies sensitively depend on both the one-electron orbital energies (negative due to the dominating nuclear attraction) and the repulsive Coulomb interactions (positive) between all the electrons in spin-orbital pairs coupled in various manners among each other.
In the 1930s–1950s it was not yet clear how many electrons were to be accommodated in the inner and the outer s, p, and d AOs of the transition metal atoms in their compounds [55, 56]. Pauling was influential, advertising 3d-4s-4p and even 4d AO populations of 3d transition elements, despite the spatial diffuseness of the 4s,p,d orbitals. This narrative is rather slowly disappearing from the textbooks. Chemists tended to assume that the, somewhat irregular, dominant electronic configurations of the ground states of the free atoms have some relevance for the chemically bound atoms. There are 27 (3 × 9), 3d, 4d, and 5d elements between the d 0 s 2 and d 10 s 2 groups 2 and 12. The free atomic ground states of 17 of these d elements are derived from (n−1)d g−2 ns 2 configurations, following the (n+ℓ,n) rule. The 10 ‘exceptions’ from the rule of the free atomic states (Cr, Cu; Nb, Mo, Ru to Ag; Pt, Au) are mentioned in most elementary textbooks, e.g., refs. [20], [21], [22], [23].
During the 1950s, transition metal chemists found that the dominant electron configurations of the metal ions M q+ (q = formal oxidation state) in common transition metal complex species, where the metal atom is surrounded by a set of closed-shell ligands (Lewis bases), are 3d g−q 4s 0 [33], [34], [35], [36], [37], [38], [39, 41], [42], [43, 50], [51], [52], [53]. The situation is different in solid metallic phases, where the metal atoms are coordinated by more ‘ligands’, though of the same ‘metallic character’ with small occupied core shells, with normal and diffuse valence shells, and without dative electron pairs [57], [58], [59], [60], [61], [62], [63].
We here investigate the dominant configurations of the elements of groups g = 2–12. We compare the free atoms M1, the atomic clusters M m (m = 1–13), a few complex molecules ML6, and the metallic phases [M∞]. Professor Dronskowski’s LOBSTER AO-Population tool [64, 65] is particularly helpful to answer our conceptual questions concerning the dominant electron configurations of the transition metals in different chemical environments.
2 Computational results and analysis of experimental data
2.1 Free atoms
We discuss some experimental ‘orbital’ energies ε exp, as displayed in Figure 1. Figure 2 presents the values of free transition metal atoms (early and late ones, light and heavy ones). ε exp is defined as the configuration-averaged single-electronic detachment energy from one orbital without changing the population of the other orbitals (ionization energy, IE; corresponding to the work function of solids). Four points must be noted.
![Figure 2:
Configuration-averaged one-electron ionization energies (IE in eV) of neutral transition metal atoms (Sc, La and Ni, Pt) derived from NIST atomic spectral data [19]. The d levels (red) vary more than the s levels (blue), in particular for the 3d series. The orbital energy patterns strongly depend on the d
g−x
s
x
orbital population distributions.](/document/doi/10.1515/znb-2021-0139/asset/graphic/j_znb-2021-0139_fig_002.jpg)
Configuration-averaged one-electron ionization energies (IE in eV) of neutral transition metal atoms (Sc, La and Ni, Pt) derived from NIST atomic spectral data [19]. The d levels (red) vary more than the s levels (blue), in particular for the 3d series. The orbital energy patterns strongly depend on the d g−x s x orbital population distributions.
First, the valence shell contains energetically near-degenerate orbitals of rather different spatial size, the compact (n−1)d and the diffuse ns orbitals. As a rule of thumb, the radius of ns is ca. three times larger than that of (n−1)d [17, 32]. The orbital energies of the compact (n−1)d orbitals change significantly for different d-s population schemes, they change more than the diffuse ns orbitals, in particular for the 3d elements, and for the late d elements. The (n−1)d ≷ ns orbital energy order may also change for different population schemes such as d g s 0, d g−1 s 1, d g−2 s 2. Hence, there is no general fixed orbital energy pattern for the (n−1)d-ns valence shell of the d-block elements!
Second, to derive sensible ε exp values, ionization processes need to be selected where no significant orbital rearrangement happens. That is not necessarily the lowest energy ionization, which is usually displayed in the textbooks. For instance, the 1st ionization of several transition metal atoms, such as V, Co, or Ni, occurs with a d-s population change. As an example, the “first” ionization of Ni is Ni0-3F4 e(3d 84s 2 ) → Ni+-5D5/2 e(3d 94s 0 ). However, for the ε exp values of Ni-3d, and Ni-4s, respectively, the ionization processes Ni0-3F4 e(3d 84s 2) → Ni+-4F4.5 e(3d 74s 2, and 3d 84s 1) are the appropriate ones.
Third, the experimentally derived orbital energies (and the exact KS density-theoretical ones [66]) cover also the relaxation of the final cation, what reduces the ε exp value, while the self-consist calculations of orbital energies refer to the orbital charge distributions in the initial state and do not cover the ionic orbital relaxation effect. Note that common density-functional approximations usually yield reasonable molecular structures and energies, while the orbital energies are much smaller in value than the ionization energies, mainly due to the incompletely corrected ‘self-interaction error’ [66].
Forth, atoms, clusters, and molecules each have their specific spatial symmetries, which lead to specific orbital level splittings and specific orbital populations. For instance, the most stable state of the half-filled d shell of a free atom in a vacuum (of O3 symmetry) is 6S5/2 e- d 5 , while for the atom in a strong ligand field (of Oh symmetry), the particularly stable states are 4A1g(3/2)- d(t 2g ) 3 and 1A1g(0)- d(t 2g ) 6 . Hence, which AO population numbers are preferred depends on the atomic environment. Accordingly, the orbital coupling of the free atomic ground state is of less relevance for chemistry in general. In any case it is better to average over the different states of a given configuration. Hence, we display the configuration-averaged one-electron ionization energies; in particular we average over the spin-orbit split levels.
The staircase-like variation of the d levels with increasing Z versus the more smooth hydrogen-like variation of the pair of s and p levels is shown in Figure 1. We deduce the following fuzzy empirical rule for the free atoms of d-block elements: “energetically (n–1)d ≈ ns at the beginning of the d series, with the tendency toward (n–1)d < ns at the end”. Pilar even claimed “4s is always above 3d”, in contrast to the textbook rule ([67], however [68]).
We here note that ε exp[(n−1)d] < ε exp[ns] in particular for the cases where the lowest energy configuration is (n−1)d g−2 ns 2. Small d population means small d-d repulsion and low d-orbital energy. Apparently, caution is recommended when comparing experimental ionization energies (the configuration averaged ones, and even more so referring to the more ‘directly observed’ ones between individual states) with the orbital energies of orbital models.
