Next Article in Journal
Epimerisation in Peptide Synthesis
Next Article in Special Issue
Experimental and Theoretical Screening of Core Gold Nanoparticles and Their Binding Mechanism to an Anticancer Drug, 2-Thiouracil
Previous Article in Journal
Structural Characterization and Properties of Modified Soybean Meal Protein via Solid-State Fermentation by Bacillus subtilis
Previous Article in Special Issue
A Series of Novel 3D Coordination Polymers Based on the Quinoline-2,4-dicarboxylate Building Block and Lanthanide(III) Ions—Temperature Dependence Investigations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic Structure and Chemical Bonding of the First-, Second-, and Third-Row-Transition-Metal Monoborides: The Formation of Quadruple Bonds in RhB, RuB, and TcB

by
Constantinos Demetriou
1,
Christina Eleftheria Tzeliou
1,
Alexandros Androutsopoulos
1 and
Demeter Tzeli
1,2,*
1
Laboratory of Physical Chemistry, Department of Chemistry, National and Kapodistrian University of Athens, Panepistimiopolis Zografou, 157 84 Athens, Greece
2
Theoretical and Physical Chemistry Institute, National Hellenic Research Foundation, 48 Vassileos Constantinou Ave., 116 35 Athens, Greece
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(24), 8016; https://doi.org/10.3390/molecules28248016
Submission received: 17 November 2023 / Revised: 1 December 2023 / Accepted: 3 December 2023 / Published: 8 December 2023
(This article belongs to the Special Issue Fundamental Aspects of Chemical Bonding)

Abstract

:
Boron presents an important role in chemistry, biology, and materials science. Diatomic transition-metal borides (MBs) are the building blocks of many complexes and materials, and they present unique electronic structures with interesting and peculiar properties and a variety of bonding schemes which are analyzed here. In the first part of this paper, we present a review on the available experimental and theoretical studies on the first-row-transition-metal borides, i.e., ScB, TiB, VB, CrB, MnB, FeB, CoB, NiB, CuB, and ZnB; the second-row-transition-metal borides, i.e., YB, ZrB, NbB, MoB, TcB, RuB, RhB, PdB, AgB, and CdB; and the third-row-transition-metal borides, i.e., LaB, HfB, TaB, WB, ReB, OsB, IrB, PtB, AuB, and HgB. Consequently, in the second part, the second- and third-row MBs are studied via DFT calculations using the B3LYP, TPSSh, and MN15 functionals and, in some cases, via multi-reference methods, MRCISD+Q, in conjunction with the aug-cc-pVQZ-PPM/aug-cc-pVQZB basis sets. Specifically, bond distances, dissociation energies, frequencies, dipole moments, and natural NPA charges are reported. Comparisons between MB molecules along the three rows are presented, and their differences and similarities are analyzed. The bonding of the diatomic borides is also described; it is found that, apart from RhB(X1Σ+), which was just recently found to form quadruple bonds, RuB(X2Δ) and TcB(X3Σ) also form quadruple σ2σ2π2π2 bonds in their X states. Moreover, to fill the gap existing in the current literature, here, we calculate the TcB molecule.

1. Introduction

Boron has an important role in chemistry, biology, and materials science [1]. It is well known that it forms single, double, and triple bonds, but it was only recently found that it can form quadruple bonds in specific diatomic molecules [2,3,4,5]. Additionally, its chemistry is quite interesting to preparative chemists, theoreticians, industrial chemists, and technologists. It is noteworthy that it is the only non-metal in group 13 of the periodic table, and it presents many similarities to its neighbor, carbon, and its diagonal relative, silicon. Hence, like C and Si, it showcases a marked propensity to form covalent, molecular compounds, but it differs greatly from them in having one less valence electron, a situation sometimes referred to as “electron deficiency”. This deficiency plays a key role in its chemistry [1].
Transition-metal borides have received considerable attention since they present common catalytic properties for the hydrogenation of alkenes and alkynes, the reduction of nitrogenous functional groups, and deoxygenation reactions [6]. They are important building blocks in many complexes and materials. Moreover, they possess remarkable physical properties, such as very high conductivity (TiB2) [7]—even superconductivity (MgB2) [8]—as well as super hardness (ReB2) [9]. In solid state, many computational and experimental studies have been carried out; see, for instance, [10,11,12,13,14,15,16,17,18]. Computationally, the DFT methodology is applied to determine the bond lengths, frequencies, and vibrational properties of solids [10,13]; density of state; bond population; charge density maps [14]; relative stability; mechanical, electronic, and magnetic properties [15]; elastic behavior; and elastic anisotropy [16]. Furthermore, ab initio molecular dynamic (AIMD) simulations at finite temperature have also been employed in order to investigate the structural stability of materials, for instance, those of Ʋ2B (Ʋ = Ti, Cr, Nb, Mo, Ta, and W) [18].
It has been reported that the electronic structure and the chemical bonding of diatomic and triatomic molecules are strongly related to their structure, the variety of their morphologies, and the properties of their 2D materials and solid state [19,20]. Therefore, an investigation of the electronic structure and the bonding of the diatomic transition-metal borides, which constitute the simplest building blocks of the compounds or materials in question, would lay the foundation for understanding the very complex solid-metal borides and even their bulk properties. Finally, it should be noted that the diatomic transition-metal borides showcase unique electronic structures presenting interesting and peculiar properties and a variety of bonding schemes.
The present work has two aims. In the first part, we present a review of the experimental and theoretical studies on the first-row-transition-metal borides, i.e., ScB, TiB, VB, CrB, MnB, FeB, CoB, NiB, CuB, and ZnB; the second-row-transition-metal borides, i.e., YB, ZrB, NbB, MoB, TcB, RuB, RhB, PdB, AgB, and CdB; and the third-row-transition-metal borides, i.e., LaB, HfB, TaB, WB, ReB, OsB, IrB, PtB, AuB, and HgB. In the second part, the second- and third-row-transition-metal borides, MBs, are studied via DFT calculations using the B3LYP, TPSSh, and MN15 functionals in conjunction with the aug-cc-pVQZ-PPM/aug-cc-pVQZB basis sets. Additionally, MRCISD(+Q) calculations were carried out to clarify the ground states of the MB molecules when their identity was not known. Bond distances, dissociation energies, frequencies, dipole moments, and Mulliken and natural NPA charges are presented. Comparisons between the MB molecules of all three rows are presented, and their differences and similarities are analyzed. Finally, transition-metal borides forming quadruple bonds are described and analyzed here, and for the first time, we report on the RuB and TcB molecules.

2. Previous Studies on Transition-Metal Monoborides, MBs

2.1. First-Row-Transition-Metal MBs

All previous theoretical and experimental data on the ground states of the first-row-transition-metal borides are summarized in Table 1 [21,22,23,24,25,26,27,28,29,30]. The first study of each first-row-transition-metal boride (MB) was carried out by Wu in 2005, who studied the MB molecules via DFT methodology, i.e., B3LYP/6-311++G(3df) [22]. In 2008, Tzeli and Mavridis, using multireference methods (MRCI), systematically studied the electronic structure and bonding of the ground and some low-lying states, up to twenty-four excited states, of all first-row-transition-metal borides (MBs). They used multireference methods employing correlation-consistent basis sets of quintuple cardinality (cc-pV5Z) [21]. Full potential-energy curves were constructed at the MRCI/cc-pV5Z level for the lowest up to five states, while about twenty states for every MB species were examined at the MRCI/cc-pVQZBANO-4ZM level of theory. At the MRCI/cc-pV5Z level, total energies, dissociation energies, dipole moments, and common spectroscopic parameters of the nine diatomic borides, MBs, M = Sc–Cu, were reported. Ground states of the MBs along with “recommended” bond distances, dissociation energies, and dipole moments were calculated, and boron atoms’ exceptional ability to participate in a variety of bonding schemes was stressed; the bonding in these MB series varied from three half bonds to full triple bonds. Furthermore, the core correlation using the cc-pwCV5Z basis set was calculated for specific molecules as well as the Scalar relativistic effects through the second-order Douglas–Kroll–Hess approximation, i.e., at C-MRCI+DKH2/cc-pwCV5Z-DK. This study [21], using very accurate methodology, calculated bond distances, dissociation energies, dipole moments, and spectroscopic parameters in excellent agreement with experimental studies which were subsequently conducted the following decade; see Table 1 and discussion below.
ScB: In 2008, the electronic structure and bonding of the ground and some low-lying states of ScB were calculated employing MRCI methodology, including the scalar relativistic effects and the correlation of the core electrons [21]. The ground state, X5Σ, was calculated at the C-MRCISD+DKH+Q/cc-pV5Z level of theory, and a value of 2.094 Å was found for the bond length, 3.309 eV was found for the dissociation energy with respect to the adiabatic atomic products, and 603 cm−1 was found for the vibrational frequency. In 2021, Merriles et al. performed resonant two-photon ionization spectroscopy (R2PI) experiments to measure the predissociation threshold in ScB and found a value of D0 (ScB) = 1.72(6) eV with respect to the ground state products [25], while the theoretical dissociation energy with respect to the atomic ground state was 1.74 eV [21], in excellent agreement with the experimental value of 1.72(6) eV [25]. The bonding in the ground state consists of three half bonds [21].
TiB: The ground state of TiB is the X6Δ state. Its bond distance was calculated to be 2.080 Å with a dissociation energy of 2.797 eV and a vibrational frequency of 583 cm−1 using icMRCISD/cc-pV5Z [21]. Recently, Merriles et al. measured the predissociation threshold in TiB to be D0 (TiB) = 1.956(16) eV via R2PI spectroscopy [25]. Finally, in the ground state, three half bonds are formed.
VB: The its ground state of VB, X7Σ+, presents similar bonding behavior to the ground state of the ScB and TiB molecules, i.e., three half bonds are formed [21]. Its bond distance is 2.043 Å, with a dissociation energy, De, of 2.381 eV with respect to the adiabatic products, and 2.143 eV with respect to the atomic ground state products, while the vibrational frequency, ωe, is 589.9 cm−1 [21]. The R2PI predissociation threshold in VB, D0, was measured to be 2.150(16) eV [25], in excellent agreement with the theoretical value.
CrB: The bonding of the CrB ground state, X6Σ+, is a σ2 bond and two half π bonds, i.e., σ2π1π1 [21]. Using the icMRCISD+Q/cc-pV5Z methodology, the Cr-B distance was calculated to be 2.183 Å with a binding energy of 1.353 eV, a rather small dissociation energy, while the ωe value was 405 cm−1. The B3LYP bond distance, 2.187 Å [22], is in very good agreement with the corresponding value obtained using the icMRCISD+Q/cc-pV5Z methodology [21].
MnB: The relative ordering of twenty-six states of MnB has been calculated [21]. The ground state is the X5Π state, with a value of 2.183 Å for the bond length, 0.854 eV for the dissociation energy, and 391.9 cm−1 for the vibrational frequency. The bonding is σ2π1π1; however, the dissociation energy is very small. On the contrary, the first excited state, A5Σ, which lies 0.09 eV above the X state, has three bonds, σ2π2π2, a short bond length of 1.832 Å, and a significant larger dissociation energy of 3.131 eV than the X state obtained using the icMRCISD+Q/cc-pV5Z methodology [21]. Note that in the A5Σ state, Mn is excited at its 5D atomic state.
FeB: In 2005, the ground state, 4Σ, of the FeB molecule was calculated at the DFT(B3LYP)/6-311++G(3df) level [22]. Then, in 2008, nineteen electronic states were calculated using MRCI, while the electronic structure and chemical bonding of the ground and the first excited states were examined [21]. At the icMRCISD+Q/cc-pV5Z level of theory, the X4Σ ground state values are re = 1.743 Å and De = 2.346 eV. In 2019, Merriles et al. measured the predissociation threshold in FeB via R2PI spectroscopy and found a value of D0 = 2.43(2) eV [26]. This work included the first experimental measurement of the BDE of FeB.
CoB: The ground state of CoB, X3Δ, was calculated in 2005 via the DFT methodology [26], and in 2008, it was calculated using icMRCISD+Q/cc-pV5Z [21]. Eighteen states were calculated. In the ground state, a triple bond is formed with a bond distance of 1.700 Å and a binding energy of 2.849 eV. In 2011, Ng et al. observed and analyzed the electronic transition spectrum of CoB in the visible region between 495 and 560 nm using laser-induced fluorescence spectroscopy [27]. The ground state of CoB was identified to be X3Δ3 with re = 1.705 Å. The ground state’s identity was reconfirmed to be X3Δ3 by Dore et al. [31]. In 2019, Merriles et al. performed resonant two-photon ionization spectroscopy experiments to measure the predissociation threshold of CoB, obtaining a value of D0 = 2.954(3) eV [26].
NiB: The X2Σ+ state of NiB is the ground state, presenting a triple bond. It was calculated via DFT methodology in 2005 [22] and the MRCI method in 2008 [21]. Via MRCI, twenty states were studied. At the icMRCISD+Q/cc-pV5Z level, its bond length was calculated to be 1.681 Å with a binding energy of 3.434 eV and ωe = 803 cm−1. Also in 2008, Balfour et al. characterized NiB spectroscopically using laser-induced fluorescence spectroscopy [28]. The ground state of NiB was identified to be X2Σ+ with an electronic configuration of 1σ22σ21π41δ43σ1, re = 1.698 Å, ωe = 778.0 cm−1, and ωexe = 4.90 cm−1 [28]. In 2010, Zhen et al. investigated NiB using LIF spectroscopy to resolve the rotational structure of a band belonging to a newly discovered band system with a 2Π3/2 upper state [32]. In 2015, Goudreau et al. [33] investigated the 0-0, 2-0, and 3-0 bands of NiB belonging to the 2Π3/2 ← X2Σ+ band system assigned by Zhen et al. [32] at high resolution. The fine structure splitting in each state was determined for the first time, confirming the assignment of the ground state as 2Σ+ with an electronic configuration of 1σ22σ21π41δ43σ1. Finally, in 2019, Merriles et al., via R2PI spectroscopy, measured the predissociation threshold in NiB to be D0 = 3.431(4) eV [26].
CuB: In 1997, Barysz and Urban investigated the spectroscopic constants and dipole moment curves of the ground states, Χ1Σ+, of the coinage metal diatomic molecules with boron, i.e., BCu, using high-level-correlated methods combined with quasi-relativistic Douglas–Kroll (no-pair) spin-averaged approximation [29]. At the CCSD(T)/[9s7p3d2f/Cu5s3p2d/B] computational level, the values re = 1.909 Å, De = 1.522 eV, and ωe = 555.0 cm−1 were found. In 2005, Wu also studied the X state via DFT [22], while in 2008, Tzeli and Mavridis investigated nineteen states using the MRCI+Q methodology [21]. At the icMRCISD+Q/cc-pV5Z level of theory, they found values of re = 1.922 Å, De = 2.129 eV, and ωe = 553 cm−1. The De value at the icMRCISD/cc-pV5Z level was significantly larger than the corresponding values at the CCSD(T)/[9s7p3d2f/Cu5s3p2d/B] level due to its significantly larger basis set. In 2023, Merriles and Morse studied CuB experimentally for the first time via R2PI spectroscopy and obtained the first BDE measurement for this molecule. They found that CuB remains bound at energies that far surpass its bond dissociation energy (BDE), and bonds break only when excited at or above an excited sharp predissociation threshold (SAL). Nevertheless, a BDE value of D0 = 2.26(15) eV was derived [30], which was in very good agreement with the calculated value of 2.129 eV using icMRCISD+Q/cc-pV5Z [21].
ZnB: Only one theoretical study was found for ZnB. Its ground state, Χ2Π, was calculated via the DFT methodology, B3LYP/6-311++G(3df) [22], obtaining a value of re = 2.274 Å with a very small binding energy of 0.370 eV. Here, we calculated the X state at the B3LYP, TPSSh, and MN15/aug-cc-pVQZ(-PP) levels of theory. Both B3LYP and TPSSh provided the same De values, i.e., 0.373 eV, while TPSSh overestimated it. Here, we found that the B3LYP/aug-cc-pVQZ(-PP) methodology is in very good agreement with the available experimental data on MBs compared with the other functionals. Thus, we consider it to be our best DFT methodology. The bond distance was found to be 2.258 Å and the dipole moment was found at 1.65 D.

