Next Article in Journal
Molecular Distillation of Lavender Supercritical Extracts: Physicochemical and Antimicrobial Characterization of Feedstocks and Assessment of Distillates Enriched with Oxygenated Fragrance Components
Next Article in Special Issue
Electrochemistry of Cobalta Bis(dicarbollide) Ions Substituted at Carbon Atoms with Hydrophilic Alkylhydroxy and Carboxy Groups
Previous Article in Journal
Nitrification Progress of Nitrogen-Rich Heterocyclic Energetic Compounds: A Review
Previous Article in Special Issue
Vibrational Property of α-Borophene Determined by Tip-Enhanced Raman Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Relative Stability of Boron Planar Clusters in Diatomic Molecular Model

by
Levan Chkhartishvili
1,2
1
Department of Engineering Physics, Georgian Technical University, 77 Merab Kostava Avenue, Tbilisi 0160, Georgia
2
Boron and Powder Composite Materials Laboratory, Ferdinand Tavadze Metallurgy and Materials Science Institute, 8b Elizbar Mindeli Street, Tbilisi 0186, Georgia
Molecules 2022, 27(5), 1469; https://doi.org/10.3390/molecules27051469
Submission received: 27 December 2021 / Revised: 10 February 2022 / Accepted: 12 February 2022 / Published: 22 February 2022
(This article belongs to the Special Issue New Science of Boron Allotropes, Compounds, and Nanomaterials)

Abstract

:
In the recently introduced phenomenological diatomic molecular model imagining the clusters as certain constructions of pair interatomic chemical bonds, there are estimated specific (per atom) binding energies of small all-boron planar clusters Bn, n = 1–15, in neutral single-anionic and single-cationic charge states. The theoretically obtained hierarchy of their relative stability/formation probability correlates not only with results of previous calculations, but also with available experimental mass-spectra of boron planar clusters generated in process of evaporation/ablation of boron-rich materials. Some overestimation in binding energies that are characteristic of the diatomic approach could be related to differences in approximations made during previous calculations, as well as measurement errors of these energies. According to the diatomic molecular model, equilibrium binding energies per B atom and B–B bond lengths are expected within ranges 0.37–6.26 eV and 1.58–1.65 Å, respectively.

1. Introduction

Nanoboron and boron-rich nanomaterials are of current academic and practical interests because of their widely variable interatomic bonding mechanism and related unique complex of physical–chemical properties useful in technological applications—see some of the reviews in the last decade [1,2,3,4,5,6,7,8,9].
Among them, the all-boron clusters Bn, n = 2, 3, 4, …, as individual species in the gas phase, play an important role, as they can serve for building blocks in the boron-rich solids chemistry [10]. For example, a quasi-planar boron cluster B35 with a double-hexagonal hole at the center has been reported [11] as a flexible structural motif for borophene allotropies, as it can be used to construct atom-thin boron sheets with various hexagonal hole densities.
In this regard, it should be noted that, depending on the number of atoms and also formation kinetics, boron clusters can take several different shapes. Joint experimental studies and computational simulations revealed [12] that boron clusters, which favor (quasi)planar, i.e., 2D, structures up to 18 atoms, prefer 3D structures beginning at 20 atoms. The B20 neutral cluster was found to have a double-ring tubular ground structure. As for the B20 anion, its tubular structure was shown to be almost isoenergetic to 2D structures. Thus, the usually observed 2D-to-3D structural transition suggests that B20 may be considered as the embryo of thinnest single-walled boron nanotubes. According to the QC (Quantum Chemical) and DFT (Density Functional Theory) investigations [13], there are two structural transitions that are expected in boron clusters: the second transition from double-ring system into triple-ring one occurs between B52 and B54.
At low (namely, from 7 to 20) nuclearities, i.e., for (quasi)planar boron clusters, the separate quantum rules of in- and out-of-plane bonding were obtained [14], using the free-particle-on-disk and rectangle models combined with DFT electronic structure calculations.
In this Introduction, the quite-rich data available on boron clusters synthesis methods, binding parameters and their potential applications are only briefly discussed.

1.1. Synthesizing

Boron quasi-planar clusters can be formed in process of thermal vaporization [15,16,17,18,19,20,21], bombardment with high-energy particles [22], even grinding of boron-rich solid materials [23,24,25] and mainly by their laser ablation [26,27,28,29].
An effective method of thermal generating of pure boron cluster-ions for their further use as a plasma-process feed gas was proposed by Becker [18]. Chamber’s electrode material is a compound of boron with Me metal(s) thermally decomposable within a suitable temperature range to provide boron in the vapor, but other species are substantially not in the vapor states. Magnetic confinement of the simultaneously released electrons causes numerous collisions, resulting in boron vapor ionization to the plasma state. This plasma is then extracted and accelerated at a suitable energy toward the workpiece. Created in this way, boron clusters can be self-assembled into nanostructures [19].
As early as in References [23,24], electron microscopic study of the elementary boron powder structure revealed that ultrafine particles (of ≤200 Å size) of freely grown boron are characterized by a stable 2D shape with almost hexagonal profile and aspect ratio of ~20:1. Computer simulations were performed [25] to model structural relaxation in 2D-clusters mimicking these boron small particles.
Theoretically, using DFT, a growth path for small boron clusters Bn was discussed [30] with a size range and isomers structure. The thermochemical parameters that were determined by using coupled-cluster theory calculations suggested [31] the evolution of geometry and resonance energy of Bn clusters through the number of B–B bonds. Based on the known geometrical characteristics of boron clusters, their general growth mechanism was proposed in Reference [32]. Moreover, a systematic structural investigation of Bn clusters established a picture of their growth behavior [33].