There are no apparent violations of the correct Aufbau rules concerning the orbital populations in single-configuration models of the ab-initio (HF or Dirac–Fock) or density-functional (KS) type. For the ab-initio orbital model “the lower orbital levels are fully occupied at first, provided the orbital energy gap is bigger than the two-electron interaction correction”. We stress that the one-electron energy values have different meanings concerning: the experimental ionization; the average-configuration derived values; the ab-initio HF approach; the exact density-functional KS approach; the common KS approximations [66]. For instance, the HF orbital energies do not account for Coulomb correlation nor for cationic relaxation, which often partially cancel each other. The common KS approximations suffer from incomplete self-interaction corrections. If (n−1)d g−2 ns 2 is the lowest energy configuration (as in many free d atoms), yet ε exp [(n−1)d] < ε exp [ns] and ε HF [(n−1)d] < ε HF [ns] may hold (violation of the simple Aufbau rule), but ε KS,appox [ns] < ε KS,approx [(n−1)d] usually holds (Janak’s rule). Indeed, we find with the PBE density functional for Sc-3d 14s 2: 4sα (−4.3 eV) < 4sβ (−4.0 eV) < 3d 1α (−3.5 eV) (see also Supplementary Material available online, Sect. S3).
2.2 Clusters of transition metal atoms
For transition metals M, we compare the atom M1 and the close-packed solid [M∞] with cluster molecules M m , m = 5, 9, and 13 (the latter has the ‘metallic’ CN = 12 for the central atom). The geometric structures are displayed in Figure 3, the calculated bond lengths and orbital populations are shown in Table 1. We have chosen two 3d elements with atomic ground configurations 3d g−24s 2, namely Mn (g = 7) and Ni (g = 10).

Ball-and-stick models of clusters of transition metal atoms (M) for M = Mn or Ni. In the M13 (D 3d) cluster, the red (thin-blue) lines represent shorter (longer) interatomic distances.
Clusters of transition metal atoms M m : total spin quantum number S, D M–M distances (in pm). Average coordination numbers CNave, average Mulliken orbital populations of the 3d, 4s, and 4p orbitals.a
Cluster | S | D M–M | CNave | 3d ave | 4s ave | 4p ave |
---|---|---|---|---|---|---|
Mn1 | 2.5 | – | 0 | 5.00 | 2.00 | 0.00 |
Mn5 | 12.5 | 272 | 0.8 | 5.31 | 1.36 | 0.33 |
Mn9 | 2.5 | 237 | 4.44 | 5.64 | 0.86 | 0.55 |
Mn13 | 2.5 | 252, 262 | 6.46 | 5.66 | 0.79 | 0.55 |
[Mn∞] | 0 | 244 | 12 | 5.99 | 0.60 | 0.41 |
|
||||||
Ni1 | 1 | – | 0 | 8.00 | 2.00 | 0.00 |
Ni5 | 5 | 218 | 0.8 | 8.61 | 1.02 | 0.37 |
Ni9 | 1 | 216 | 4.44 | 8.80 | 0.76 | 0.44 |
Ni13 | 1 | 236, 240 | 6.46 | 8.83 | 0.68 | 0.49 |
[Ni∞] | 0 | 244 | 12 | 8.97 | 0.69 | 0.34 |
-
aObtained by density functional calculations with the PBE potential at the scalar relativistic ZORA level using Slater-type TZP basis sets. Metallic phases [M∞] see Supplementary Material, Sect. 2.3.
From the single atom to the metallic phase, the 3d population increases by ca. one unit, while the 4s population decreases, 4s partially hybridized with 4p. These trends may be rationalized as follows. The slight overlap of the rather compact 3d orbitals causes bonding by populating the ‘band’ of d orbital levels. The 4s-4p population is more complex. These diffuse AOs strongly overlap with each other, yielding bonding and strongly antibonding levels. In addition there is an overall Pauli repulsion by the occupied core shells of the adjacent metal atoms. Only the bottom of the s-p band of molecular orbitals is bonding and low in energy and will be occupied. Due to the large overlap values, in particular for high CNs, the atomic reference basis becomes nearly degenerate, and the AO population values sometimes become numerically unstable, i.e., inaccurate.
In summary, while the single atoms in vacuum have a tendency toward high population of the diffuse ns orbital (paradigmatic 3d g−24s 2), the clustering of the metal atoms toward the metallic phase tends to medium sp population (paradigmatic 3d g−14(sp)1) with filling the lower, bonding half of the s-type band.
2.3 Solid transition metals
Three different low-energy crystal structures were investigated for the solid metals Ca to Zn of groups g = 2–12 in period 4 (and for the heavier homologs, yielding similar trends, see the Supplementary Material, Sect. S1): the hexagonal close-packed (hcp) one, the cubic close-packed (ccp = fcc, face centered cubic) one, both with (CN = 12), and the body-centered cubic (bcc) one (CN = 8); for technical details see the Methodology (Sect. 4, below).
Figure 4 displays the quantum-chemically optimized interatomic nearest neighbor distances (at T = 0 K and p = 0 bar). The experimentally determined values (for STP) of the 4th, and also of the 5th and 6th periods are very similar (see Sect. 4 of the Supplementary Material) (For the calculations of Mn at 0 K, we chose the high-temperature fcc structure instead of the unusually complex, highly coordinated, spatially expanded structure at STP). Four points are not new, yet noteworthy: (i) The quantum chemical calculations correctly reproduce the empirical trends. (ii) The inter-atomic distances vary smoothly with group number, the shortest ones occur around groups 8 and 9. (iii) The irregularity of the electronic configurations of the ground levels of the free atoms ((n−1)d g−2 ns 2 or (n−1)d g−1 ns 1 or (n−1)d g ns 2) appears not to affect the bond distances, and also not the melting and boiling temperatures of the solid phases (Supplementary Material, Sects. S1.3 and S1.4). (iv) The secondary periodicity is obvious, meaning that period four differs from the mutually rather similar periods 5 and 6.

Solid transition metals: Calculated (for T = 0 K) interatomic distances (in pm) of the most stable hcp, ccp, or bcc structure (but for Mn, the close-packed high-T structure at T = 0 K, instead of the complex, less-dense low-T structure). Fourth period: black squares ■; 5th period: red circles ; 6th period: blue triangles
; all with eye-guiding connections (no relation to the d
g−2
s
2 vs. d
g−1
s
1 configurational differences of the ground states of the free atoms is visible).
The orbital populations of the atoms in the metallic phases of the elements in the 4th period are shown in Figure 5 (for more details and for the heavier homologs, see the Supplementary Material, Sect. S1.2). Again, the variation is rather smooth. If the occupied valence band structure is projected onto the 3d and 4s AOs, one obtains for the metals Sc to Cu approximately (within ±0.2) 3d g−0.94s 0.9, and for the projection onto 3d, 4s, and 4p, approximately 3d g−1.44(sp)1.4. Calcium and Zink from groups 2 and 12 differ a little. The Ca metal with ca. 3d 0.6 has slightly more 4s population, but significantly less than the 3d 04s 2 of the free atom. Zn metal has a filled 3d 10 core shell with only a little additional 4d polarization of the 4sp valence band. Our AO populations agree within ca. ±0.2 with the ones in the review book of Papaconstantopoulos, giving on the average 3d g−1.05 4s 0.65 4p 0.4 [69, 70].