2.2. Second-Row-Transition-Metal MBs

All previous theoretical and experimental data on the ground states of the diatomic metal borides including the second-row transition metals are summarized in Table 2. Below, they are discussed in detail. There is no experimental or theoretical study on TcB except a D0 value for the 5Σ, obtained via DFT(B97-1). Note that Tc is a synthetic element, and all its isotopes are radioactive. In this paper, we fill this gap by studying the TcB molecule.
YB: The first report on YB was in 1969 by K. A. Gingerich [24], who estimated the dissociation energy of the molecule to be 2.99 eV using the Pauling method of electronegativities [24,25]. In 2009, Kharat et al. calculated the ground spin states of the second-row (4d) transition metals (except for Tc) and their cationic and anionic counterparts at the DFT(B3LYP)/LANL2DZ level. They calculated the bond distances, re; binding energies, De; electron affinities (EA); ionization potentials (IP); vibrational frequencies, ωe; and dipole moments, μ [34]. For the diatomic YB, a quintet (S = 2) ground spin state was established. The report lacks details about its electronic configuration, and as such, its ground spin state electronic symmetry is not included. The associated bond distance was found to be 2.254 Å, the binding energy was 2.17 eV, the EA was 0.69 eV, the IP was 6.16 eV, ωe = 582.4 cm−1, and μ = 4.65 D. The most recent investigation into the electronic structure of the YB dimer was made in 2021 by Merriles et al. [25]. They measured the predissociation thresholds of several early-transition-metal boride diatomics using resonant two-photon ionization (R2PI) spectroscopy. For the YB molecule, a D0 value of 2.057(3) eV was obtained. Additionally, they provided an insight into the chemical bonding and electronic structures of those same species by performing quantum chemical calculations using the DFT (B97-1) methodology. Computational results showed excellent agreement with the measurements for YB, and its ground state was computed to be the wavefunction X5Σ (with a dominating 1σ22σ13σ11π2 configuration determinant), resulting in a bond distance re of 2.306 Å, a dissociation energy D0 of 1.96 eV, and ωe = 517 cm−1.
ZrB: This molecule has only been studied together with other similar species, once in 2009 by Kharat et al. [34] and in 2021 by Merriles et al. [25]. In the first study, a sextet (S = 5/2) ground spin state was determined, with re = 2.189 Å, De =3.92 eV, EA = 0.48 eV, IP = 7.03 eV, ωe = 582.4 cm−1, and μ = 3.48 D. In the latter, the predissociation threshold revealed a D0 value of 2.573(5) eV, and the ground electronic spin state corresponded to the X6Δ symmetry wavefunction (with a dominating 1σ22σ13σ11π21δ1 configuration determinant), resulting in a 2.159 Å bond distance, a 2.61 eV dissociation energy, D0, and an ωe value of 610 cm−1.
NbB: Similarly to the previous molecule, there are two studies on NbB [25,34]. In the first study [34], a triplet (S = 1) ground spin state was found, with a bond distance of 1.996 Å, a binding energy of 3.40 eV, EA = 1.05 eV, IP = 7.03 eV, ωe = 662.7 cm−1, and μ = 3.84 D. In the latter [25], the predissociation threshold revealed a D0 value of 2.989(12) eV, and the computed ground spin state corresponded to the superposition X5Π/5Φ (with a 1σ22σ13σ11π31δ1 configuration determinant), due to the calculations using real forms of the 1π and 1δ orbitals, providing a bond distance of 1.988 Å, a dissociation energy of 3.07 eV, and a vibrational frequency of 698 cm−1. Here, we carried out MRCISD(+Q)/aug-cc-PVQZ(-PP) calculations, and we clarified that the X state is the X5Π state, while the A5Φ is located 0.084(0.093) eV above the X state; see discussion below.
MoB: In the first report of MoB [34], it was claimed that the ground state is doublet (S = 1/2), with a 1.817 Å bond distance, a 6.40 eV binding energy, EA = 0.21 eV, IP = 8.68 eV, ωe = 826 cm−1, and μ = 4.05 D. In 2011, Borin and Gobbo [35] investigated the electronic structure of the Χ state and the low-lying electronic states of MoB and its cationic MoB+ counterpart by employing quantum computational CASSCF protocols on the CASPT2+DKH/4ζ-8s7p5d3f2g-ANO-RCCMo/4ζ-5s4p3d2fB level. MoB’s ground spin state was computed to be the wavefunction X6Π (with a dominant 1σ22σ13σ11δ21π3 configuration determinant) with a bond distance, re, of 1.968 Å, a vibrational frequency, ωe, of 664 cm−1, and a dipole moment, μ, of 2.7 D. The binding energy was also determined to be 2.18 eV.
TcB: TcB is the least studied molecule. Its only study resulted in a D0 value calculated to be 3.31 eV [25] via the B97-1/(aug)-cc-pVTZ-(PP) methodology for the X5Σ state. Tc is the lightest element, and all its isotopes are radioactive. In this paper, we fill this research gap, and we study three states of the TcB molecule. The main spectroscopic data are provided; see discussion below and tables of Section 3 below. Here, we found that the ground state is an X3Σ state which presents a quadruple bond; see discussion below.
RuB: This molecule was studied experimentally via mass spectrometry for the first time by Auwera-Mahieu et al. in 1970 [36]. Via the Knudsen effusion method, they determined its dissociation energy, D0, to be 4.59(22) eV. The ωe value of 915 cm−1 was estimated from the values of the corresponding carbides using the D0(A)/D0(B) = μAωA2/μBωB2 equation, where μ is the reduced mass. The internuclear distance was estimated from the values of the carbides using the formula rMB = rMC + ½ (rB₂rC₂), resulting in a value of 1.75 Å. Finally, they proposed that the X state is a 2Σ state. It took nearly four decades for this molecule to be inspected again, and in the 2009 work of Kharat et al. [34], a doublet (S = ½) ground spin state was reported for the diatomic RuB, with a bond distance of 1.761 Å, a binding energy of 6.48 eV, an EA of 0.35 eV, an IP of 9.06 eV, an ωe of 910.8 cm−1, and a dipole moment of 3.49 D. In 2012, Wang et al. [37] studied the laser-induced fluorescence spectrum of RuB in the visible region between 500 nm and 575 nm. They determined that the ground state symmetry is X2Δ, consistent with an electronic configuration obtained using molecular orbital energy level diagrams, while the bond length, r0, is 1.7099 Å. In 2019, Merriles et al. [26] used R2PI spectroscopy and accurately assigned the bond dissociation energies of the diatomic late-transition-metal monoborides from the measurement of a predissociation threshold. The measured predissociation threshold resulted in D0 = 4.815(3) eV. Continuing their work, in 2022, Merriles et al. [45] investigated the ionization energies and the cationic dissociation energies of the diatomic second- and third-row-late-transition-metal borides they had previously examined. Resonant two-photon ionization spectroscopy was employed, and the ionization threshold of RuB was measured to be 7.879(9) eV. Regarding the ground state, it was characterized as X2Σ via Knudsen effusion [36] and 2Δ5/2 via LIF spectroscopy. Here, we clarify that the X2Δ state is the ground state, and that it presents a quadruple bonding; see discussion in the next section.
RhB: In 1970, Auwera-Mahieu et al. [36], via the Knusden effusion method, yielded a value of 4.89(22) eV for the D0, dissociation energy, of RhB. The vibrational frequency, the bond distance, and the ground state were proposed to be 915 cm−1, 1.75 Å, and 1Σ, respectively. In 2006, Chowdhury and Balfour [38] measured the gas phase electronic spectrum of RhB in the visible region; it was elucidated that the ground electronic state is of X1Σ+ symmetry, with an internuclear distance of 1.691(2) Å. The following year, in 2007, Gobbo and Borin [39] studied the low-lying 1Σ+ states of RhB at the CASSCF/MS-CASPT2/4ζ-ANO-RCCRh/14s9p4d3fB level to answer some questions raised by the previous experiments. In agreement with the experiment, their results indicated that the ground electronic state corresponded to the X1Σ+ wavefunction (with a dominant 1σ22σ21π41δ4 configuration determinant) with an internuclear distance, r0, of 1.698 Å. In the same year, another study by Chowdhury and Balfour [40] resumed their previous spectroscopic studies, with a clear focus on the intrinsic details of the emerging bands. In 2008, A.C. Borin and J.P. Gobbo [41], in order to gain further insight into the structural and spectroscopic properties of RhB, investigated its first two atomic dissociation channels. The first regarded the adiabatic coupling of the two atoms in their ground atomic states, B(2s22p;2P) and Rh(4d8(3F)5s;4F), while in the second, the rhodium atom participated in its first excited atomic electronic state, Rh(4d9;2D), at the CASSCF/MS-CASPT2/4ζ-ANO-RCCRh/14s9p4d3fB level of theory. Results showcased that X1Σ+ correlates with the second atomic dissociation limit. The researchers predicted a 5.6 eV dissociation energy, D0; a 924 cm−1 vibrational frequency, ωe; and a dipole moment of 4.54 D. The Mulliken population analysis yielded a charge of +0.35e on Rh. In the 2009 work of Kharat et al. [34], a singlet (S = 0) ground state was reported for RhB, with a 1.745 Å bond distance, a 6.48 eV binding energy, EA = 0.85 eV, IP = 8.19 eV, and μ = 2.84 D. Later, in the 2019 work by Merriles et al. [26], a predissociation threshold of D0 = 5.252(3) eV was measured with the use of R2PI spectroscopy. In 2020, two works concerning the bonding structure of RhB were published, where the formation of a quadruple bond was reported. The first study conducted by Cheung et al. [2] explored the bonding nature of RhB(BO) and RhB species. With the use of PES, an electronic fingerprint was obtained, and the electron affinity of the dimer was measured experimentally to be 0.961 eV. In all computational levels, ADF, DFT, CCSD(T), it was found that the electronic ground state corresponds to an X1Σ+ wavefunction (with a dominant 1σ21π42σ21δ4 configuration determinant), resulting in a quadruple bond consisting of two π bonds formed between the Rh 4dxz/4dyz and B 2px/2py orbitals and two σ bonds between the Rh 4dz² and B 2s/2pz orbitals, followed by internuclear distances ranging from 1.685 Å to 1.689 Å. At the ADF/PBE/TZ2P level, it was also possible to obtain a 5.27 eV value for the dissociation energy. The second study, carried out by Tzeli and Karapetsas [4], investigated the bond occurring inside isoelectronic species between transition metals and the main group elements TcN, RuC, RhB and PdBe. For the RhB molecule, at various levels of theory, i.e., MRCISD, MRCISD+Q, and RCCSD(T)/aug-cc-pV5Z-PPRh aug-cc-pV5ZB, the common spectroscopic constants were computed, presenting great agreement among themselves, as well as with the experimental values. Specifically, values of re =1.6872 Å, De = 5.490 eV, and ωe = 942.1 cm−1 were found, along with anharmonic corrections of ωexe =3.78 cm−1 and μ = 2.965 D [4]. It was deduced that the ground electronic state (X1Σ+) has a dominant 1σ22σ21π41δ4 configuration determinant and correlates adiabatically to the atomic electronic spin states B(X2P;2s22p) + Rh(a2D;4d9), resulting in a four-fold quadruple bond. Tzeli and Karapetsas [4] found that, except for the ground state of RhB, its two lowest excited states, i.e., a3Δ and A1Δ, also present quadruple bonds [4]. Additionally, in the ground and the first excited states of the RhB anion, Χ2Σ+ and A2Δ quadruple bonds are also formed [5]. In 2021, Schoendorff et al. [42] also studied the bonding in RhB both qualitatively and quantitively at a MCSCF-IOTC/DKH-TZP-2012Rh/Sapporo-TZP-2012B level of theory and reached the same results as those in previous works. They concluded that the ground electronic spin state corresponded to the X1Σ+ symmetry wavefunction (with a dominant 1σ22σ21π41δ4 configuration determinant), with an equilibrium bond length of 1.701 Å and a binding energy of 5.165 eV. Finally, in 2022, Merriles and Morse [45] measured the ionization potential and obtained a value of 8.234(10) eV.
PdB: Via the Knusden effusion method, the dissociation energy, D0, of PdB was measured to be 4.89(22) eV, with ω = 650 cm−1 and a bond distance of 2.00 Å, while the proposed ground state was 2Σ [36]. In 1992, Knight Jr. et al. [43], via electron spin resonance (ESR) spectroscopy, revealed that the ground electronic state of the dimer is X2Σ. Unrestricted Hartree–Fock (UHF) calculations were also carried out and a 1.608 Å internuclear distance was deduced. In 2009, a DFT study [34] predicted a doublet (S = ½) ground state with a bond distance of 1.856 Å and a binding energy of 3.33 eV. The most recent study on PdB was published in 2012 by Ng et al. [44], marking its first electronic spectroscopic investigation using laser-induced fluorescence spectroscopy in the visible region between 465 and 520 nm. An X2Σ+ ground state was revealed, with r0 = 1.7278 Å. Moreover, a molecular-orbital-energy-level diagram was designed to understand the observed ground state, and the proposed configuration was determined to be 1σ22σ21π41δ43σ1.
AgB: In 1997, Barysz and Urban [29] investigated the spectroscopic constants of AgB at many levels of theory and the plotted dipole moment curves of AgB using high-level-correlated methods combined with quasi-relativistic Douglas–Kroll (no-pair) spin-averaged approximation. The obtained ground state was X1Σ+ in all of them. At the DK-CCSD(T)-20/NpPolMe level of theory, a bond length of 2.115 Å and a dissociation energy of 0.910 eV were obtained, while the corresponding values at the DK-CASPT2/NpPolMe level were 2.098 Å and 1.684 eV. It was highly advised that both relativistic as well as correlation effects be carefully considered to obtain accurate results. Note that due to the relativistic shrinkage of s valence shell electrons, stronger bonds are formed which would, otherwise, not be described successfully. Finally, Kharat et al. [34] predicted via DFT a singlet (S = 0) ground state with a 2.187 Å bond distance and a 1.60 eV binding energy.
CdB: There is only one DFT study in the literature. Kharat et al. [34] reported a doublet (S = ½) ground state with a 2.668 Å bond distance, a 0.22 eV binding energy, an EA of 0.09 eV, an IP of 7.05 eV, ωe = 198.3 cm−1, and μ = 1.67 D.

2.3. Third-Row-Transition-Metal MBs

Table 3 summarizes the previous experimental and theoretical data obtained for LaB, HfB, TaB, WB, ReB, OsB, IrB, PtB, AuB, and HgB. The calculations are mainly DFT apart from those of HfB [46], PtB [47], and AuB [29,47], for which CCSD(T), CASPT2, and MRCI calculations were carried out.
LaB: In 1969, K. A. Gingerich [24] estimated the binding energy at 3.51 eV. In 2010, Kalamse et al. [48] studied the 5d-metal mononitrides and monoborides using DFT methodology with the B3LYP functional set and both LANL2DZ and SDD basis sets. The lowest electronic spin states at these two DFT levels of theory, as well as EA, IP, binding energies, and electronic configurations of the MBs, were discussed, while the orbitals involved in bond formation were identified. At the B3LYP/LANL2DZ level, it was deduced that the most stable state for LaB is X5Σ, with a bond length of 2.435 Å, a binding energy of 2.49 eV, μ = 4.29 D, EA = 1.01 eV, and IP = 5.61 eV. On the other hand, for the B3LYP/SDD level, the most stable state was found to be X3Π, with re = 2.336 Å and De = 2.49 eV. In 2018, Elkahwagy et al. [49] studied LaB and its anionic LaB counterpart with the diffusion Monte Carlo method in combination with three different trial functions to calculate the potential energy curves for the lowest electronic states of those species, along with some spectroscopic constants of neutral LaB. Irrespectively of the functional used, it was found that the quintet state of LaB is the ground state, while the triplet state is higher in energy, elucidating the mystery that the previous work arose. The 2021 study by Merriles et al. [25] also suggested that the ground spin state corresponded to an X5Σ symmetry wavefunction (with a dominant 2σ13σ11π2 configuration determinant) with a dissociation energy, D0, of 2.54 eV; a bond length, re, of 2.372 Å; and a vibrational frequency, ωe, of 521 cm−1. The experimental part of their study yielded a 2.086(18) eV value for the dissociation energy from the predissociation threshold. Here, both the lowest triplet and quintet states were calculated, and we found that the ground one is the X5Σ state.
Table 3. Previous theoretical and experimental data on the ground states of the 3rd-row-transition-metal boride molecules, MBs (M = La, Hf, Ta, W, Re, Os, Ir, Pt, Au, and Hg): bond lengths re (Å), dissociation energies De (eV) and/or D0 (eV), electron affinities EA (eV), ionization potentials IP (eV), vibrational frequencies ωe (cm−1), anharmonic corrections ωexe (cm−1), and dipole moments μ (μFF = δE/δε) (Debye).
Table 3. Previous theoretical and experimental data on the ground states of the 3rd-row-transition-metal boride molecules, MBs (M = La, Hf, Ta, W, Re, Os, Ir, Pt, Au, and Hg): bond lengths re (Å), dissociation energies De (eV) and/or D0 (eV), electron affinities EA (eV), ionization potentials IP (eV), vibrational frequencies ωe (cm−1), anharmonic corrections ωexe (cm−1), and dipole moments μ (μFF = δE/δε) (Debye).
MBMethodologyRef.StatereDe aD0ωeωexeμ (μFF)
LaBDFT: B3LYP/LANL2DZ[48]X5Σ2.4352.49 511 4.29
DFT/B3LYP/SDD[48]X3Π2.3362.10 496 5.02
DMC(B3LYP)/CRENBS-ECPLaBurkatzki-PPB[49]S = 22.1503.37 4.22
DMC(B3PW91)/CRENBS-ECPLa Burk.-PPB[49]S = 22.1453.81 4.19
R2PI spectroscopy[25] 2.086(18)
DFT: UB97-1/AVTZ-PPLa VTZB[25]X5Σ2.3722.572.54521
Pauling method[24] 3.51
HfBDFT: B3LYP/LANL2DZ[48]X4Σ2.1572.70 613 2.44
DFT/B3LYP/SDD[48]X4Σ2.1952.64 580 2.68
R2PI spectroscopy[25] 2.593(3)
DFT: UB97-1/AVTZ-PPLa VTZB[25]X4Σ2.1282.642.60584
c-CCSD(T)/wCV5Z-PPHf AV5ZB[46]X4Σ2.1442.8412.8406072.8
MRCI/c-CCSD(T)/wCVQZ-PPHf AVQZB[46]X4Σ2.174 6103.0
BP86/def2-QZVP[46] 3.182
BLYP/def2-QZVP[46] 2.803
BPE/def2-QZVP[46] 3.311
MN15-L/def2-QZVP[46] 3.071
LRC-ωPBEh/def2-QZVP[46] 2.174
DSD-PBEB95-D3BJ/def2-QZVP[46] 1.923
TaBDFT: B3LYP/LANL2DZ[48]X3Σ+2.0012.49 721 2.68
DFT/B3LYP/SDD[48]X5Δ2.1842.48 555 1.44
R2PI spectroscopy[25] 2.700(3)
DFT: UB97-1/AVTZ-PPTa VTZB[25]X5Δ2.0853.002.95754
WBDFT: B3LYP/LANL2DZ[48]X6Σ2.1612.74 526 2.61
DFT/B3LYP/SDD[48]X6Σ1.9902.88 725 2.67
R2PI spectroscopy[25] 2.730(4)
DFT: UB97-1/AVTZ-PPW VTZB [25]X6Σ+1.8912.942.89730
ReBDFT: B3LYP/LANL2DZ[48]X3Σ1.8422.77 867 2.99
DFT: B3LYP/SDD[48]X5Σ1.8754.18 853 2.29
OsBR2PI spectroscopy[26]GS 4.378(3)
B3LYP/aug-cc-pVQZ-PP[45]X4Σ1.770 938 2.24
B3LYP/LANL2DZ[48]Χ4Δ1.8133.99 955 2.17
IrBLIF spectroscopy[50]Χ3Δ31.7675
B2PLYP/AVQZ(-PP)M[47]Χ3Δ1.763
CCSD(T)/AVQZ(-PP)[47]Χ3Δ 5.085
R2PI spectroscopy[26]GS 4.928(10)
Mass spectrometry[36]GS 5.27(18)
B3LYP/LANL2DZ[48]Χ3Δ1.8064.86 936 1.72
PtBR2PI spectroscopy[26]GS 5.235(3)
Mass spectrometry[51]GS 4.91(17)
LIF spectroscopy[52]Χ2Σ+1.741 903.6
B3LYP/LANL2DZ[48]Χ2Σ+1.8095.43 906 1.18
B2PLYP/AVQZ(-PP)[47]Χ2Σ+1.755
CCSD(T)/AVQZ(-PP)[47]Χ2Σ+ 5.668
AuBMass spectrometry[36]GS 3.50(16)
Knudsen cell experiment[53]GS 3.773
B3LYP/LANL2DZ[48]Χ1Σ+1.9972.96 559 0.68
B2PLYP/AVQZ(-PP)[47]Χ1Σ+1.906 710
CCSD(T)/AVQZ(-PP)[47]Χ1Σ+ 3.734
Nonrelativistic CASPT2/PolMe[29]Χ1Σ+2.2561.261 3623.06
No-pair DK CASPT2/NpPolMe[29]Χ1Σ+1.9313.519 6764.76
No-pair DK CCSD(T)-20/NpPolMe[29]Χ1Σ+1.9602.709 6634.01
R2PI spectroscopy[30]Χ1Σ+ 3.724(3)
HgBB3LYP/LANL2DZ[48]Χ2Σ+4.3810.002 19 0.27
a Dissociation energy with respect to the atomic ground state products.
HfB: The B3LYP/SDD methodology [48] was used to predict X4Σ as the ground state of HfB with a bond length of 2.195 Å, μ = 2.60 D, EA = 1.05 eV, IP = 5.01 eV, and a binding energy of 2.64 eV. The researchers found that, in contrast to the rest of the 5d-metal monoborides, it is the 5d orbital of the metal that loses an electron in the case of HfB. In the 2021 study by Merriles et al. [25], the authors measured the predissociation threshold, i.e., D0(HfB) = 2.593(3) eV, and the calculated ground spin state corresponded to the X4Σ symmetry wavefunction (with a dominant 2σ13σ21π2 configuration determinant) with a bond distance of 2.128 Å, a dissociation energy of 2.60 eV, and a vibrational frequency of 584 cm−1 at UB97-1/AVTZ-PPLaVTZB. In 2022, Ariyarathna et al. [46] conducted a comparative study on the outcomes of several computational methods on the basis of high-level, multi-reference configuration interaction theory and coupled cluster quantum chemical calculations, with large quadruple-ζ and quintuple-ζ quality-correlation-consistent basis sets, as well as DFT, with numerous exchange-correlation functionals that span multiple rungs of “Jacob’s ladder”, through the inspection of HfO and HfB, to investigate their performance on species containing third-row transition metals. Ab initio studies of HfB showed, unanimously, that X4Σ is the ground electronic state, owing to a 1σ22σ13σ21π2 configuration dominant determinant with a bond length of 2.144 Å, dissociation energy at 2.841 eV, and ωe = 607 cm−1 at the c-CCSD(T)/wCV5Z-PPHf AV5ZB level. DFT calculations were performed on the closed-shell single-reference ground state to evaluate the errors in the predictions from distinct density functionals. The resulting trends were discussed, plotted, and compared.
TaB: DFT calculations predict different states to be the ground state of TaB depending on the methodology. B3LYP/LANL2DZ predicted X3Σ+ to be the ground state, with a 2π45σ16σ1 electronic configuration, re = 2.001 Å, and De = 2.49 eV. However, at the B3LYP/SDD level, the calculated ground state was X5Δ state, with re = 2.184 Å, μ = 1.44 D, ΕA = 1.43 eV, ΙP = 7.35 eV, and De = 2.48 eV [48]. In 2021, Merriles et al. [25] determined the D0 value to be 2.700(3) eV, and they found that at the UB97-1/AVTZ-PPTaVTZB level, the ground state corresponded to X5Δ (with a dominant 2σ13σ21π21δ1 configuration determinant) with re = 2.085 Å, De = 2.95 eV, and ωe = 754 cm−1.
WB: In WB, as in TaB, DFT calculations predict different states to be the ground state, depending on the methodology. Kalamse et al. [48], at both the B3LYP/LANL2DZ and SDD levels, found that the ground state is X6Σ, where re = 1.990 Å, μ = 2.67 D, and De = 2.88 eV at the B3LYP/SDD level [50]. In 2021, Merriles et al. [25] measured D0 = 2.730(4) eV and calculated the ground state to be X6Σ+, with a dominant 2σ13σ21π21δ2 configuration determinant, re = 1.891 Å, De = 2.89 eV, and ωe = 730 cm−1, using the UB97-1/AVTZ-PPWVTZB methodology [25]. Here, we found that the ground state is the X6Π state, while the X6Σ+ state lies 0.137 eV above the X state; see tables of the Section 3 and discussion below.
ReB: There is only a single theoretical study on the ReB molecule at a DFT level; however, different basis sets predict different states to be the ground state. B3LYP/LANL2DZ predicted the ground state to be X3Σ, with re = 1.842 Å and De = 2.77 eV, while B3LYP/SDD predicted X5Σ to be the ground state, with re = 1.875 Å and De = 4.18 eV [48]. Here, we found that the ground state is X5Σ, while the a3Σ state lies 0.099 eV above the X state; see tables of Section 3 and discussion below.
OsB: Kalamse et al., via B3LYP/LANL2DZ and B3LYP/SDD, predicted that the ground state of OsB is the Χ4Δ state. It corresponds with a π4δ35σ16σ1 electronic configuration, with re = 1.813 Å, ωe = 955 cm−1, μ = 2.17 D, and De = 3.99 eV at the B3LYP/LANL2DZ level [48]. In 2019, Merriles et al. performed R2PI spectroscopy experiments to measure the predissociation threshold, and it was found to be D0 = 4.378(3) eV [26]. Furthermore, they investigated the electronic ground state at the B3LYP/Def2TZVPP [26] and B3LYP/aug-cc-pVQZ-PP levels [45], and they found that it is a triply bonded 1σ22σ21π41δ23σ1, at the X4Σ state, which contradicts with the prediction of Kalamse et al. [48], with a bond distance of 1.770 Å. In 2022, Merriles et al. measured, for the first time, the ionization energy of OsB using resonant two-photon ionization spectroscopy, obtaining a value of IE (OsB) = 7.955(9) eV [45].
IrB: In 1969, Auwera-Mahieu et al. determined the dissociation energy of IrB to be 5.27(18) eV via a mass spectrometric study at high temperatures [36]. In 2005, Ye et al. investigated IrB using LIF spectroscopy [50]. In this study, the ground state of IrB was found to be X3Δ3, with an electronic configuration of 1σ21π41δ32σ1 and re = 1.7675 Å. In 2010, using the B3LYP/LANL2DZ methodology, it was also found that the most stable state for IrB is Χ3Δ raised from a π4δ35σ26σ1 electronic configuration, with re = 1.806 Å, μ = 1.72 D, ΕA = 5.06 eV, ΙP = 4.66 eV, and De = 4.86 eV [48]. In 2011, Pang et al. investigated the electronic transition spectrum of IrB in the spectral region between 400 and 545 nm using LIF spectroscopy [54]. Its ground state was also identified to be X3Δ3 with an electronic configuration of 1σ22σ2π4δ33σ1. In 2019, Merriles et al., using R2PI spectroscopy, measured the predissociation threshold of IrB, obtaining a value of D0 = 4.928(10) eV [26]. In 2020, Wang et al. investigated the nature of the chemical bonding in IrB, employing high-resolution photoelectron imaging and theoretical B2PLYP and CCSD(T)/aug-cc-pVQZ(-PP) calculations. They calculated a bond distance of 1.763 Å using B2PLYP and a dissociation energy of 5.085 eV at the CCSD(T) level of theory, in good agreement with the experimental values [47]. In 2022, Merriles and Morse measured, for the first time, the ionization energy of IrB using R2PI spectroscopy, obtaining IE = 8.301(15) eV [45].
PtB: In 1968, via Knudsen effusion mass spectrometry, McIntyre et al. measured the dissociation energy of gaseous PtB to be 4.91(17) eV [51]. Via the B3LYP/LANL2DZ (B3LYP/SDD) levels of theory, the ground state was assigned to be Χ2Σ+, derived from a π45σ2δ46σ1 electronic configuration, with re = 1.809(1.815) Å and De = 5.43(4.86) eV [48]. In 2012, Ng et al. investigated the optical spectrum of PtB in the visible region between 455 and 520 nm using LIF spectroscopy [52]. The ground state of PtB was identified to be Χ2Σ+, determined using an electronic configuration of 1σ22σ21π41δ43σ1, with re = 1.741 Å and ωe = 903.60 cm−1. In 2019, Merriles et al. performed R2PI spectroscopy measurements, and they found D0 = 5.235(3) eV [26]. In 2020, Wang et al. investigated the nature of the chemical bonding in PtB, employing high-resolution photoelectron imaging and theoretical calculations using the B2PLYP and CCSD(T)/aug-cc-pVQZ(-PP) methodologies [47]. They calculated a bond distance of 1.755 Å using B2PLYP and a dissociation energy of 5.668 eV at the CCSD(T) level of theory [47]. In 2022, Merriles and Morse measured, for the first time, the ionization energy of PtB using R2PI spectroscopy, obtaining a value of IE = 8.524(10) eV [45].
AuB: In 1969, via mass spectrometry at high temperatures, Auwera-Mahieu et al. measured the dissociation energy of AuB to be 3.50(16) eV [36]. In 1971, Gingerich investigated AuB using a combination of Knudsen effusion and mass spectroscopic techniques. The reaction enthalpies determined by the second and third law method yielded D0 = 3.77 eV [53]. In 1997, Barysz and Urban investigated the spectroscopic constants and dipole moment curves of the AuB ground state, Χ1Σ+, using high-level-correlated methods combined with quasi-relativistic Douglas–Kroll (no-pair) spin-averaged approximation [29]. At the CCSD(T)/[13s11p7d4f/Au5s3p2d/B] computational level, they found values of re = 1.960 Å, De = 2.709 eV, ωe = 663.0 cm−1, and ωexe = 4.01 cm−1. In 2010, via B3LYP/LANL2DZ and SSD, values of De = 2.96 eV and 3.04 eV, respectively, were calculated [48]. In 2020, Wang et al. investigated the nature of the chemical bonding in AuB, employing high-resolution photoelectron imaging and theoretical calculations [47]. They calculated re = 1.906 Å using B2PLYP)/aug-cc-pVQZ(-PP) and a dissociation energy of 3.734 eV at the CCSD(T) level of theory [47]. This dissociation energy was in excellent agreement with the experimental value of 3.724(3) eV measured by Merriles and Morse in 2023 [30], who examined the AuB molecule experimentally using R2PI spectroscopy. They found that it remains bound at energies that far surpass its bond dissociation energy (BDE), and bonds break only when excited at or above an excited sharp predissociation threshold (SAL) [30].
HgB: There is only one theoretical study on HgB. The ground state was assigned to be the Χ2Σ+ via the B3LYP/LANL2DZ and B3LYP/SDD methodologies. Values of re = 4.381 (2.535) Å, ωe = 19 (177) cm−1, μ = 0.27 (1.20) D, EA = 0.63 (0.90) eV, IP = 7.10 (7.40) eV, and De = 0.002(0.17) eV were calculated at the B3LYP/LANL2DZ(B3LYP/SDD) levels of theory [48]. The differences between the two methods indicate that more accurate calculations should be performed using larger basis sets than the double zeta quality of the LANL2DZ and SDD basis sets.