Atomic orbital (AO) Mulliken populations (3d, 4s, 4s, and 4p) of face-centered cubic phases of 4th row metals Ca to Zn (From density functional calculations with the SCAN+rVV10 meta-gradient-approximation with dispersion). The full (open) circles connected by full (dotted) lines refer to calculations with (without) 4p orbitals for the population analysis, respectively. Other density functionals, other cubic and hexagonal phases, yield similar trends (see the Supplementary Material, Sect. S1.3).
The s,p,d populations in the transition metals may be derived from various spectroscopic data or from band structure computations. The literature data scatter quite a bit (e.g., refs. [47], [48], [49]). In some cases small or even negative ns populations were reported, and in other cases also higher populations of ns 1.5 to ns >2. This (non-systematic) scattering of deduced parameters does not really prove that the atoms at high CN = 12 in hexagonal or cubic solid environment follow the d,s population behavior of free atoms in vacuum with CN = 0.
Figure S5 in the Supplementary Material displays good correlations between the calculated cohesive energies and the boiling and melting temperatures of the metallic elements. Figure S6 displays the variation of the melting and boiling temperatures along the series of transition metals. Both Figures indicate an increasing bond strength where the s(p) bonding of the metals of groups 2 or 12 is supported by d bonding, which is strongest for group 6 metallic phases with dominant (n−1)d 5 ns 1 configuration and with half-filled valence bands. There is no apparent correlation with the d g s 0, d g−1 s 1, d g−2 s 1 configurations of the free, chemically isolated atoms.
2.4 Some transition metal complexes
Since the development of the crystal and ligand field models before and after WWII, transition metal chemists describe the lower electronic states of complexes of metal cations M q+ by the dominant configuration (n−1)d g−q ns 0. We have reproduced this concept for three complexes, formally with 5, 6, and 7 electrons in the valence shell of the metal atom, using the same computational approaches as applied so far.
For Td symmetric Na4[MnCl6], we obtain a spin sextet and populations Mn-3d 5.494s 0.124p 0.66, meaning Na+ 4[Mn2+ (3d 54s 0)Cl− 6] where the Cl− ligands donate ca. 6 × 0.5 into the half-populated 3d shell, and ca. 0.1 and 0.7 e into the diffuse 4s and 4p shells, respectively. For the geometric structure, effective charges and spin populations of this high-spin and the respective excited low-spin complexes, see the Supplementary Material, Sect. S2.
For Oh symmetric [Cr(CO)6], we obtain a spin singlet and populations Cr-3d 5.334s 0.004p 0.77, meaning [Cr0 (3d 64s 0) (CO0)6]0, where the six CO ligands donate ca. 2e from their C-2pσ lone-pairs into the empty 3d(eg) and 4p(t1u) orbitals, while ca. 2e from the filled 3d(t2g) orbitals are back-donated into the empty CO-π* orbitals.
Adding 1 e, we obtain the Oh symmetric spin doublet complex anion [Cr(CO)6]− with populations Cr-3d 5.294s 1.04 4p 0.76, meaning [Cr− (3d 64s 1) (CO0)6]− (see also the Supplementary Material, Sect. S2). The attached electron populates the diffuse Cr-4s orbital, a typical phenomenon for formally anionic transition metal atoms. In contrast, cationic metal complexes with formally seven valence electrons are d 7 s 0 systems. Examples are the Rh2+ or Co2+ complexes, e.g., the spin quartet [Co2+ (3d 74s 0) (H2O,NH3)0 6]2+ [44], [45], [46].
3 Our messages
Quantum chemistry explains the currently accepted structure of the Periodic Table of elements. Most common textbook presentations that determine the chemical convictions of the majority of chemists, physicists, educators, and philosophers, have so far only partially absorbed this wisdom.
The physical origin of the chemical periodicity. It is due to the ‘chemically’ large energy gap Δε (indicated by ≪ in relation 4) between the sets of still rather hydrogen-like one-electron (n−1)s,p and (n)s,p states in many-electron atoms. The order of s,p orbital energies for the seven periods of the periodic system is very systematic and given by relation (4). Educators should prefer this order (Here we do not comment on the decreasing gap sizes above the filled n(sp)8 ‘noble gas’ shells for increasing n, what causes the end of the common chemical periodicity around periods 7 and 8).
The sp block. Relation (4) is the reason for only one period of length 2, but more periods of length 8 (and 18 and 32). The sets of adjacent ns,np shells give rise to a joint block of sp elements with s-p mixing tendency of bonded atoms. The extent of hybridization is mainly determined by the atomic core and valence s and p orbital radii, besides the energies [71].
The d (and f) blocks. The centrifugal force ∼ ℓ(ℓ+1)/r 3 at distances r from the nucleus is small for ns and np orbitals (ℓ = 0 and 1), but significant for the higher angular momentum nd and nf orbitals (ℓ = 2 and 3). Hence, electrons in nd (and nf) orbitals are efficiently screened from nuclear attraction by the inner electron density. The nd (and nf) orbitals are less stabilized for increasing element or nuclear-charge number Z, as long as the inner-more 1s to np shells are populated (Figure 1). The np shells are populated directly after the ns shells; but the nd shells (n ≥ 3) are not populated directly after the ns and np shells. When for increasing Z, the (n−1)(sp)8 noble gas shell has just been filled up, the (n–1)d level is energetically still much higher in groups 0 and 1, namely near the (n+1)s level! However, when the ns shell is filled, the (n−1)d level begins to ‘collapse’ for increasing Z, i.e., it contracts spatially and decreases energetically down to or below the ns level (Figure 1) [72]. Therefore, the (n−1)d series with 10 elements is inserted near the beginning of the nsp series (i.e., after group 2), after the first two periods of 8 (and the (n−3)f series with 14 elements is inserted near the beginning of the (n−1)d series (i.e., after group 3), after the first two periods of 18).
Various d AO energy patterns. The first nine transition elements in groups g = 3–11 have g valence electrons in the d,s valence shell, with ε((n–1)d) ≈ ε(ns) at the beginning of the d block for neutral non-confined atoms, but ε((n–1)d) < ε(ns) for the later d atoms (Figure 2), and for the ionized and the ligated d atoms. Only in negatively charged atoms, there is ε(ns) < ε((n–1)d). Eventually, ε((n–1)d) < ε(ns) in group 12, and hence (n–1)(spd)18 becomes the chemically inert atomic core shell from group 12 (inclusive) onward. A chemically more useful rule of thumb for the orbital energies is relation (5).