3. Results and Discussion

DFT calculations were carried out for the ground states of the ZnB and the second- and third-row-transition-metal monoborides, MBs, and for some low-lying excited states of the NbB, TcB, LaB, TaB, ReB, and HgB molecules. Three functionals were used, i.e., B3LYP [55], MN15 [56], and TPSSh [57], in conjunction with the aug-cc-pVQZ basis set for B and the aug-cc-pVQZ-PP basis set [58,59,60,61] for all M except La; for La, the def2-QZVPPD basis set was used [62]. Bond distances, dissociation energies with respect to the adiabatic products and ground state products, frequencies, dipole moments, and natural NPA charges were calculated; Table 4 and Table 5. In the case of the NbB molecule, additional MRCISD(+Q)/aug-cc-pVQZ(-PP) calculations were carried out to clarify its ground state. Finally, the chemical bonding was analyzed, and it is presented in Table 6.

3.1. Second-Row-Transition-Metal Borides

The least studied molecule is the TcB species, where only a calculated D0 value of 3.31 eV via the B97-1/(aug)-cc-pVTZ-(PP) methodology has been reported [25]. Here, we calculated three electronic states, X3Σ, 5Σ, and 7Σ, and all the main spectroscopic data are provided; see Table 4. The ground state is a triplet state with a bond distance of 1.746 Å and a dissociation energy of 4.838 eV with respect to the adiabatic products Tc(4D)[4d6(5D)5s] + B(3P). At equilibrium, the Tc atom is excited at the 4F[4d7] state and it forms a quadruple bond, σ2σ2πx2πy2, with the boron atom; see Table 6. The dissociation energy of the ground state with the atomic ground state atoms is 3.894 eV; see discussion below. The lowest quintet and septet states are of Σ symmetry, i.e., 5Σ, and 7Σ, and they lie 0.312 eV and 2.512 eV above the ground state. They are adiabatically correlated with the Tc(6S[4d55s2]) + B(3P) dissociation channel, but at equilibrium, the Tc atom is excited at the Tc(6D[4d6(5D)5s]) + B(3P) channel. Their dissociation energies with respect to the atomic ground state products are 3.582 eV and 1.382 eV, respectively. The corresponding bond distances are 1.829 Å and 2.080 Å and the formed bonds are σ1πx2πy2 and σ1π1, respectively. The decrease in the formed bonds from the triplet to the quintet and septet can be perceived in the corresponding decrease in the dissociation energy and the corresponding increase in the bond distance. The dissociation energy per bond is about 1.2 eV in all three states, i.e., 1.21 eV (X3Σ), 1.02 eV (5Σ), and 1.38 eV (7Σ).
Table 4. Bond lengths re (Å), dissociation energies De (eV) with respect to the adiabatic atomic products (with respect to the ground state atomic products, in parenthesis), vibrational frequencies ωe (cm−1), dipole moments μ (Debye), and charge on metal qM via natural population analysis of ground and some low-lying states of the 2nd-row-transition-metal boride molecules, MBs (M = Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, and Cd) at the B3LYP, TPSSh, and MN15/aug-cc-pVQZ(-PP) levels of theory and for the NbB molecule using the MRCISD and MRCISD+Q/aug-cc-pVQZ(-PP) methodologies.
Table 4. Bond lengths re (Å), dissociation energies De (eV) with respect to the adiabatic atomic products (with respect to the ground state atomic products, in parenthesis), vibrational frequencies ωe (cm−1), dipole moments μ (Debye), and charge on metal qM via natural population analysis of ground and some low-lying states of the 2nd-row-transition-metal boride molecules, MBs (M = Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, and Cd) at the B3LYP, TPSSh, and MN15/aug-cc-pVQZ(-PP) levels of theory and for the NbB molecule using the MRCISD and MRCISD+Q/aug-cc-pVQZ(-PP) methodologies.
MBStatereDeωeμqM
B3LYP
YBX5Σ2.2313.261 (2.094)571.94.894 +0.59
ZrBX6Δ2.1623.035 (2.787)604.23.550 +0.30
NbBX5Π a1.9882.862 692.53.025 +0.15
a3Σ+1.8702.755 (2.460)805.82.984+0.18
MoBX6Π1.9732.247 654.92.221 +0.00
TcBX3Σ1.7464.838 (3.894)848.33.767 +0.00
a5Σ1.8293.582 796.31.785−0.13
7Σ2.0801.382 516.02.240+0.23
RuBX2Δ1.7005.210 (4.463)935.63.480 −0.09
RhBX1Σ+1.6795.491 (4.969)960.62.877 −0.18
PdBX2Σ+1.7773.781 759.71.281 −0.26
AgBX1Σ+2.0701.900 457.51.296 −0.11
CdBX2Π2.4660.305 237.11.684 +0.24
TPSSh
YBX5Σ2.2353.418 (2.475)574.14.893 +0.60
ZrBX6Δ2.1663.247 (3.301) 608.43.691 +0.31
NbBX5Π a1.9923.010 690.63.229 +0.17
MoBX6Π1.9822.338 646.02.416 +0.02
TcBX3Σ1.7655.210 (4.222)835.73.766 +0.01
RuBX2Δ1.7115.518 (4.443)915.03.604 −0.07
RhBX1Σ+1.6855.412 (4.991)958.43.079 −0.16
PdBX2Σ+1.7814.020 768.81.543 −0.24
AgBX1Σ+2.0561.980 474.91.555 −0.10
CdBX2Π2.4130.483 274.71.953 +0.26
MN15
YBX5Σ2.2103.649 (1.969)590.95.095 +0.59
ZrBX6Δ2.1373.347 (2.651)624.73.670 +0.29
NbBX5Π a1.9683.056 718.33.006 +0.12
MoBX6Π1.9422.465 709.01.991 −0.06
TcBX3Σ1.7245.390 (4.899)915.13.589 −0.06
RuBX2Δ1.6845.221 (5.169)986.93.375 −0.14
RhBX1Σ+1.6666.064 (5.895)1002.62.850 −0.22
PdBX2Σ+1.7624.056 783.41.084 −0.30
AgBX1Σ+2.0532.151 470.51.155 −0.15
CdBX2Π2.4620.366 218.51.238 +0.20
MRCISD
NbBX5Π2.0172.799 b709.03.096+0.16
A5Φ2.0182.715 b710.42.921+0.16
MRCISD+Q
X5Π2.0182.901 b708.9
A5Φ2.0192.808 b710.3
a The X5Π and 5Φ are almost energetically degenerate. A5Φ lies 0.084(0.093) eV above the X5Π state at the MRCISD(MRCISD+Q) level of theory. b Dissociation energy obtained via potential energy cure at infinite re values.
Regarding the ground state of NbB, the X state correlates to the atomic ground state products Nb(6D[4d45s1]) + B(2P) and it maintains this character in the minimum. DFT cannot predict if the state is a 5Π(1) or 5Φ(3), i.e., m(Nb) = ±2, +m(B) = ±1, and thus multireference techniques are necessary to clarify the electronic state. MRCISD(+Q)/aug-cc-pVQZ(-PP) calculations were carried out and it was found that the global minimum is an X5Π state and the A5Φ lies 0.084(0.093) eV above the X state. Their re values were calculated to be 2.017(2.018) Å for the X5Π state and 2.018(2.019) Å for the A5Φ, respectively. The lowest triplet state is the a3Σ+, which lies 0.402 eV above the X state. The bonding in the X state is 1 2 ( σ 1 π x 2 π y 1 ± σ 1 π x 1 π y 2 ) , while in the a3Σ+ it is σ1πx2πy2; see Table 6. The multiple bonding of 2 in the X state in contradiction to 2.5 in the 3Σ+ affects the corresponding bond distances, i.e., 1.988 vs 1.870, where a decrease of 0.1 Å is observed.
The bond distances of the ground states of the second-row-transition-metal borides, MBs, with respect to the different M atoms, are plotted in Figure 1a. Note that all three functionals present the same geometry; see Table 4. YB has a bond distance of 2.231 Å, and as the M changes along the row of the periodic table, the bond distance decreases up to RhB (1.679 Å) and then increases up to CdB at 2.466 Å.
The dissociation energy with respect to the adiabatic products (De) as the M atom changes from left to right is plotted in Figure 1d. The De value is decreased from YB up to MoB; then, the TcB, RuB, and RhB molecules have similar De values at ~5eV, i.e., 4.838 eV, 5.210 eV, and 5.491 eV, respectively; finally, the De values decrease constantly up to CdB at 0.305 eV. The dissociation energy with respect to the atomic ground state products at the zero-vibrational level, D0,GS, with respect to the change in the M atom from left to right, is plotted in Figure 1e, where we observe an increase in the D0,GS value up to NbB, then a small decease for the MoB, and then an increase up to Rh. Furthermore, the available experimental D0 values determined via R2PI and Knudsen spectrometry are plotted in Figure 1e. The calculated D0,GS values are in very good agreement with the experimental values; the % differences range from 0.00 to 0.09%.
The dipole moment, μ, follows a decreasing pattern from 5.10 Debye (YB) to 1.99 D (MoB), followed by an immediate jump at 3.59 D (TcB), and then by a small decrease at 3.38 D (RuB) and at 2.85 D (RhB). Note that a quadruple bond is formed in these three molecules. Finally, the last three MBs have similar dipole moments, i.e., 1.08 D (PdB), 1.16 D (AgB), and 1.24 D CdB; see Figure 1b.
The NPA charge on metal is positive in YB, +0.6e, i.e., about half of the electron has been moved to the B atom, and it decreases to 0 for MoB and TcB. In these two molecules, there is a charge transfer from M to B and back. Then, from Ru to Ag, the metal has a total negative charge, i.e., the charge is transferred from B to M, while in CdB, again the M is charged positively, which is expected since the M has an s2d10 atomic configuration (Figure 1c). The opposite trend is observed in harmonic vibrational frequencies, where the ωe values increase from Y (571.9 cm−1) to Rh (960.6 cm−1) and then decreased to Cd (237.1 cm−1); see Figure 1f.
The bonding of the ground and excited states of the MBs is reported in Table 6. In the X states of the first two MB molecules, YB(X5Σ) and ZrB(X6Δ), three half bonds, σ1π1π1, are formed. The additional electron in ZrB is added to the non-bonding δ(dx²−y² or dxy) orbital. As a result, they have re values that differ only by about 0.07 Å, and their De values differ only by about 0.2 eV. The next NbB(X5Π) and MoB(X6Π) form two half bonds and one whole bond, σ1π2π1. The additional e from Zr to Nb is placed in the half-occupied π orbital, while the additional electron of MoB is added to the non-bonding δ(dx²−y² or dxy) orbital. As a result, the X states of NbB and MoB have re values that differ only by about 0.01 Å and De values that differ by about 0.6 eV.
Then, in TcB(X3Σ), RuB(X2Δ), and RhB(X1Σ+), there is a significant change in the bonding compared to the first four MBs of the second row. The bonding is formed from an atomic state which has an empty 5s orbital from, for example, Tc(4F[4d7]), Ru(b3F[4d8]), and Rh(a2D[4d9]). As a result, the 5s of the metal is empty and thus it can accept electrons, forming a dative bond. The bonding in X states is 1σ22σ21πx21πy2; see Figure 2. Specifically, the bonding is 1σ2 = (M4dz²)2(B2pz)0, which is a dative bond from the M to the B atom; 2σ2 = (M5sM4dz²)0(B2s)2, also a dative bond from the B to the M; and 1πx2 = (M4dxz)1 − (B2px)1 and 1πy2 = (M4dyz)1 − (B2py)1, which are both π covalent bonds. The similar bonding leads to similar short re values, smaller up to 0.3 Å with respect to MoB, and dissociation energies of about 5 eV (4.84 up to 5.49 eV). Note that there is a 5s5dz² hybridization in M and a 2s2pz hybridization in B. The added electrons from Tc to Rh are added to non-bonding δ orbitals, i.e., δ1δ1 in TcB, δ2δ1 in RuB, and δ2δ2 in RhB.
The PdB(X2Σ+) also has an X state that is formed from an atomic state with an empty 5s orbital in Pd(1S[4d10]). However, there is a triple bond which consists of three dative bonds, i.e., 1σ2 = (Μ5s)0(Β2s)2, 1πx2 = (Μ4dxz)2(Β2px)0, and 1πy2 = (Μ4dyz)2(Β2py)0, while there is a non-bonding σ1 which corresponds to the (Β2pz)1 of B. The AgB(X1Σ+) molecule also has a triple bond with one covalent and two dative bonds, i.e., 1σ2 = (M5s)1-(B2pz)1, 1πx2 = (M4dxz)2(B2px)0 and 1πy2 = (M4dyz)2(B2py)0. Finally, CdB(X2Π) has two dative bonds, i.e., 1πx2 = (M4dxz)2(B2px)0 and 1πy2 = (M4dyz)2(B2py)0, resulting in a positive charge of +0.24e on Cd.