Rational enlightenment accepts the evidence of nature and the sovereignty of human reason, which starts from the mathematical simplicity of the basic physical laws and leads to the intricacy of the emerging complexity of chemical matter. There is no empirical or theoretical proof of a general (n+ℓ,n) rule in chemistry. The rule correctly specifies for which group of elements the d orbital first becomes significantly populated. However, it neither specifies where the large, periodicity determining orbital energy gaps appear (rule 4), nor does it correctly inform on the variation of the orbital energy patterns (rule 5).
Orbital populations. If one knows where the large AO energy gaps occur (between the sp shells of different principal quantum number), one may apply the simplistic Aufbau rule to predict or memorize the dominant atomic electron configurations that govern the chemistry of the main group elements. It depends on the core and valence orbital energies and radii. However, for the valence shells of the more or less adjacent s and d (and f) orbitals of the transition metals M, the simplistic Aufbau rule fails too often. The d-s orbital populations depend more on the chemical environment than on the irregularities specific for the free atoms. More important in chemistry than teaching some questionable orbital energy rule is teaching the actual orbital populations in compounds, namely:
The diffuse ns orbitals of most cations Mq+ ligated by Lewis bases are destabilized and non-populated (ns 0). Although the ns level is slightly above the (n−1)d level, it becomes populated in metallic phases due to the stabilizing delocalized metallic bonding (ca. ns 1). The free transition metal atoms do not serve well as a paradigm for atomic behavior in chemical compounds including the metallic phases (Figures 4 and 5).
4 Computational methodology
Calculations with the self-consistent spin-polarized KS Density-Functional Theory (DFT) were performed, applying the Amsterdam Density Functional (ADF) [73] package for the molecules (metallic clusters; metal complexes) and the Vienna Ab-initio Simulation Package (VASP) [74] for the solid metals.
Three types of exchange-correlation functionals were used for comparison, i.e., the Local-Spin-Density-Approximation (LSDA) [75], the Generalized-Gradient-Approximation (GGA) of the Perdew–Burke–Ernzerhof (PBE) functional [76], and the meta-Generalized-Gradient-Approximation (meta-GGA) with the new SCAN functional and the long-range van-der-Waals interaction rVV10 form Perdew et al. [77, 78].
For the atoms, clusters and complexes, the triple-Zeta Slater-type basis sets were used [79]. The zero-order regular approximation [80] for the scalar relativistic effects was applied. The frozen core approximation was applied for the [1s 2] inner shells of the O, C, Na, and Cl atoms, and for the [1s 2-2p 6] inner shells of the Cr, Mn, Ni, and Sc atoms in the calculated molecules.
For the solids, the Γ-centered Monkhorst-Pack k -meshes 12 × 12 × 12, 10 × 10 × 10, and 17 × 17 × 8 were used, respectively, for the body-centered cubic close-packed (bcc/ccp), the face-centered cubic (fcc) and the hexagonal close packed (hcp) structures. The 18 × 18 × 18, 14 × 14 × 14, and 23 × 23 × 11 k -meshes, respectively, were used in the population analyses [81]. The Projector-Augmented-Wave (PAW) projector-pseudo-potentials were applied, treating 1s 22s 22p 6 as frozen core shells for Ca to Ti, 1s 22s 22p 6+3s 2 for Cr to Mn, and 1s 22s 22p 63s 2+3p 6 for Fe to Zn; and 3s 23p 6(3d4s4p) g , 3p 6(3d4s4p) g , and (3d4s4p) g , respectively, as self-consistently optimized semi-core and valence shells [82]. A plane-wave basis energy-cutoff of 600 eV was applied. Structural optimizations were achieved with a threshold of energy difference of less than 10−6 eV and all forces smaller than 10−3 eV/Å2 [83], [84], [85], [86].
The solid-phase population analyses were performed with the LOBSTER program package, with reference to the “pbevaspfit2015” contracted Slater-type basis set [85, 86], starting from the self-consistency calculations, but using a denser k -mesh (see above) and with the symmetry restriction switched off.
5 Supporting information
S1: Bond distances, populations, cohesive energies, and melting and boiling temperatures of transition metals. S2: Structures and populations of transition metal complex molecules. S3: Configurations and orbital energies of the Sc atom are given as supplementary material available online (https://doi.org/10.1515/znb-2021-0139).
Dedicated to: Professor Richard Dronskowski of the RWTH Aachen on the occasion of his 60th birthday.
Funding source: Alexander von Humboldt Foundation
Funding source: RWTH Aachen University
Award Identifier / Grant number: JARA0179
Funding source: Computational Chemistry Laboratory of the Department of Chemistry
Acknowledgments
We feel deeply obliged to Professor Dronskowski for his contributions during the early stages of this work. We thank J. Autschbach, E. J. Baerends, G. Frenking, J. Li, P. Pyykkö, P. Salvador, E. Scerri, P. Schwerdtfeger, R. Vernon, and the two unknown reviewers for their constructive comments. WHES thanks for hospitality and support by the research groups in Beijing and Siegen.
-
Author contributions: All authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: KC gratefully acknowledges the financial support from the Alexander von Humboldt Foundation. The computational work was supported by the IT center of RWTH Aachen University under Grant JARA-HPC (JARA0179). WLL thanks for support by the Computational Chemistry Laboratory of the Department of Chemistry under the Tsinghua Xuetang Talents Program.