3.2. Third-Row-Transition-Metal Borides

The ground states of the third-row MB molecules were calculated, while for the LaB, TaB, ReB, and HgB molecules, an additional electronic state was also included; see Table 5 and Table 6. The ground state of the LaB is the Χ5Σ state, which correlates to La(4F[5d26s]) + B(2P) and maintains this character in the minimum. Its dissociation energy, D0, with respect to the ground state products is 2.086 eV, in excellent agreement with the R2PI experimental value of 2.086 eV [25]. On the contrary, the 3Π state, which is 0.448 eV above the X state, correlates to an excited La(2D[5d26s]), but the in situ La atom is b4F[5d3]. The X state has a σ1πx1πy1 bonding scheme, while the 3Π state presents a σ1πx2πy1 bonding scheme i.e., one less bonding electron, and, as a result, a shorter bond distance by 0.1 Å. The X state of the HfB is a Χ4Σ state and it has three half bonds with a dissociation energy of 2.810 eV, in good agreement with experimental value of 2.593 eV [25]. Two states have been calculated for the TaB; the X5Δ state, which has a whole bond and two half bonds, and the 3Σ+ state, which lies 0.260 eV above the ground and presents two and a half bonds. Again, the addition of one electron in the bonding results in a decrease in the bond distance by 0.1 Å. The dissociation energy of X5Δ with respect to the adiabatic products is 3.495 eV. The X state of WB is the Χ6Π, where a σ1π2π1 bond scheme is formed. It presents a dissociation energy of 2.907 eV and re = 1.955 Å. The 6Σ+ lies 0.137 eV above the X state; it forms a σ2πx1πy1 bond with a dissociation energy of 2.770 eV and re = 2.125 Å. The ground state of ReB is the Χ5Σ, state which correlates to Re(6S[5d56s2]), but the in situ atom is the Re(6D[5d66s1]), which forms two and a half bonds. The dissociation energy is 3.106 eV and the bond distance at 1.834 Å. The a3Σ state lies only 0.099 eV above the ground state, and it has a stronger bonding, i.e., three bonds (σ2πx2πy2). Its bond distance is 1.809 Å, and its dissociation energy is 4.967 eV with respect to the correlated products, while with respect to the atomic ground state products, it is 3.008 eV.
Table 5. Bond lengths, re (Å); dissociation energies, De (eV) with respect to the adiabatic atomic products (with respect to the ground state atomic products, in parenthesis); vibrational frequencies, ωe (cm−1); dipole moments, μ (Debye); and charge on metal qM via natural population analysis of ground and some low-lying states of the 3rd-row-transition-metal boride molecules, MBs (M = La, Hf, Ta, W, Re, Os, Ir, Pt, Au, and Hg): at the B3LYP, TPSSh, and MN15/aug-cc-pVQZB(-PP)M and def2-QZVPPDLa levels of theory.
Table 5. Bond lengths, re (Å); dissociation energies, De (eV) with respect to the adiabatic atomic products (with respect to the ground state atomic products, in parenthesis); vibrational frequencies, ωe (cm−1); dipole moments, μ (Debye); and charge on metal qM via natural population analysis of ground and some low-lying states of the 3rd-row-transition-metal boride molecules, MBs (M = La, Hf, Ta, W, Re, Os, Ir, Pt, Au, and Hg): at the B3LYP, TPSSh, and MN15/aug-cc-pVQZB(-PP)M and def2-QZVPPDLa levels of theory.
MoleculeStatere (Å)DeωeμqM
B3LYP
LaBΧ5Σ2.3842.874(2.528)512.74.03+0.55
a3Π2.2632.080478.56.08+0.74
HfBΧ4Σ2.1512.810594.82.55+0.35
TaBX5Δ 1.9743.495739.63.58+0.22
a3Σ+1.8703.236838.23.78+0.28
WBΧ6 Π1.9552.907 a735.02.63+0.04
6Σ+2.1252.770 a548.22.86+0.10
ReBΧ5Σ1.8343.106865.52.28−0.07
a3Σ1.8094.967(3.008)850.72.85−0.05
OsBΧ4Σ1.7724.482935.12.18−0.13
IrBΧ3Δ1.7625.338926.91.62−0.24
PtBΧ2Σ+1.7595.377905.21.08−0.32
AuBΧ1Σ+1.9253.451642.20.87−0.26
HgBX2Π2.3970.237223.41.38+0.17
HgBA2Σ+4.0290.00331.70.350.00
TPSSh
LaBΧ5Σ2.3843.032(2.786)518.64.40+0.56
HfBΧ4Σ2.1493.004606.12.55+0.36
TaBX5Δ1.9793.585733.53.65+0.23
WBΧ6 Π1.9632.938 a720.52.74+0.06
6Σ+2.1262.731 a549.52.99+0.12
ReBΧ5Σ1.8383.519863.32.41−0.05
OsBΧ4Σ1.7794.646909.92.59−0.07
IrBΧ3Δ1.7655.300928.11.81−0.22
PtBΧ2Σ+1.7595.547917.41.19−0.31
AuBΧ1Σ+1.9213.604657.61.04−0.25
HgBX2Π2.3410.395266.51.62+0.19
HgBA2Σ+3.6580.01642.90.50−0.009
MN15
LaBΧ5Σ2.3653.237(2.402)526.14.64+0.55
HfBΧ4Σ2.1333.123603.52.46+0.33
TaBX5Δ1.9573.560767.03.42+0.19
WBΧ6 Π1.9323.031 a778.12.470.00
ReBΧ5Σ1.8143.346904.42.02−0.11
OsBΧ4Σ1.7584.943958.71.98−0.17
IrBΧ3Δ1.7505.633944.71.46−0.27
PtBΧ2Σ+1.7445.744934.80.87−0.35
AuBΧ1Σ+1.9023.933673.90.69−0.30
HgBX2Π2.3370.364234.41.12+0.14
HgBA2Σ+3.3880.08566.90.68−0.02
a With respect to the W(7S) state, which is the lowest atomic state with respect to the average J term.
The OsB(Χ4Σ), IrB(Χ3Δ), and PtB(Χ2Σ+) molecules present triple bonds in their ground states, with similar bond distances of 1.772 Å, 1.762 Å, and 1.759 Å, which is expected due to the same bonding. Note that the additional electron moving from Os to Pt is added to the non-bonding d+2 or d−2 atomic orbital. The dissociation energies with respect to the adiabatic products are 4.482 eV for OsB, 5.338 eV for IrB, and 5.377 eV for PtB. The IrB and OsB molecules correlate to their in situ atoms, while for OsB, the state of the M atom in the minimum differs from the correlated one. Thus, the dissociation energy of OsB with respect to its in situ atoms is 5.12 eV, only 0.2 eV less than the De value of IrB and PtB. On the contrary, whilst a triple bond is also formed in the X state of AuB (Χ1Σ+), the two π2 bonds which are dative, i.e., Au(5dxz)2B(2px)2 and Au(5dyz)2B(2py)0, are weak with a small charge transfer. Finally, two states were calculated for the HgB molecule, X2Π and A2Σ+. The X state has a weak double bond with an elongated bond distance of 2.397 Å and a dissociation energy of 0.237 eV, while the A2Σ+ is a van der Waals state, where very weak interactions are formed between atoms. Comparing the D0 dissociation energies (with respect to the ground state products) with the corresponding experimental values, there is a very good agreement, i.e., calculated[expt] values of 4.424[4.378] eV for OsB, 5.281[4.928] eV for IrB, 5.321[5.235] eV for PtB, and 3.411[3.724] eV for AuB. Thus, differences in D0 that range from 0.05 eV to 0.35 eV are observed.
The bond distances of the ground states of the third-row-transition-metal MB molecules, with respect to the different M atoms, are plotted in Figure 3a. Note that all three functionals present the same geometry (Table 5). The first molecule of the row, LaB, has a bond distance of 2.384 Å, and, as the M changes along the row of the periodic table, the bond distance decreases up to PtB (1.759 Å) and then increases up to HgB at 2.397 Å. It should be noted that OsB(Χ4Σ), IrB(Χ3Δ), and PtB(Χ2Σ+) present similar bond distances, i.e., 1.772 Å, 1.762 Å, and 1.759 Å, respectively, and similar triple bonding, σ2π2π2.
The dipole moment, μ, follows a decreasing pattern from 4.03 Debye (LaB) to 0.87 D (AuB), except for the increase from Hf to Ta. Finally, there is an increase from AuB to HgB at 1.38 D; see Figure 3b. The NPA charge on the metal is positive in LaB, +0.55e, i.e., about half of the electrons are moved to the B atom, and the charge decreases to WB (+0.04 e); see Figure 3c. Then, from Re to Au, the metal has a total negative charge, i.e., the charge is transferred from B to M, while in HgB, again, the M is charged positively (+0.17e), which is expected since the M has an s2d2 atomic configuration. In PtB, the Pt demonstrates the most negative charge of all the metals. Finally, the harmonic vibrational frequencies generally increase from La to OsB. OsB, IrB, and PtB have similar ωe values at about ~920 cm−1; then, they decrease to HgB (223.4 cm−1); see Figure 3f.
The dissociation energy with respect to the adiabatic products, De, as the M atom changes from left to right is plotted in Figure 3d. The first five MBs have similar De values, with some discrepancies. Then, in OsB, IrB, and PtB, the De value increases; their values are 4.482 eV, 5.338 eV, and 5.377 eV, respectively, and finally, the De values decrease constantly up to HgB at 0.237 eV. The dissociation energy with respect to the atomic ground state products at the zero-vibrational level, D0,GS, with respect to the change in the M atom is plotted in Figure 3e. The available experimental D0 values are also plotted in Figure 3e via R2PI and Knudsen spectrometry. The calculated D0,GS values for HfB, OsB, WB, and PtB are in excellent agreement, and for LaB and IrB, they are in good agreement, while the largest difference is observed for TaB. Finally, it should be noted that the B3LYP-obtained values are in better agreement with the experimental ones than the values obtained via the MN15 and TPSSh functional methods.
The bonding of the ground and excited states of the MBs is reported in Table 6. In the ground states of the first two MB molecules, i.e., LaB and HfB, three half bonds, σ1πx1πy1, are formed. As a result, they have the same dissociation energies, i.e., 2.874 eV and 2.810 eV. It should be noted that there is a strong 5dz²6s hybridization in the M atom. The TaB and WB form two half bonds and one whole bond, i.e., σ1πx2πy1. The following ReB has one half and two bonds, i.e., σ1πx1πy2, while the next four molecules, OsB, IrB, PtB, and AuB, form a triple bond, i.e., σ2πx2πy2. Note that, contrary to the second row, where the atomic states of M atoms with empty 5s orbitals are involved in the bond, the corresponding atomic states of the M atoms occupy the 6s atomic orbital. Finally, the last MB of this row, i.e., HgB, has two dative bonds, i.e., 1πx2 = M(5dxz)2B(2px)0 and 1πy2 = M(5dyz)2B(2py)0, similar to the CdB of the second row, thus leading to similar bond lengths, 2.466 Å (CdB) and 2.397 Å (HgB), and dissociation energies, 0.305 eV and 0.237 eV, respectively.