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
1. Bohr, N. The Structure of the Atom. Noble Lectures, 1922; pp. 37. www.nobelprize.org/prizes/physics/1922/bohr/lecture/.Search in Google Scholar
2. Bohr, N., Coster, D. Röntgenspektren und periodisches System der Elemente. Z. Phys. 1923, 12, 342–374; https://doi.org/10.1007/bf01328104.Search in Google Scholar
3. Condon, E. U., Shortley, G. H. The Theory of Atomic Spectra; Cambridge University Press: London, 1935.Search in Google Scholar
4. Bloch, F. Über die Quantenmechanik der Elektronen in Kristallgittern. Z. Phys. 1929, 52, 555–600.10.1007/BF01339455Search in Google Scholar
5. Sommerfeld, A., Frank, N. H. The statistical theory of thermoelectric, galvano- and thermomagnetic phenomena in metals. Rev. Mod. Phys. 1931, 3, 1–42; https://doi.org/10.1103/revmodphys.3.1.Search in Google Scholar
6. Heitler, W., London, F. Wechselwirkung neutraler Atome und homöopolare Bindung nach der Quantenmechanik. Z. Phys. 1927, 44, 455–472; https://doi.org/10.1007/bf01397394.Search in Google Scholar
7. Hund, F. Zur Frage der chemischen Bindung. I & II. Z. Phys. 1932, 73, 1–30 and 565–577; https://doi.org/10.1007/bf01342005.Search in Google Scholar
8. Schwarz, W. H. E. Measuring orbitals: provocation or reality? Angew. Chem. Int. Ed. 2006, 45, 1508–1517; https://doi.org/10.1002/anie.200501333.Search in Google Scholar PubMed
9. Zhao, L., Pan, S., Holzmann, N., Schwerdtfeger, P., Frenking, G. Chemical bonding and bonding models of main-group compounds. Chem. Rev. 2019, 119, 8781–8845; https://doi.org/10.1021/acs.chemrev.8b00722.Search in Google Scholar PubMed
10. Zhao, L., Schwarz, W. H. E., Frenking, G. The Lewis electron-pair bonding model: the physical background, one century later. Nat. Rev. Chem. 2019, 3, 35–47; https://doi.org/10.1038/s41570-018-0052-4.Search in Google Scholar
11. Bacskay, G. B., Nordholm, S., Ruedenberg, K. The virial theorem and covalent bonding. J. Phys. Chem. 2018, 122, 7880–7893; https://doi.org/10.1021/acs.jpca.8b08234.Search in Google Scholar PubMed
12. West, A. C., Schmidt, M. W., Gordon, M. S., Ruedenberg, K. Intrinsic resolution of molecular electronic wave functions and energies in terms of Quasi-atoms and their interactions. J. Phys. Chem. 2017, 121, 1086–1105; https://doi.org/10.1021/acs.jpca.6b10911.Search in Google Scholar PubMed
13. Schmidt, M. W., Ivanic, J., Ruedenberg, K. The physical origin of covalent bonding. In The Chemical Bond. Fundamental Aspects; Frenking, G., Shaik, S., Eds., Wiley VCH: Weinheim, 2014, pp. 1–67.10.1002/9783527664696.ch1Search in Google Scholar
14. Schmidt, M. W., Ivanic, J., Ruedenberg, K. Covalent bonds are created by the drive of electron waves to lower their kinetic energy through expansion. J. Chem. Phys. 2014, 140, 204104.https://doi.org/10.1063/1.4875735.Search in Google Scholar PubMed PubMed Central
15. Ruedenberg, K., Schmidt, M. W. Physical understanding through variational reasoning: electron sharing and covalent bonding. J. Phys. Chem. A 2009, 113, 1954–1968; https://doi.org/10.1021/jp807973x.Search in Google Scholar PubMed
16. Bitter, T., Ruedenberg, K., Schwarz, W. H. E. Toward a physical understanding of electron-sharing two-center bonds. I. general aspects. J. Comput. Chem. 2007, 28, 411–422; https://doi.org/10.1002/jcc.20531.Search in Google Scholar PubMed
17. Frenking, G., Fröhlich, N. The nature of the bonding in transition-metal compounds. Chem. Rev. 2000, 100, 717–774; https://doi.org/10.1021/cr980401l.Search in Google Scholar PubMed
18. Moore, C. E. Atomic Energy Levels as Derived from the Analyses of Optical Spectra, Vol. 3; NBS: Washington DC, 1949–1958.Search in Google Scholar
19. Kramida, A., Ralchenko, A, et mult. al. NIST Standard Reference Database #78. Atomic Spectra. www.nist.gov/pml/atomic-spectra-database (accessed Sep, 2021).Search in Google Scholar
20. Huheey, J., Keiter, E., Keiter, R. Inorganic Chemistry, 4th ed.; Prentice-Hall: Hoboken NJ, 1993.Search in Google Scholar
21. Brown, T. L., LeMay, H. E., Bursten, B. E., Murphy, C. J., Woodward, P. M., Stoltzfus, M. W. Chemistry The Central Science, 13th ed.; Pearson: Boston, 2015.Search in Google Scholar
22. Miessler, G. L., Fischer, P. J., Tarr, D. A. Inorganic Chemistry, 5th ed.; Pearson: Upper Saddle River NJ, 2014.Search in Google Scholar
23. Atkins, P. W., Jones, L. L., Laverman, L. E. Chemical Principles, 7th ed.; Freeman: New York, 2016.Search in Google Scholar
24. Simmons, L. M. The display of electronic configuration by a periodic table. J. Chem. Educ. 1948, 25, 658–661; https://doi.org/10.1021/ed025p658.Search in Google Scholar
25. Eichinger, J. W. Teaching electron configurations. J. Chem. Educ. 1957, 34, 504–505; https://doi.org/10.1021/ed034p504.Search in Google Scholar
26. Wong, D. P. Theoretical justification of Madelung’s rule. J. Chem. Educ. 1979, 56, 714–717; https://doi.org/10.1021/ed056p714.Search in Google Scholar
27. Carpenter, A. K. 4s, 3d, what? J. Chem. Educ. 1983, 60, 562.https://doi.org/10.1021/ed060p562.Search in Google Scholar
28. Stewart, P. J. From telluric helix to telluric remix. Found. Chem. 2020, 22, 3–14.10.1007/s10698-019-09334-7Search in Google Scholar
29. Latter, R. Atomic energy levels for the Thomas-Fermi and Thomas-Fermi-Dirac potential. Phys. Rev. 1955, 99, 510–519; https://doi.org/10.1103/physrev.99.510.Search in Google Scholar
30. Herman, F., Skillman, S. Atomic Structure Calculations; Prentice-Hall: Englewood Cliffs NJ, 1963.Search in Google Scholar
31. Clementi, E. Tables of Atomic Functions; IBM: San Jose, CA, 1965.10.1147/JRD.1965.5392159Search in Google Scholar
32. Desclaux, J. P. Relativistic Dirac-Fock expectation values for atoms with Z = 1 to Z = 120. Atomic Data Nucl. Data Tables 1973, 12, 311–406; https://doi.org/10.1016/0092-640x(73)90020-x.Search in Google Scholar
33. Jørgensen, C. K. Energy Levels of Complexes and Gaseous Ions; Gjellerups Forlag: Copenhagen, 1957.Search in Google Scholar
34. Jørgensen, C. K. Oxidation Numbers and Oxidation States; Springer: Berlin, 1969.10.1007/978-3-642-87758-2Search in Google Scholar
35. Ballhausen, C. J. Introduction to Ligand Field Theory; McGraw-Hill: New York, 1962.Search in Google Scholar
36. Orgel, L. E. An Introduction to Transition-Metal Chemistry: Ligand-Field Theory; Methuen: London, 1960.Search in Google Scholar
37. Glasstone, S. Textbook of Physical Chemistry; Van Nostrand: New York, 1940.Search in Google Scholar
38. Kauzmann, W. Quantum Chemistry. An Introduction; Academic Press: New York, 1957.10.1016/B978-1-4832-2745-0.50005-6Search in Google Scholar
39. Levine, I. N. Quantum Chemistry; Pearson: Boston, 1970.Search in Google Scholar
40. Gold, V. Compendium of Chemical Terminology. IUPAC Gold Book (version 2.3.2), 2012. goldbook.iupac.org.Search in Google Scholar
41. Cao, C., Vernon, R. E., Schwarz, W. H. E., Li, J. Understanding periodic and non-periodic chemistry in periodic tables. Front. Chem. 2021, 8, 813.https://doi.org/10.3389/fchem.2020.00813.Search in Google Scholar PubMed PubMed Central
42. Cao, C., Hu, H., Li, J., Schwarz, W. H. E. Physical origin of chemical periodicities in the system of elements. Pure Appl. Chem. 2019, 91, 1969–1999; https://doi.org/10.1515/pac-2019-0901.Search in Google Scholar
43. Wang, S. G., Qiu, Y. X., Fang, H., Schwarz, W. H. E. The challenge of the so-called electron configurations of the transition metals. Chem. Eur J. 2006, 12, 4101–4114; https://doi.org/10.1002/chem.200500945.Search in Google Scholar PubMed
44. Goel, S., Masunov, A. E. Dissociation curves and binding energies of diatomic transition metal carbides from density functional theory. Int. J. Quant. Chem. 2011, 111, 4276–4287; https://doi.org/10.1002/qua.22950.Search in Google Scholar
45. Sanati, D. A., Andrae, D. Low-lying electronic terms of diatomic molecules AB (A = Sc–Ni, B = Cu/Ag/Au). Mol. Phys. 2020, 118, e1772514; https://doi.org/10.1080/00268976.2020.1772514.Search in Google Scholar
46. Cheung, L. F., Chen, T.-T., Kocheril, G. S., Chen, W.-J., Czekner, J., Wang, L.-S. Observation of four-fold boron–metal bonds in RhB(BO−) and RhB. J. Phys. Chem. Lett. 2020, 11, 659–663; https://doi.org/10.1021/acs.jpclett.9b03484.Search in Google Scholar PubMed
47. Raj, S., Padhi, H. C., Palit, P., Basa, D. K., Polasik, M., Pawłowski, F. Relative K x-ray intensity studies of the valence electronic structure of 3d transition metals. Phys. Rev. B 2002, 65, 193105.https://doi.org/10.1103/physrevb.65.193105.Search in Google Scholar
48. Uğurlu, M., Demir, L. Relative, K X-ray intensity ratios of the first and second transition elements in the magnetic field. J. Mol. Struct. 2020, 1203, 127458.10.1016/j.molstruc.2019.127458Search in Google Scholar
49. Daoudi, S., Kahoul, A., Aylikci, N. K., Sampaio, J. M., Marques, J. P., Aylikci, V., Sahnoune, Y., Kasri, Y., Deghfel, B. Review of experimental photon-induced Kβ/Kα intensity ratios. Atomic Data Nucl. Data Tables 2020, 132, 101308.10.1016/j.adt.2019.101308Search in Google Scholar
50. Hillier, I. H., Saunders, V. R. Ab initio molecular orbital calculations of transition metal complexes. Mol. Phys. 1971, 22, 1025–1034; https://doi.org/10.1080/00268977100103341.Search in Google Scholar
51. Demuynck, J., Veillard, A. Electronic structure of the nickel tetracyanonickelate Ni(CN)42− and nickel carbonyl Ni(CO)4. An ab-initio LCAO-MO-SCF calculation. Theor. Chim. Acta 1973, 28, 241–265; https://doi.org/10.1007/bf00533488.Search in Google Scholar
52. Böhm, M. C. The electronic structure of closed shell metallocenes in the ground state and in the cationic hole-states. Z. Naturforsch. 1982, 37a, 1193–1204; https://doi.org/10.1515/zna-1982-1011.Search in Google Scholar
53. Weber, J., Geoffroy, M., Goursot, A., Pénigault, E. Application of the multiple scattering Xα molecular orbital method to the determination of the electronic structure of metallocene compounds. 1. Dibenzenechromium and its cation. J. Amer. Chem. Soc. 1978, 100, 3995–4003.10.1002/chin.197839061Search in Google Scholar
54. Löwdin, P.-O. Some comments on the periodic system of the elements. Int. J. Quantum Chem. Symp. 1969, 3, 331–334.10.1002/qua.560030737Search in Google Scholar
55. Pauling, L. The Nature of the Chemical Bond and the Structure of Molecules and Crystals; Cornell University Press: Ithaca NY, 1939.Search in Google Scholar
56. Pauling, L. General Chemistry, an Introduction to Descriptive Chemistry and Modern Chemical Theory; Freeman: San Francisco CA, 1947.Search in Google Scholar
57. Mott, N. F., Jones, H. The Theory of the Properties of Metals and Alloys; Clarendon Press: Oxford, 1936.Search in Google Scholar
58. Mott, N. F., Stevens, K. W. H. The band structure of the transition metals. Philos. Mag. A 1957, 2, 1364–1386; https://doi.org/10.1080/14786435708243213.Search in Google Scholar
59. Goodenough, J. B. Band structure of transition metals and their alloys. Phys. Rev. 1960, 120, 67–83; https://doi.org/10.1103/physrev.120.67.Search in Google Scholar
60. Mattheiss, L. F. Energy bands for the iron transition series. Phys. Rev. 1964, 134, 970–973; https://doi.org/10.1126/science.145.3635.970.Search in Google Scholar
61. Gomès, A. A., Campbell, I. A. Remarks on the electronic structure of transition metal alloys. J. Phys. C Solid State Phys. 1968, 1, 253–264; https://doi.org/10.1088/0022-3719/1/1/328.Search in Google Scholar
62. Pettifor, D. G. Accurate resonance-parameter approach to transition-metal band structure. Phys. Rev. B 1970, 2, 3031–3034; https://doi.org/10.1103/physrevb.2.3031.Search in Google Scholar
63. Engels, S. Zum Begriff des „Valenzelektrons” in metallischen Systemen. Z. Chem. 1969, 9, 161–170.10.1002/zfch.19690090502Search in Google Scholar
64. Ertural, C., Steinberg, S., Dronskowski, R. Development of a robust tool to extract Mulliken and Löwdin charges from plane waves and its application to solid-state materials. RSC Adv. 2019, 9, 29821–29830; https://doi.org/10.1039/c9ra05190b.Search in Google Scholar PubMed PubMed Central
65. Nelson, R., Ertural, C., George, J., Deringer, V. L., Hautier, G., Dronskowski, R. LOBSTER: Local orbital projections, atomic charges, and chemical-bonding analysis from projector-augmented-wave-based density-functional theory. J. Comput. Chem. 2020, 41, 1931–1940; https://doi.org/10.1002/jcc.26353.Search in Google Scholar PubMed
66. Baerends, E. J. Density functional approximations for orbital energies and total energies of molecules and solids. J. Chem. Phys. 2018, 149, 054105.https://doi.org/10.1063/1.5026951.Search in Google Scholar PubMed