3.3. Comparison of MBs of All Three Rows and Bonding Analysis

Quantum chemical computations provide details on the chemical bonds and electronic structures of these species. In general, the bond lengths of the transition-metal borides increase in the periodic table from left to right, or as one goes down a group of elements. Of course, anomalies occur, but the data are the result of the variety of bonding schemes that are formed in the MB molecules. These bonding schemes depend on the bonding and the filling of σ, π, and δ orbitals. MBs can form one-and-a-half, double, triple, and even quadruple bonds, with the latter being recently discovered in RhB and RhB [2,4,5], while here, we found that RuB and TcB also form quadruple bonds.
The bond distances of the ground states of the MBs of all three rows, with respect to the different M atoms, are plotted in Figure 4a. Moving from up (first row) to down (third row) the bond distances, in the cases where there is the same bonding, the re values increase by about 1 Å. Also, moving from left to right, the re value decreases up to the seventh or eighth MB, and then the re value increases sharply. All differences are a result of the type of bonding. In all rows, there are three MBs that present similar strong bonding. Thus, along the first row, from the sixth to the eight MB, i.e., FeB, CoB, and NiB, a triple bond and similar re values, around 1.7 Å, are observed. Similarly, along the third row, from the sixth to the eighth MB, OsB, IrB, and PtB, a triple bond and similar re values, around 1.76 Å, are found. On the contrary, in the second row, from the fifth to the seventh MB, i.e., TcB, RuB, and RhB, a quadruple bond is formed with similar re values, around 1.7 Å.
In Figure 4d,e, the dissociation energies with respect to the adiabatic products (De) and the atomic ground state products at the zero-vibrational level (D0,GS) are plotted with respect to the change in the M atom from left to right for the three rows. The general shape is the same as that of the lowest dissociation energies being observed for the M, with atomic configurations of (ds)6 or (ds)7 and d10s2, while the highest values are observed for the M with atomic configurations of (ds)8–10 for the first and third rows and (ds)7–9 for the second row, and the largest values are found for the second and third rows. Finally, it should be noted that the first two and the last two MBs present similar dissociation energy regardless of the row.
The dipole moment, μ, with respect to the M follows roughly the same trend as the M changes from left to right in each row, i.e., a general reduction in the μ value. Any differences are related to the differences in the bonding; see Figure 4b. The smallest differences among the three rows are observed in d1s2, (ds)5, and d10s2, while the first and third rows present similar μ values for the d5s2 metals, where the second-row MB presents a μ value larger by 1.5 D than the corresponding MnB and ReB values due to the quadruple bond of TcB.
Table 6. Atomic states of transition metals forming the bonding (in situ M) of the ground MB state (X) and the atomic state of M in RM-B infinity; and of the atomic ground state XM and the energy difference between the ground atomic state and the atomic state forming the bonding Te (eV) (MJ-averaged experimental).
Table 6. Atomic states of transition metals forming the bonding (in situ M) of the ground MB state (X) and the atomic state of M in RM-B infinity; and of the atomic ground state XM and the energy difference between the ground atomic state and the atomic state forming the bonding Te (eV) (MJ-averaged experimental).
MBXConfigurationBondIn Situ MIn RM-B Infinity XMΤe b
1st row MB M
ScBX5Σ21111σ1πx1πy14F[3d24s1] a4F[3d24s1]2D[3d14s2]1.428(1.427)
TiBX6Δ211111σ1πx1πy15F[3d3(4F)4s1] a5F[3d3(4F)4s1]a3F[3d24s2 ]0.813(0.806)
VBX7Σ+2111111σ1πx1πy16D[3d4(5D)4s1] a6D[3d4(5D)4s1]a4F[3d34s2]0.262(0.245)
CrBX6Σ+2211111σ2πx1πy17S[3d5(6S)4s1] a7S[3d5(6S)4s1]7S[3d5(6S)4s1]0
MnBX5Π2212111σ2πx2πy16S[3d54s2] a6S[3d54s2]a6S[3d54s2]0
FeBΧ4Σ2212211σ2πx2πy2a5F[3d7(4F)4s1] aa5D[3d64s2]a5D[3d64s2]0.859(0.875)
CoBΧ3Δ2212221σ2πx2πy2b4F[3d8(3F)4s1] aa4F[3d74s2] a4F[3d74s2]0.432(0.417)
NiBΧ2Σ+2212222σ2πx2πy2a3D[3d9(2D)4s1] aa3F[3d8(3F)4s2]a3F[3d8(3F)4s2]0.025(-0.030)
CuBΧ1Σ+2222222σ2πx2πy22S[3d10(1S)4s1] a2S[3d10(1S)4s1]2S[3d10(1S)4s1]0
2nd row
ZnBΧ2Π22222122σ2π21S[3d104s2]1S[3d104s2]1S[3d104s2]0
YBX5Σ21111σ1πx1πy14F[4d25s1]a4F[4d2(3F)5s]a2D[4d5s2]1.356(1.359)
ZrBX6Δ211111σ1πx1πy15F[4d3(4F)5s1]a5F[4d3(4F)5s1]a3F[4d25s2]0.604(0.588)
NbB5Π/5Φ211211σ1πx2πy1a6D[4d4(5D)5s1]a6D[4d4(5D)5s1]a6D[4d4(5D)5s1]0
3Σ+21122σ1πx2πy2a4D[4d45s1]a4F[4d35s3]a6D[4d4(5D)5s1]1.043(1.049)
MoBX6Π2112111σ1πx2πy1a7S[4d5(6S)5s]a7S[4d5(6S)5s]a7S[4d5(6S)5s]0
TcBX3Σ222211σ2σ2πx2πy24F[4d7]4D[4d6(5D)5s]6S[4d55s2]1.827(2.332)
5Σ2112211σ1πx2πy26D[4d65s1]6S[4d55s2]6S[4d55s2]0.319(0.406)
7Σ21121111σ1π16D[4d65s1]6S[4d55s2]6S[4d55s2]0.319(0.406)
RuBX2Δ222221σ2σ2πx2πy2b3F[4d8]a3F[4d7(a4F)5s]a5F[4d7(a4F)5s]1.131(1.092)
RhBX1Σ+222222σ2σ2πx2πy2a2D[4d9]a2D[4d9]a4F[4d8(3F)5s]0.410(0.342)
PdBX2Σ+2212222σ2πx2πy21S[4d10]1S[4d10]1S[4d10]0
AgBX1Σ+2222222σ2πx2πy22S[4d105s]2S[4d105s]2S[4d105s]0
CdBX2Π22222122σ2π21S[4d105s2]1S[4d105s2]1S[4d105s2]0
3rd row
LaBX5Σ21111σ1πx1πy14F[5d2(3F)6s]4F[5d2(3F)6s]2D[5d6s2]0.331(0.355)
3Π2121σ1πx2πy1b4F[5d3]2D[5d6s2]2D[5d6s2]1.541(1.560)
HfBX4Σ22111σ1πx1πy1a3F[5d26s2]a3F[5d26s2]a3F[5d26s2]0
TaBX5Δ211211σ1πx2πy1a6D[5d46s1]a4F[5d36s2]a4F[5d36s2]1.210(1.038)
3Σ+21122σ1πx2πy2 a4F[5d36s2]a4F[5d36s2]
WBX6Π2112111σ1πx2πy17S[5d56s1]5D[5d46s2]5D[5d46s2]0.366(−0.187)
6Σ+2211111σ2πx1πy17S[5d56s1]5D[5d46s2]5D[5d46s2]0.366(−0.187)
ReBX5Σ2112211σ1πx2πy2a6D[5d66s1]a6S[5d56s2]a6S[5d56s2]1.457(1.759)
3Σ222211σ2πx2πy2a4P[5d56s2]a4P[5d56s2]a6S[5d56s2]1.436(1.603)
OsBΧ4Σ2212211σ2πx2πy2a5F[5d7(4F)6s1]a5D[5d66s2]a5D[5d66s2]0.638(0.757)
IrBΧ3Δ2212221σ2πx2πy2a4F[5d76s2]a4F[5d76s2]a4F[5d76s2]0
PtBΧ2Σ+2212222σ2πx2πy23D[5d96s1]3D[5d96s1]3D[5d96s1]0
AuBΧ1Σ+2222222σ2πx2πy22S[5d106s1]2S[5d106s1]2S[5d106s1]0
HgBΧ2Π22222122σ2π21S[5d106s2]1S[5d106s2]1S[5d106s2]0
Χ2Σ+222122222π2)1S[5d106s2]1S[5d106s2]1S[5d106s2]0
a Ref. [21]. b Expt values of the energy separation between the term with the lowest in energy spin–orbit coupling angular momentum quantum number J (average term).
On the contrary, the NPA charge on the metals as the M changes from left to right has the same shape for all three rows apart, from the MnB of the first row; see Figure 4c. Finally, the change in the vibrational frequencies with respect to M for the three rows has the same trend; see Figure 4f. In most MB molecules, the second- and third-row MBs have larger ωe values than the MBs of the first row.
The dissociation energy per bond, i.e., per two bonding electrons, and the number of the formed bonds for all MBs of the first, second, and third rows are depicted in Figure 5. Regarding the diagram of the De/bond, the general shape is the same for all three rows. The largest De/bond, except for the first two and the last MBs, is observed for the MBs of the third row. Thus, the fact that the 4d series exhibits greater intrinsic bond energies [26] is explained with the formation of more multiple bonds than in some MBs of the 3d and 5d series, while the De/bond is the highest for the third row, as expected.
Regarding the number of formed bonds, it seems that the MB molecules of the first and third rows present the same multiple bonds, except the (n)d3(n + 1)s2 atoms, where, for VB, the ground state is X7Σ+ and for ΤaB it is X5Δ. The difference in spin multiplicities results in different orders of bonding. The largest observed difference between MBs of the second row with the corresponding MBs of the first and third rows is for the fifth MB, i.e., TcB (quadruple bond) vs MnB and ReB (two and a half bonds). Finally, the X states of the RuB and RhB molecules of the second row also have quadruple bonds, while the corresponding MB molecules of the other rows have triple bonds.
The formation of quadruple bonding in the MB part of the complexes has been reported in the literature. For instance, in the anionic complex FeB(CO)3, a quadruple bond is formed in the FeB part. The BFe(CO)3 anion was calculated via DFT and DLPNO-CCSD(T) methodologies and identified using mass-selected infrared photodissociation spectroscopy in the gas phase [63].
Finally, it should be noted that in the transition-metal molecules, highly correlated electrons are involved in the spin−orbit interactions, and relativistic effects exist. The use of basis sets with pseudopotentials of course simplified the complicated effects of the motion of the core (non-valence) electrons. Furthermore, 5d MB molecules suffer from strong spin−orbit effects, which here have been neglected. The use of atomic spin−orbit stabilization [64] may lead to reductions in the calculated dissociation energies [25]. However, our calculated values are in very good agreement with the experimental ones, with the exception of TaB, since both M and MB species are subject to the spin–orbit effects; thus, via the cancellation of errors, the De results are good. Specifically, our B3LYP/aug-cc-PVQZ(-PP) values are in better agreement with the experimental ones, and the energy differences between experimental and calculated values range from 0.046 (OsB) to 0.410 eV (LaB), apart from TaB, where the calculated value overestimates the experimental D0 by 0.749 eV. On the contrary, for the second-row MBs, the energy differences between experimental and calculated values range from 0.002 eV to 0.343 eV. Finally, note that the main aim of the present study is to study the periodic bonding schemes and trends for MB molecules over three rows, providing a quantitative and qualitative picture of the chemical bonding in the MB species, which is presented here adequately.

4. Computational Details

The ground states of the ZnB and the second- and third-row-transition-metal monoborides, MBs, were calculated employing the B3LYP [55,65], MN15 [56], and TPSSh [57] functionals, along with the correlation-consistent basis sets of Dunning et al., i.e., the aug-cc-pVQZ-PP basis set for all M samples except La, and the aug-cc-pVQZ basis set for B [58,59,60,61]. For the La atom, the def2-QZVPPD basis set was used [62]. Additionally, low-lying excited states were calculated for the NbB, TcB, LaB, TaB, ReB, and HgB molecules. Bond lengths, dissociation energies with respect to the adiabatic products and with respect to ground state products, frequencies, and dipole moments were calculated. The charges were obtained via the natural population analysis, NPA. In the case of the NbB molecule, additional MRCISD (MRCISD+Q)/aug-cc-pVQZ(-PP) calculations were carried out to clarify the ground state. In their reference CASSCF calculations, the eight valence electrons were distributed in ten orbitals (5s4d + 2s2p), resulting in 2060 configuration state functions (CSFs), while the number of MRCISD CSFs were 5.9 × 107 and they were reduced to 1.1 × 107 after the internally contracted approach. All DFT calculations were performed using the GAUSSIAN package [66]. The multireference calculations were carried out using the MOLPRO package [67].

5. Summary and Conclusions

Boron plays an important role in chemistry, biology, and materials science. Diatomic transition-metal borides are important building blocks of many complexes and materials and thus the knowledge of their dissociation energy, bond distances, and bonding analysis dipole moments are very useful. It is interesting that boron forms a variety of orders of bonding from single to quadruple bonds and bonds of different types, i.e., covalent, dative, and ionic bonds. In the present paper, the diatomic borides of transition metals are reviewed and studied.
In the first part, a review on the available experimental and theoretical studies on the first-row-transition-metal borides, i.e., ScB, TiB, VB, CrB, MnB, FeB, CoB, NiB, CuB, and ZnB; the second-row-transition-metal borides, i.e., YB, ZrB, NbB, MoB, RuB, RhB, PdB, AgB, and CdB; and the third-row-transition-metal borides, i.e., LaB, HfB, TaB, WB, ReB, OsB, IrB, PtB, AuB, and HgB, is presented. There was a gap in the literature regarding TcB, which is studied here for the first time in detail. While where there were doubts regarding which state is the ground state for some MBs, here, it is clarified.
In the second part, the second- and third-row-transition-metal borides, MBs, are studied via DFT calculations using the B3LYP, TPSSh, and MN15 functionals in conjunction with the aug-cc-pVQZ-PPM/aug-cc-pVQZB basis sets. In the case of the NbB molecule, additional multireference calculations, i.e., MRCISD(MRCISD+Q)/aug-cc-pVQZ(-PP), were carried out to clarify which state is the ground state, because DFT could not decipher it. Bond distances, dissociation energies, frequencies, dipole moments, and natural NPA charges are presented. Comparisons between the MB molecules of all three rows and their bonding are presented. All differences and similarities are analyzed. Both result from the differences in bonding schemes.
It was found that, apart from RhB which was recently reported to form quadruple bonds [2,4], RuB and TcB form quadruple bonds in their ground states as well. The X states of these three molecules present the same quadruple bonding, σ2σ2πx2πy2, and as the metal is escalated from Tc to Rh, the additional valence electron is added to the single occupied d orbitals, while the X states change from X3Σ (TcB) to X2Δ (RuB) and then to X1Σ+ (RhB). Finally, here, we studied the TcB molecule, i.e., three states were calculated, filling the gap that existed in the literature.
As a final remark, it has been reported that the states of the diatomic and triatomic molecules of sulfides are involved in complexes and solid or 2D materials as building blocks, explaining the variety of their morphologies [19,20]. Similarly, the present data may present a new approach to exploring the properties of solid state and 2D metastable polymorphic materials involving transition-metal borides, while they may assist in explaining the catalytic properties of complexes, including transition-metal boride bonds.