67. Pilar, F. L. 4s is always above 3d! Or, how to tell the orbitals from the wavefunctions. J. Chem. Educ. 1978, 55, 2–6; https://doi.org/10.1021/ed055p2.Search in Google Scholar
68. Carlton, T. S. 4s sometimes is below 3d. J. Chem. Educ. 1979, 56, 767; https://doi.org/10.1021/ed056p767.1.Search in Google Scholar
69. Papaconstantopoulos, D. A. Handbook of the Band Structure of Elemental Solids from Z = 1 To Z = 112, 2nd ed.; Springer: New York, 2015.10.1007/978-1-4419-8264-3Search in Google Scholar
70. Sigalas, M., Papaconstantopoulos, D. A., Bacalis, N. C. Total energy and band structure of the 3d, 4d, and 5d metals. Phys. Rev. B 1992, 45, 5777–5783; https://doi.org/10.1103/physrevb.45.5777.Search in Google Scholar PubMed
71. Wang, Z.-L., Hu, H.-S., Von Szentpály, L., Stoll, H., Fritzsche, S., Pyykkö, P., Schwarz, W. H. E., Li, J. Understanding the uniqueness of 2p elements in periodic tables. Chem. Eur J. 2020, 26, 15558–15564; https://doi.org/10.1002/chem.202003920.Search in Google Scholar PubMed PubMed Central
72. Connerade, J. P. Orbital collapse in extended homologous sequences. J. Phys. B Atom. Mol. Opt. Phys. 1991, 24, L109–L115; https://doi.org/10.1088/0953-4075/24/5/001.Search in Google Scholar
73. SCM. ADF. Theoretical Chemistry; Vrijie Universiteit: Amsterdam, 2017.Search in Google Scholar
74. Kresse, G., Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186; https://doi.org/10.1103/physrevb.54.11169.Search in Google Scholar PubMed
75. Kohn, W., Sham, L. J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, 1133–1138; https://doi.org/10.1103/physrev.140.a1133.Search in Google Scholar
76. Perdew, J. P., Burke, K., Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1997, 78. 1396; https://doi.org/10.1103/physrevlett.78.1396.Search in Google Scholar
77. Sun, J., Ruzsinszky, A., Perdew, J. P. Strongly constrained and appropriately normed semilocal density functional. Phys. Rev. Lett. 2015, 115, 036402; https://doi.org/10.1103/physrevlett.115.036402.Search in Google Scholar PubMed
78. Peng, H., Yang, Z.-H., Perdew, J. P., Sun, J. Versatile van der Waals Density Functional Based on a Meta-Generalized Gradient Approximation. Phys. Rev. X 2016, 6, 041005; https://doi.org/10.1103/physrevx.6.041005.Search in Google Scholar
79. Van Lenthe, E., Baerends, E. J. Optimized Slater-type basis sets for the elements 1–118. J. Comput. Chem. 2003, 24, 1142–1156; https://doi.org/10.1002/jcc.10255.Search in Google Scholar PubMed
80. Van Lenthe, E., Baerends, E. J., Snijders, J. G. Relativistic regular two‐component Hamiltonians. J. Chem. Phys. 1993, 99, 4597–4610; https://doi.org/10.1063/1.466059.Search in Google Scholar
81. Monkhorst, H. J., Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192; https://doi.org/10.1103/physrevb.13.5188.Search in Google Scholar
82. Kresse, G., Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775; https://doi.org/10.1103/physrevb.59.1758.Search in Google Scholar
83. Maintz, S., Deringer, V. L., Tchougréeff, A. L., Dronskowski, R. Analytic projection from plane-wave and PAW wavefunctions and application to chemical-bonding analysis in solids. J. Comput. Chem. 2013, 34, 2557–2567; https://doi.org/10.1002/jcc.23424.Search in Google Scholar PubMed
84. Maintz, S., Deringer, V. L., Tchougréeff, A. L., Dronskowski, R. LOBSTER: A tool to extract chemical bonding from plane-wave based DFT. J. Comput. Chem. 2016, 37, 1030–1035; https://doi.org/10.1002/jcc.24300.Search in Google Scholar PubMed PubMed Central
85. Chen, K., Dronskowski, R. First-principles study of divalent 3d transition-metal carbodiimides. J. Phys. Chem. 2019, 123, 9328–9335; https://doi.org/10.1021/acs.jpca.9b05799.Search in Google Scholar PubMed
86. Chen, K., Fehse, M., Laurita, A., Arayamparambil, J. J., Sougrati, M. T., Stievano, L., Dronskowski, R. Quantum-chemical study of the FeNCN conversion-reaction mechanism in lithium- and sodium-ion batteries. Angew. Chem. Int. Ed. 2020, 59, 3718–3723; https://doi.org/10.1002/anie.201914760.Search in Google Scholar PubMed PubMed Central
Supplementary Material
The online version of this article offers supplementary material (https://doi.org/10.1515/znb-2021-0139).
© 2021 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- In this issue
- Laudatio/Preface
- Celebrating the 60th birthday of Richard Dronskowski
- Review
- Orbital-selective electronic excitation in phase-change memory materials: a brief review
- Research Articles
- Solving the puzzle of the dielectric nature of tantalum oxynitride perovskites
- d- and s-orbital populations in the d block: unbound atoms in physical vacuum versus chemical elements in condensed matter. A Dronskowski-population analysis
- Single-crystal structures of A 2SiF6 (A = Tl, Rb, Cs), a better structure model for Tl3[SiF6]F, and its novel tetragonal polymorph
- Na2La4(NH2)14·NH3, a lanthanum-rich intermediate in the ammonothermal synthesis of LaN and the effect of ammonia loss on the crystal structure
- Linarite from Cap Garonne
- Salts of octabismuth(2+) polycations crystallized from Lewis-acidic ionic liquids
- High-temperature diffraction experiments and phase diagram of ZrO2 and ZrSiO4
- Thermal conversion of the hydrous aluminosilicate LiAlSiO3(OH)2 into γ-eucryptite
- Crystal structure of mechanochemically prepared Ag2FeGeS4
- Effect of nanostructured Al2O3 on poly(ethylene oxide)-based solid polymer electrolytes
- Sr7N2Sn3: a layered antiperovskite-type nitride stannide containing zigzag chains of Sn4 polyanions
- Exploring the frontier between polar intermetallics and Zintl phases for the examples of the prolific ALnTnTe3-type alkali metal (A) lanthanide (Ln) late transition metal (Tn) tellurides
- Zwitterion coordination to configurationally flexible d 10 cations: synthesis and characterization of tetrakis(betaine) complexes of divalent Zn, Cd, and Hg
- An approach towards the synthesis of lithium and beryllium diphenylphosphinites
- Synthesis, crystal and electronic structure of CaNi2Al8
- Crystal and electronic structure of the new ternary phosphide Ho5Pd19P12
- Synthesis, structure, and magnetic properties of the quaternary oxysulfides Ln 5V3O7S6 (Ln = La, Ce)
- Synthesis, crystal and electronic structure of BaLi2Cd2Ge2
- Structural variations of trinitrato(terpyridine)lanthanoid complexes
- Preparation of CoGe2-type NiSn2 at 10 GPa
- Controlled exposure of CuO thin films through corrosion-protecting, ALD-deposited TiO2 overlayers
- Experimental and computational investigations of TiIrB: a new ternary boride with Ti1+x Rh2−x+y Ir3−y B3-type structure
- Synthesis and crystal structure of the lanthanum cyanurate complex La[H2N3C3O3]3 · 8.5 H2O
- Cd additive effect on self-flux growth of Cs-intercalated NbS2 superconducting single crystals
- 14N, 13C, and 119Sn solid-state NMR characterization of tin(II) carbodiimide Sn(NCN)
- Superexchange interactions in AgMF4 (M = Co, Ni, Cu) polymorphs
- Copper(I) iodide-based organic–inorganic hybrid compounds as phosphor materials
- On iodido bismuthates, bismuth complexes and polyiodides with bismuth in the system BiI3/18-crown-6/I2
- Synthesis, crystal structure and selected properties of K2[Ni(dien)2]{[Ni(dien)]2Ta6O19}·11 H2O
- First low-spin carbodiimide, Fe2(NCN)3, predicted from first-principles investigations
- A novel ternary bismuthide, NaMgBi: crystal and electronic structure and electrical properties
- Magnetic properties of 1D spin systems with compositional disorder of three-spin structural units
- Amine-based synthesis of Fe3C nanomaterials: mechanism and impact of synthetic conditions
- Enhanced phosphorescence of Pd(II) and Pt(II) complexes adsorbed onto Laponite for optical sensing of triplet molecular dioxygen in water
- Theoretical investigations of hydrogen absorption in the A15 intermetallics Ti3Sb and Ti3Ir
- Assembly of cobalt-p-sulfonatothiacalix[4]arene frameworks with phosphate, phosphite and phenylphosphonate ligands
- Chiral bis(pyrazolyl)methane copper(I) complexes and their application in nitrene transfer reactions
- UoC-6: a first MOF based on a perfluorinated trimesate ligand
- PbCN2 – an elucidation of its modifications and morphologies
- Flux synthesis, crystal structure and electronic properties of the layered rare earth metal boride silicide Er3Si5–x B. An example of a boron/silicon-ordered structure derived from the AlB2 structure type
Articles in the same Issue
- Frontmatter
- In this issue
- Laudatio/Preface
- Celebrating the 60th birthday of Richard Dronskowski
- Review
- Orbital-selective electronic excitation in phase-change memory materials: a brief review
- Research Articles
- Solving the puzzle of the dielectric nature of tantalum oxynitride perovskites
- d- and s-orbital populations in the d block: unbound atoms in physical vacuum versus chemical elements in condensed matter. A Dronskowski-population analysis
- Single-crystal structures of A 2SiF6 (A = Tl, Rb, Cs), a better structure model for Tl3[SiF6]F, and its novel tetragonal polymorph
- Na2La4(NH2)14·NH3, a lanthanum-rich intermediate in the ammonothermal synthesis of LaN and the effect of ammonia loss on the crystal structure
- Linarite from Cap Garonne
- Salts of octabismuth(2+) polycations crystallized from Lewis-acidic ionic liquids
- High-temperature diffraction experiments and phase diagram of ZrO2 and ZrSiO4
- Thermal conversion of the hydrous aluminosilicate LiAlSiO3(OH)2 into γ-eucryptite
- Crystal structure of mechanochemically prepared Ag2FeGeS4
- Effect of nanostructured Al2O3 on poly(ethylene oxide)-based solid polymer electrolytes
- Sr7N2Sn3: a layered antiperovskite-type nitride stannide containing zigzag chains of Sn4 polyanions
- Exploring the frontier between polar intermetallics and Zintl phases for the examples of the prolific ALnTnTe3-type alkali metal (A) lanthanide (Ln) late transition metal (Tn) tellurides
- Zwitterion coordination to configurationally flexible d 10 cations: synthesis and characterization of tetrakis(betaine) complexes of divalent Zn, Cd, and Hg
- An approach towards the synthesis of lithium and beryllium diphenylphosphinites
- Synthesis, crystal and electronic structure of CaNi2Al8
- Crystal and electronic structure of the new ternary phosphide Ho5Pd19P12
- Synthesis, structure, and magnetic properties of the quaternary oxysulfides Ln 5V3O7S6 (Ln = La, Ce)
- Synthesis, crystal and electronic structure of BaLi2Cd2Ge2
- Structural variations of trinitrato(terpyridine)lanthanoid complexes
- Preparation of CoGe2-type NiSn2 at 10 GPa
- Controlled exposure of CuO thin films through corrosion-protecting, ALD-deposited TiO2 overlayers
- Experimental and computational investigations of TiIrB: a new ternary boride with Ti1+x Rh2−x+y Ir3−y B3-type structure
- Synthesis and crystal structure of the lanthanum cyanurate complex La[H2N3C3O3]3 · 8.5 H2O
- Cd additive effect on self-flux growth of Cs-intercalated NbS2 superconducting single crystals
- 14N, 13C, and 119Sn solid-state NMR characterization of tin(II) carbodiimide Sn(NCN)
- Superexchange interactions in AgMF4 (M = Co, Ni, Cu) polymorphs
- Copper(I) iodide-based organic–inorganic hybrid compounds as phosphor materials
- On iodido bismuthates, bismuth complexes and polyiodides with bismuth in the system BiI3/18-crown-6/I2
- Synthesis, crystal structure and selected properties of K2[Ni(dien)2]{[Ni(dien)]2Ta6O19}·11 H2O
- First low-spin carbodiimide, Fe2(NCN)3, predicted from first-principles investigations
- A novel ternary bismuthide, NaMgBi: crystal and electronic structure and electrical properties
- Magnetic properties of 1D spin systems with compositional disorder of three-spin structural units
- Amine-based synthesis of Fe3C nanomaterials: mechanism and impact of synthetic conditions
- Enhanced phosphorescence of Pd(II) and Pt(II) complexes adsorbed onto Laponite for optical sensing of triplet molecular dioxygen in water
- Theoretical investigations of hydrogen absorption in the A15 intermetallics Ti3Sb and Ti3Ir
- Assembly of cobalt-p-sulfonatothiacalix[4]arene frameworks with phosphate, phosphite and phenylphosphonate ligands
- Chiral bis(pyrazolyl)methane copper(I) complexes and their application in nitrene transfer reactions
- UoC-6: a first MOF based on a perfluorinated trimesate ligand
- PbCN2 – an elucidation of its modifications and morphologies
- Flux synthesis, crystal structure and electronic properties of the layered rare earth metal boride silicide Er3Si5–x B. An example of a boron/silicon-ordered structure derived from the AlB2 structure type