Author Contributions

C.D., C.E.T. and A.A.: Investigation, data curation, formal analysis, writing—original draft preparation; D.T.: conceptualization, methodology, resources, validation, supervision, writing—original draft, review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are provided in the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Greenwood, N.N.; Earnshaw, A. Chemistry of the Elements, 2nd ed.; Butterworth-Heinemann: Oxford, UK, 1998. [Google Scholar]
  2. Cheung, L.F.; Chen, T.-T.; Kocheril, G.S.; Chen, W.-J.; Czekner, J.; Wang, L.-S. Observation of Fourfold Boron-Metal Bonds in RhB(BO) and RhB. J. Phys. Chem. Lett. 2020, 11, 659–663. [Google Scholar] [CrossRef] [PubMed]
  3. Shaik, S.; Danovich, D.; Wu, W.; Su, P.; Rzepa, H.S.; Hiberty, P.C. Quadruple Bonding in C2 and Analogous Eight-Valence Electron Species. Nat. Chem. 2012, 4, 195–200. [Google Scholar] [CrossRef] [PubMed]
  4. Tzeli, D.; Karapetsas, I. Quadruple Bonding in the Ground and Low-Lying Excited States of the Diatomic Molecules TcN, RuC, RhB, and PdBe. J. Phys. Chem. A 2020, 124, 6667–6681. [Google Scholar] [CrossRef] [PubMed]
  5. Tzeli, D. Quadruple chemical bonding in the diatomic anions TcN¯, RuC¯, RhB¯, and PdBe¯. J. Comput. Chem. 2021, 42, 1126–1137. [Google Scholar] [CrossRef] [PubMed]
  6. Ganem, B.; Osby, J.O. Synthetically useful reactions with metal boride and aluminide catalysts. Chem. Rev. 1986, 86, 763–780. [Google Scholar] [CrossRef]
  7. Will, G. Electron deformation density in titanium diboride chemical bonding in TiB2. J. Solid State Chem. 2004, 177, 628–631. [Google Scholar] [CrossRef]
  8. Choi, H.J.; Roundy, D.; Sun, H.; Cohen, M.L.; Louie, S.G. The origin of the anomalous superconducting properties of MgB(2). Nature 2002, 418, 758–760. [Google Scholar] [CrossRef]
  9. Chung, H.; Weinberger, M.B.; Levine, J.B.; Kavner, A.; Yang, J.; Tolbert, S.H.; Kaner, R.B. Synthesis of Ultra-Incompressible Superhard Rhenium Diboride at Ambient Pressure. Science 2007, 316, 436–439. [Google Scholar] [CrossRef]
  10. Wdowik, U.D.; Twardowska, A.; Rajchel, B. Vibrational Spectroscopy of Binary Titanium Borides: First-Principles and Experimental Studies. Adv. Condens. Matter Phys. 2017, 18, 4207301. [Google Scholar] [CrossRef]
  11. Decker, B.F.; Kasper, J.S. The crystal structure of TiB. Acta Crystallogr. 1954, 7, 77–80. [Google Scholar] [CrossRef]
  12. Spear, K.E.; McDowell, P.; McMahon, F. Experimental evidence for the existence of the Ti3B4 phase. J. Am. Ceram. Soc. 1986, 69, C-4–C-5. [Google Scholar] [CrossRef]
  13. Viswanathan, E.; Sundareswari, M.; Jayalakshmi, D.S.; Manjula, M. Fermi surface and hardness enhancement study on ternary scandium and vanadium-based borides by first principles investigation. Comput. Mater. Sci. 2019, 157, 107–120. [Google Scholar] [CrossRef]
  14. Zhang, Y.; Liu, D.; Zhao, Y.; Li, W.; Gao, Y.; Duan, M.; Hou, H. Physical Properties and Electronic Structure of Cr2B Under Pressure. Phys. Status Solidi B 2020, 258, 2000212. [Google Scholar] [CrossRef]
  15. Gou, H.; Steinle-Neumann, G.; Bykova, E.; Nakajima, Y.; Miyajima, N.; Li, Y.; Ovsyannikov, S.V.; Dubrovinsky, L.S.; Dubrovinskaia, N. Stability of MnB2 with AlB2-type structure revealed by first-principles calculations and experiments. Appl. Phys. Lett. 2013, 102, 061906. [Google Scholar] [CrossRef]
  16. Panda, K.B.; Ravi Chandran, K.S. First principles determination of elastic constants and chemical bonding of titanium boride (TiB) on the basis of density functional theory. Acta Mater. 2006, 54, 1641–1657. [Google Scholar] [CrossRef]
  17. Li, P.; Zhou, R.; Cheng Zeng, X. Computational Analysis of Stable Hard Structures in the Ti-B System. Appl. Mater. Interfaces 2015, 7, 15607–15617. [Google Scholar] [CrossRef]
  18. Wang, M.; Liu, C.; Wen, M.; Li, Q.; Ma, Y. Investigations on structural determination of semi-transition-metal borides. Phys. Chem. Chem. Phys. 2017, 19, 31592–31598. [Google Scholar] [CrossRef]
  19. Tzeli, D.; Karapetsas, I.; Merriles, D.M.; Ewigleben, J.C.; Morse, M.D. The molybdenum-sulfur bond: Electronic structure of low-lying states of MoS. J. Phys. Chem. A 2022, 126, 1168–1181. [Google Scholar] [CrossRef]
  20. Mermigki, M.A.; Karapetsas, I.; Tzeli, D. Electronic structure of low-lying states of triatomic MoS2 molecule: The building block of 2D MoS2. Chem. Phys. Chem. 2023, 24, e202300365. [Google Scholar] [CrossRef]
  21. Tzeli, D.; Mavridis, A. Electronic structure and bonding of the 3D–transition metal borides, MB, M = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, and Cu through all electron ab initio calculations. J. Chem. Phys. 2008, 128, 034309. [Google Scholar] [CrossRef]
  22. Wu, Z. Density functional study of 3D-metal monoborides. J. Mol. Struct. Theochem. 2005, 728, 167–172. [Google Scholar]
  23. Černušák, I.; Dallos, M.; Lischka, H.; Müller, T.; Uhlár, M. On the ground and some low-lying excited states of ScB: A multiconfigurational study. J. Chem. Phys. 2007, 126, 214311. [Google Scholar] [CrossRef] [PubMed]
  24. Gingerich, K.A. Gaseous Metal Borides. I. Dissociation Energy of the Molecules ThB, ThP, and Th2, and Predicted Dissociation Energies of Selected Diatomic Transition-Metal Borides. High Temp. Sci. 1969, 1, 258–267. [Google Scholar]
  25. Merriles, D.M.; Nielson, C.; Tieu, E.; Morse, M.D. Chemical Bonding and Electronic Structure of the Early Transition Metal Borides: ScB, TiB, VB, YB, ZrB, NbB, LaB, HfB, TaB, and WB. J. Phys. Chem. A 2021, 125, 4420–4434. [Google Scholar] [CrossRef] [PubMed]
  26. Merriles, D.M.; Tieu, E.; Morse, M.D. Bond dissociation energies of FeB, CoB, NiB, RuB, RhB, OsB, IrB, and PtB. J. Chem. Phys. 2019, 151, 044302. [Google Scholar] [CrossRef] [PubMed]
  27. Ng, Y.W.; Pang, H.F.; Cheung, A.S.-C. Electronic transitions of cobalt monoborides. J. Chem. Phys. 2011, 135, 204308. [Google Scholar] [CrossRef]
  28. Balfour, W.J.; Chowdhury, P.K.; Li, R. Ni+B2H6: Spectroscopic observations on NiB and NiH. Chem. Phys. Lett. 2008, 463, 25–28. [Google Scholar] [CrossRef]
  29. Barysz, M.; Urban, M. Molecular Properties of Boron–Coinage Metal Dimers: BCu, BAg, Bau. Adv. Quantum Chem. 1997, 28, 257. [Google Scholar]
  30. Merriles, D.M.; Morse, M.D. CrN, CuB, and AuB: A Tale of Two Dissociation Limits. J. Phys. Chem. Lett. 2023, 14, 7361–7367. [Google Scholar] [CrossRef]
  31. Dore, J.M.; Adam, A.G.; Tokaryk, D.W.; Linton, C. Hyperfine analysis of the (2, 0) [18.3]3–X3Δ3 transition of cobalt monoborides. J. Mol. Spectrosc. 2019, 360, 44–48. [Google Scholar] [CrossRef]
  32. Zhen, J.F.; Wang, L.; Qin, C.B.; Zhang, Q.; Chen, Y. Laser-induced Fluorescence and Dispersed Fluorescence Spectroscopy of NiB: Identification of a New 2Π State in 19000–22100 cm−1. Chin. J. Chem. Phys. 2010, 23, 626–629. [Google Scholar] [CrossRef]
  33. Goudreau, E.S.; Adam, A.G.; Tokaryk, D.W.; Linton, C. High resolution laser spectroscopy of the [20.6]0.5–X2Σ+ transition of nickel monoboride, NiB. J. Mol. Spectrosc. 2015, 314, 13–18. [Google Scholar] [CrossRef]
  34. Kharat, B.; Deshmukh, S.B.; Chaudhari, A. 4d Transition Metal Monoxides, Monocarbides, Monoborides, Mononitrides, and Monofluorides: A Quantum Chemical Study. J. Quantum Chem. 2009, 109, 1103–1115. [Google Scholar] [CrossRef]
  35. Borin, A.C.; Gobbo, J.P. Electronic Structure of the Ground and Low-Lying Electronic States of MoB and MoB+. Int. J. Quantum Chem. 2011, 111, 3362–3370. [Google Scholar] [CrossRef]
  36. Auwera-Mahieu, A.V.; Peeters, R.; McIntyre, N.S.; Drowa, J. Mass Spectrometric Determination of Dissociation Energies of the Borides and Silicides of some Transition Metals. Trans. Faraday Soc. 1970, 66, 809–816. [Google Scholar] [CrossRef]
  37. Wang, N.; Ng, Y.W.; Cheung, A.S.-C. Laser induced fluorescence spectroscopy of ruthenium monoborides. Chem. Phys. Lett. 2012, 547, 21–23. [Google Scholar] [CrossRef]
  38. Chowdhury, P.K.; Balfour, W.J. A spectroscopic characterization of the electronic ground state of rhodium monoborides. J. Chem. Phys. 2006, 124, 216101. [Google Scholar] [CrossRef]
  39. Gobbo, J.P.; Borin, A.C. The nature of the [20.0] 1Sigma+ electronic state of RhB: A multiconfigurational study. J. Chem. Phys. 2007, 126, 011102. [Google Scholar] [CrossRef]
  40. Chowdhury, P.K.; Balfour, W.J. A spectroscopic study of the rhodium monoboride molecule. Mol. Phys. 2007, 105, 1619–1624. [Google Scholar] [CrossRef]
  41. Borin, A.C.; Gobbo, J.P. Low-Lying Singlet and Triplet Electronic States of RhB. J. Phys. Chem. A 2008, 112, 4394–4398. [Google Scholar] [CrossRef]
  42. Schoendorff, G.; Ruedenberg, K.; Gordon, M.S. Multiple Bonding in Rhodium Monoboride. Quasi-atomic Analyses of the Ground and Low-Lying Excited States. J. Phys. Chem. A 2021, 125, 4836–4846. [Google Scholar] [CrossRef] [PubMed]
  43. Knight, L.B., Jr.; Babb, R.; Hill, D.W.; McKinley, A.J. Laser vaporization generation of the diatomic radicals PdB, 105PdB, PdAl, and 105PdAl: Electron spin resonance investigation in neon matrices at 4 K. J. Chern. Phys. 1992, 97, 2987. [Google Scholar] [CrossRef]
  44. Ng, Y.W.; Pang, H.F.; Qian, Y.; Cheung, A.S.-C. Electronic Transition of Palladium Monoboride. J. Phys. Chem. A 2012, 116, 11568–11572. [Google Scholar] [CrossRef] [PubMed]
  45. Merriles, D.M.; Morse, M.D. Ionization energies and cationic bond dissociation energies of RuB, RhB, OsB, IrB, and PtB. J. Chem. Phys. 2022, 157, 074303. [Google Scholar] [CrossRef] [PubMed]
  46. Ariyarathna, I.R.; Duan, C.; Kulik, H.J. Understanding the chemical bonding of ground and excited states of HfO and HfB with correlated wavefunction theory and density functional approximations. J. Chem. Phys. 2022, 156, 184113. [Google Scholar] [CrossRef]
  47. Cheung, L.F.; Kocheril, G.S.; Czekner, J.; Wang, L.S. The nature of the chemical bonding in 5d transition-metal diatomic borides MB (M = Ir, Pt, Au). J. Chem. Phys. 2020, 152, 174301. [Google Scholar] [CrossRef]
  48. Kalamse, V.; Gaikwad, S.; Chaudhari, A. Computational study of 5d transition metal mononitrides and monoborides using density functional method. Bull. Mater. Sci. 2010, 33, 233–238. [Google Scholar] [CrossRef]
  49. Elkahwagy, N.; Ismail, A.; Maize, S.M.A.; Mahmoud, K.R. Diffusion Monte Carlo calculations on LaB molecule. Chin. Phys. B 2018, 27, 093102. [Google Scholar] [CrossRef]
  50. Ye, J.; Pang, H.F.; Wong, A.M.-Y.; Leung, J.W.-H.; Cheung, A.S.-C. Laser spectroscopy of iridium monoborides. J. Chem. Phys. 2008, 128, 154321. [Google Scholar] [CrossRef]
  51. McIntyre, N.S.; Auwera-Mahieu, A.V.; Drowart, J. Mass spectrometric determination of the dissociation energies of gaseous RuC, IrC and PtB. Trans. Faraday Soc. 1968, 64, 3006–3010. [Google Scholar] [CrossRef]
  52. Ng, Y.W.; Wong, Y.S.; Pang, H.F.; Cheung, A.S.C. Electronic transitions of platinum monoborides. J. Chem. Phys. 2012, 137, 124302. [Google Scholar] [CrossRef]
  53. Gingerich, K.A. Gaseous Metal Borides. III. The Dissociation Energy and Heat of Formation of Gold Monoboride. J. Chem. Phys. 1971, 54, 2646–2650. [Google Scholar] [CrossRef]
  54. Pang, H.F.; Ng, Y.W.; Xia, Y.; Cheung, A.S.C. Electronic transitions of iridium monoborides. Chem. Phys. Lett. 2011, 501, 257–262. [Google Scholar] [CrossRef]
  55. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  56. Yu, H.S.; He, X.; Li, S.L.; Truhlar, D.G. MN15: A Kohn-Sham Global-Hybrid Exchange-Correlation Density Functional with Broad Accuracy for Multi-Reference and Single-Reference Systems and Noncovalent Interactions. Chem. Sci. 2016, 7, 5032–5051. [Google Scholar] [CrossRef] [PubMed]
  57. Tao, J.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the Density Functional Ladder: Nonempirical Meta-Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401. [Google Scholar] [CrossRef] [PubMed]
  58. Peterson, K.A.; Figgen, D.; Dolg, M.; Stoll, H. Energy-consistent relativistic pseudopotentials and correlation consistent basis sets for the 4d elements Y-Pd. J. Chem. Phys. 2007, 126, 124101. [Google Scholar] [CrossRef]
  59. Peterson, K.A.; Puzzarini, C. Systematically convergent basis sets for transition metals. II. Pseudopotential-based correlation consistent basis sets for the group 11 (Cu, Ag, Au) and 12 (Zn, Cd, Hg) elements. Theor. Chem. Acc. 2005, 114, 283–296. [Google Scholar] [CrossRef]
  60. Figgen, D.; Peterson, K.A.; Dolg, M.; Stoll, H. Energy-consistent pseudopotentials and correlation consistent basis sets for the 5d elements Hf–Pt. J. Chem. Phys. 2009, 130, 164108. [Google Scholar] [CrossRef]
  61. Peterson, K.A.; Dunning, T.H. Accurate correlation consistent basis sets for molecular core–valence correlation effects: The second row atoms Al–Ar, and the first row atoms B–Ne revisited. J. Chem. Phys. 2002, 117, 10548–10560. [Google Scholar] [CrossRef]
  62. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  63. Chi, C.; Wang, J.-Q.; Hu, H.-S.; Zhang, Y.-Y.; Li, W.-L.; Meng, L.; Luo, M.; Zhou, M.; Li, J. Quadruple bonding between iron and boron in the BFe(CO)3 complex. Nat. Commun. 2019, 10, 4713. [Google Scholar] [CrossRef]
  64. Sevy, A.; Huffaker, R.F.; Morse, M.D. Bond Dissociation Energies of Tungsten Molecules: WC, WSi, WS, WSe, and WCl. J. Phys. Chem. A 2017, 121, 9446–9457. [Google Scholar] [CrossRef] [PubMed]
  65. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed]
  66. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  67. Werner, H.-J.; Knowles, P.J.; Manby, F.R.; Black, J.A.; Doll, K.; Heßelmann, A.; Kats, D.; Köhn, A.; Korona, T.; Kreplin, D.A.; et al. MOLPRO version 2022. Chem. Phys. 2020, 152, 144107. [Google Scholar]
Figure 1. (a) Bond lengths, re, (b) dipole moments, μ, (c) charge on metal, qM, via natural population analysis, (d) dissociation energies, De (eV), with respect to the adiabatic atomic products, (e) dissociation energies, D0,GS, with respect to the ground state atomic products, and (f) vibrational frequencies, ωe, of the ground states of the 2nd-row-transition-metal boride molecules, MBs (M = Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, and Cd), at the B3LYP, TPSSh, and MN15/aug-cc-pVQZ(-PP) level of theory.
Figure 1. (a) Bond lengths, re, (b) dipole moments, μ, (c) charge on metal, qM, via natural population analysis, (d) dissociation energies, De (eV), with respect to the adiabatic atomic products, (e) dissociation energies, D0,GS, with respect to the ground state atomic products, and (f) vibrational frequencies, ωe, of the ground states of the 2nd-row-transition-metal boride molecules, MBs (M = Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, and Cd), at the B3LYP, TPSSh, and MN15/aug-cc-pVQZ(-PP) level of theory.
Molecules 28 08016 g001
Figure 2. Molecular orbitals of the X3Σ (TcB), X2Δ (RuB), and X1Σ+ (RhB) states presenting quadruple bonds.
Figure 2. Molecular orbitals of the X3Σ (TcB), X2Δ (RuB), and X1Σ+ (RhB) states presenting quadruple bonds.
Molecules 28 08016 g002
Figure 3. (a) Bond lengths, re, (b) dipole moments, μ, (c) charge on metal, qM, via natural population analysis, (d) dissociation energies, De (eV) with respect to the adiabatic atomic products, (e) dissociation energies, D0,GS, with respect to the ground state atomic products, and (f) vibrational frequencies, ωe, of the ground states of the 3rd-row-transition-metal boride molecules, MBs (M = La, Hf, Ta, W, Re, Os, Ir, Pt, Au, and Hg) at the B3LYP, TPSSh, and MN15/aug-cc-pVQZB(-PP)M and def2-QZVPPDLa levels of theory.
Figure 3. (a) Bond lengths, re, (b) dipole moments, μ, (c) charge on metal, qM, via natural population analysis, (d) dissociation energies, De (eV) with respect to the adiabatic atomic products, (e) dissociation energies, D0,GS, with respect to the ground state atomic products, and (f) vibrational frequencies, ωe, of the ground states of the 3rd-row-transition-metal boride molecules, MBs (M = La, Hf, Ta, W, Re, Os, Ir, Pt, Au, and Hg) at the B3LYP, TPSSh, and MN15/aug-cc-pVQZB(-PP)M and def2-QZVPPDLa levels of theory.
Molecules 28 08016 g003
Figure 4. (a) Bond lengths, re, (b) dipole moments, μ, (c) charge on metal, qM, via natural population analysis, (d) dissociation energies, De (eV), with respect to the adiabatic atomic products, (e) dissociation energies, D0,GS, with respect to the ground state atomic products, and (f) vibrational frequencies, ωe, of the ground states of the 1st-, 2nd-, and 3rd-row-transition-metal boride molecules; 1st row: MRCISD+Q/aug-cc-pV5Z [21]; 2nd- and 3rd-row MBs: B3LYP/aug-cc-pVQZB(-PP)M and def2-QZVPPDLa levels of theory.
Figure 4. (a) Bond lengths, re, (b) dipole moments, μ, (c) charge on metal, qM, via natural population analysis, (d) dissociation energies, De (eV), with respect to the adiabatic atomic products, (e) dissociation energies, D0,GS, with respect to the ground state atomic products, and (f) vibrational frequencies, ωe, of the ground states of the 1st-, 2nd-, and 3rd-row-transition-metal boride molecules; 1st row: MRCISD+Q/aug-cc-pV5Z [21]; 2nd- and 3rd-row MBs: B3LYP/aug-cc-pVQZB(-PP)M and def2-QZVPPDLa levels of theory.
Molecules 28 08016 g004
Figure 5. (a) Dissociation energies, De (eV), per bond with respect to the adiabatic atomic products; (b) number of formed bonds of the ground states of the 1st-, 2nd-, and 3rd-row-transition-metal boride molecules, MBs.
Figure 5. (a) Dissociation energies, De (eV), per bond with respect to the adiabatic atomic products; (b) number of formed bonds of the ground states of the 1st-, 2nd-, and 3rd-row-transition-metal boride molecules, MBs.
Molecules 28 08016 g005
Table 1. Previous theoretical and experimental data on the ground states of the 1st-row-transition-metal boride molecules, MBs (M = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Zn): bond lengths re (Å), dissociation energies De (eV) and/or D0 (eV) with respect to the adiabatic products, vibrational frequencies ωe (cm−1), anharmonic corrections ωexe (cm−1), and dipole moments μ (μFF = δE/δε) (Debye).
Table 1. Previous theoretical and experimental data on the ground states of the 1st-row-transition-metal boride molecules, MBs (M = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Zn): bond lengths re (Å), dissociation energies De (eV) and/or D0 (eV) with respect to the adiabatic products, vibrational frequencies ωe (cm−1), anharmonic corrections ωexe (cm−1), and dipole moments μ (μFF = δE/δε) (Debye).
MBMethodologyRef.StatereDe aD0ωeωexeμ (μFF)
ScBicMRCISD/cc-pV5Z[21]X5Σ2.1283.2573.220584.53.84.02(4.16)
icMRCI+Q/cc-pV5Z[21] 2.1323.2873.2515793.8(4.23)
icMRCISD+DKH+Q/cc-pV5Z[21] 2.0943.3093.2716032.6(4.13)
B3LYP/6-311++G(3df)[22]X5Σ2.0841.90 a 612 3.95
MRCI+Q[23] 1.787 a
Pauling method[24] 2.8 a
R2PI spectroscopy[25] 1.72(6) a
TiBicMRCISD/cc-pV5Z[21]X6Δ2.0772.7232.687587.63.53.13(3.42)
icMRCI+Q/cc-pV5Z[21] 2.0802.7972.7615833.5(3.51)
B3LYP/6-311++G(3df)[22]X6Δ2.0392.402.362621 3.26
Pauling method[24] <3.1
R2PI spectroscopy[25] 1.956(16)
VBicMRCISD/cc-pV5Z[21]X7Σ+2.0392.2682.231589.93.752.73(3.20)
icMRCI+Q/cc-pV5Z[21] 2.0432.3812.3445853.8(3.30)
B3LYP/6-311++G(3df)[22]X7Σ+2.0112.172.132 a615 2.82
R2PI spectroscopy[25] 2.150(16) a
CrBicMRCISD/cc-pV5Z[21]X6Σ+2.1661.141 42018.52.05
icMRCI+Q/cc-pV5Z[21]X6Σ+2.1831.353 405201.43
B3LYP/6-311++G(3df)[22]X6Σ+2.187 415 2.41
MnBicMRCISD/cc-pV5Z[21]X5Π2.1900.8460.821391.93.792.21(2.45)
icMRCI+Q/cc-pV5Z[21] 2.1830.8540.8313833.9(2.39)
B3LYP/6-311++G(3df)[22] 1.8281.271.232621 2.47
FeBR2PI spectroscopy[26]GS 2.43(2)
B3LYP/aug-6-311++G(3df)[22]Χ4Σ1.6952.22 743 2.27
icMRCISD/cc-pV5Z[21]Χ4Σ1.7432.303 642.712.501.67(2.13)
icMRCI+Q/cc-pV5Z[21]Χ4Σ1.7472.346 64513.20(2.20)
CoBR2PI spectroscopy[26]GS 2.954(3)
LIF spectroscopy[27]Χ3Δ31.705
B3LYP/aug-6-311++G(3df)[22]Χ3Δ1.6762.62 783 1.89
icMRCISD/cc-pV5Z[21]Χ3Δ1.6962.736 756.96.141.03(1.83)
icMRCI+Q/cc-pV5Z[21]Χ3Δ1.7002.849 7576.10(1.98)
NiBR2PI spectroscopy[26]GS 3.431(4)
LIF spectroscopy[28]Χ2Σ+1.698 7784.90
B3LYP/aug-6-311++G(3df)[22]Χ2Σ+1.6762.83 793 1.66
icMRCISD/cc-pV5Z[21]Χ2Σ+1.6763.239 805.13.930.80(2.16)
icMRCI+Q/cc-pV5Z[21]Χ2Σ+1.6813.434 8033.98(2.41)
CuBicMRCISD/cc-pV5Z[21]Χ1Σ+1.9341.839 496.95.541.21(1.71)
icMRCI+Q/cc-pV5Z[21]Χ1Σ+1.9222.129 5534.80(1.62)
Nonrelativistic CASPT2/PolMe[29]Χ1Σ+1.9102.806 5184.25
No-pair DK CASPT2/NpPolMe[29]Χ1Σ+1.8652.354 5634.34
No-pair DK CCSD(T)-20/NpPolMe[29]Χ1Σ+1.9091.522 555.04.33
R2PI spectroscopy[30]Χ1Σ+ 2.26(15)
B3LYP/aug-6-311++G(3df)[22]Χ1Σ+1.9522.12 513 1.61
ZnBB3LYP/aug-6-311++G(3df)[22]Χ2Π2.2740.370 286 1.70
B3LYP/aug-cc-pVQZ(-PP)bΧ2Π2.2580.373 0.329 282.5 1.65
TPSSh/aug-cc-pVQZ(-PP)bΧ2Π2.2170.573 0.529 322.7 1.84
MN15/aug-cc-pVQZ(-PP)bΧ2Π2.3300.374 0.317 234.1 1.32
a Dissociation energy with respect to the atomic ground state products. b This work.
Table 2. Previous theoretical and experimental data on the ground states of the 2nd-row-transition-metal boride molecules, MBs (M = Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, and Cd): bond lengths re (Å), dissociation energies De (eV) and/or D0 (eV), vibrational frequencies ωe (cm−1), anharmonic corrections ωexe (cm−1), and dipole moments μ (μFF = δE/δε) (Debye).
Table 2. Previous theoretical and experimental data on the ground states of the 2nd-row-transition-metal boride molecules, MBs (M = Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, and Cd): bond lengths re (Å), dissociation energies De (eV) and/or D0 (eV), vibrational frequencies ωe (cm−1), anharmonic corrections ωexe (cm−1), and dipole moments μ (μFF = δE/δε) (Debye).
MBMethodologyRef.StatereDe aD0ωeωexeμ (μFF)
YBR2PI spectroscopy[25] 2.057(3)
DFT: B97-1/AVTZ-PPY/VTZB[25]X5Σ2.306(1.99)1.96517
DFT: B3LYP/LANL2DZ[34]S = 22.2542.17 582.4 4.65
Pauling method[24] 2.99
ZrBR2PI spectroscopy[25] 2.573(5)
DFT: B97-1/AVTZ-PPZr/VTZB[25]X6Δ2.159(2.65)2.61
DFT: B3LYP/LANL2DZ[34]S = 2.52.1893.92 610.2 3.48
NbBR2PI spectroscopy[25] 2.989(12)
DFT: B97-1/AVTZ-PPNb/VTZB[25]5Π/5Φ1.988(3.11)3.07698
DFT: B3LYP/LANL2DZ[34]S = 11.9963.40 662.7 3.84
MRCISD+Q/aug-cc-pVQZ(-PP)bX5Π2.0182.901 708.9 3.10
MRCISD+Q/aug-cc-pVQZ(-PP)bA5Φ2.0192.808 710.3 2.92
MoBCASPT2/CASSCF/ANO-RCC-4ζ[35]X6Π1.9682.18 664 2.7
DFT: B3LYP/LANL2DZ[34]S = 0.51.8176.40 826 4.05
TcBDFT: B97-1/(A)TcVTZB-(PP)Tc[25]X5Σ 3.31
RuBR2PI spectroscopy[26] 4.815(3)
Knudsen effusion[36]X2Σ1.75 4.59(22)
LIF spectroscopy[37]2Δ5/21.7099
DFT: B3LYP/LANL2DZ[34]S = 0.51.7616.48 910.8 3.49
RhBR2PI spectroscopy[26] 5.252(3)
Knudsen effusion[36]X1Σ1.75 4.89(22)915
LIF spectroscopy[38]X1Σ+1.691(2)
MS-CASPT2/ANO-RCC-4ζ[39]X1Σ+1.698 4.42
LIF spectroscopy[40]X1Σ+1.691(2)
MS-CASPT2/ANO-RCC-4ζ[41]X1Σ+1.694(5.7)5.6924 4.54
MRCISD+Q/AV5Z-PPRh AV5ZB [4]X1Σ+1.68735.4735.414938.34.32(3.160)
RCCSD(T)/AV5Z-PPRh AV5ZB [4]X1Σ+1.68725.4905.431942.13.78(2.865)
ADF/PBE/TZ2P[2]X1Σ+ 5.27
DFT: TPSSh/AVQZ-PPRh AVQZB [2]X1Σ+1.685
CCSD(T)/AVQZ-PPRh AVQZB[2]X1Σ+1.689
MRCI/AVQZ-PPRh AVQZB[2]X1Σ+1.687
MCSCF/Sapporo-(DKH)Rh-TZP[42]X1Σ+1.7015.165
DFT: B3LYP/LANL2DZ[34]S = 01.7454.96 932.7 2.84
PdBESR spectroscopy[43]X2Σ
UHF/STO-3G*[43] 1.608
LIF spectroscopy[44]X2Σ+1.7278 650
Knudsen effusion[36]X2Σ2.00 3.37(22)
DFT: B3LYP/LANL2DZ[34]S = 0.51.8563.33 725.6 1.44
AgBNonrelativistic CASPT2/PolMe[29]X1Σ+2.2581.2481.233412.32?
No-pair DK CASPT2/NpPolMe[29]X1Σ+2.0981.6841.664253.41?
No-pair DK CCSD(T)-20/NpPolMe[29]X1Σ+2.1150.9100.8834403.26?
DFT: B3LYP/LANL2DZ[34]S = 02.1871.60 415.6 1.41
CdBDFT: B3LYP/LANL2DZ[34]S = 0.52.6680.22 198.3 1.67
a Dissociation energy with respect to the atomic ground state products. b This work.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Demetriou, C.; Tzeliou, C.E.; Androutsopoulos, A.; Tzeli, D. Electronic Structure and Chemical Bonding of the First-, Second-, and Third-Row-Transition-Metal Monoborides: The Formation of Quadruple Bonds in RhB, RuB, and TcB. Molecules 2023, 28, 8016. https://doi.org/10.3390/molecules28248016

AMA Style

Demetriou C, Tzeliou CE, Androutsopoulos A, Tzeli D. Electronic Structure and Chemical Bonding of the First-, Second-, and Third-Row-Transition-Metal Monoborides: The Formation of Quadruple Bonds in RhB, RuB, and TcB. Molecules. 2023; 28(24):8016. https://doi.org/10.3390/molecules28248016

Chicago/Turabian Style

Demetriou, Constantinos, Christina Eleftheria Tzeliou, Alexandros Androutsopoulos, and Demeter Tzeli. 2023. "Electronic Structure and Chemical Bonding of the First-, Second-, and Third-Row-Transition-Metal Monoborides: The Formation of Quadruple Bonds in RhB, RuB, and TcB" Molecules 28, no. 24: 8016. https://doi.org/10.3390/molecules28248016

Article Metrics

Back to TopTop