1.2. Structure and Binding

According to References [15,34], the diboron molecule B2 experimental dissociation energy is within range of 2.82 ± 0.24 or 2.69 ± 0.43 eV, respectively. The absorption transition at 3200–3300 Å observed in B2 indicated [35] that its ground electronic state should be the lower state of Σ-type. Moreover, as B2 was not observed via ESR (Electron Spin Resonance), the ground electronic state was identified with 3Σg. An accurate CI (Configuration Interaction) calculation confirmed [36] that mentioned transition is from the first excited state of 3Σu type. Molecular binding energy for B2 at the HF (Hartree–Fock) level of theory [37] is 2.861 eV, and its scaled ground-state harmonic frequency equals to 0.120 eV. From the DFT calculations [30], these parameters are 2.718 and 0.134 eV, respectively. By using the MO (Molecular Orbitals) method, the dissociation energy of B2 ground state was calculated to be 2.71 eV [38]. Handbook [39] recommends the diboron molecule ground electron state dissociation energy of 3.02 eV. In this state, its bond length and vibration quantum are 1.590 Å and 0.130 eV, and in the first excited state, they are 1.625 Å and 0.116 eV, respectively. The term of corresponding transition equals to 3.79 eV.
Potential energy curves for the states of B2 were constructed by the complete-active-space SCF (Self Consistent Field) method at the multi-reference CI level [36]. According to other multi-reference CI study [40], its ground-state curve parameters—bond length, dissociation energy and relative vibration quantum are of 1.600–1.607 Å, 2.70–2.78 and 0.128–0.129 eV, respectively. For the B2 molecule ground-state interatomic potential energy, P curve P d , where d is the inter-nuclear distance, constructed [41,42] within quasi-classical approach (Figure 1), the curve’s parameters are as follows: dissociation energy of 2.80 eV, equilibrium bond length of 1.78 Å and vibration quantum of 0.13 eV. Based on explicit expressions for intersite distances in boron nanotubes of regular geometry in terms of B−B bonds length and using this quasi-classical B−B pair potential, there were estimated some ground-state parameters of boron nanotubes [43,44,45].
The potential curves, transition energies, bond lengths and vibration frequencies of ground and some of low-lying excited electronic states of B2+ and B2 were obtained by using a CI approach [46]. The B2+ cation ground state, which was found to be 2Σg+, shows a rather shallow potential curve with a bond length of 2.125 Å and vibration quantum of 0.052 eV, when compared with that of the 3Σg state of B2 neutral: 1.592 Å and 0.131 eV. The first excited state of B2+, 2Πu, lies at 0.30 eV. As a result of loss of bonding electron, the ground-state dissociation energy for B2+ with calculated value of 1.94 eV is smaller than that of B2.
In Reference [47], the local nature of different types of boron–boron bonds from the topological analysis of ELF (Electron Localization Function) perspective was investigated in 23 boron-containing molecules.
Initially, Boustani demonstrated that stable structures of neutral bare boron clusters, Bn, with n = 12, 16, 22, 32, 42 and 46, can easily be constructed with the help of the so-called Aufbau Principle suggested on the basis of HF SCF direct CI [48], as well as QC, DFT and LMTO (Linear Muffin Tin Orbital) [49,50] studies.
Based on different theoretical approaches, such as LDA (Local Density Approximation) and LSD (Local Spin Density) versions of DFT, HF, complete active space and scattered-wave SCF; correlated CI, QC, MO, PES (Potential Energy Surface) coupled-cluster, Born–Oppenheimer and full-potential LMTO versions of MD (Molecular Dynamics); and other methods, there are reported the key structural and binding parameters of boron clusters calculated or scaled from measured ones: B3 [37,51]; B3 and B4 in neutral and anionic forms [52]; B4 [53]; B2, B3 and B4 [54]; B5, B5+ and B5 isomers [55]; B5 [56]; B6 and B6 [57]; B7 and B7 [58]; B2+–B8+ [59]; Bn with n = 2–8 in both the neutral and cationic states [60]; Bn clusters with 4 ≤ n ≤ 8 [61]; 8- and 9-atom boron clusters [62]; small boron clusters with up to 10 atoms [63]; B2–12 and B2–12+ [64]; small neutral Bn clusters with n = 2–12 [30]; icosahedral cluster B12 [65]; B12 [66,67]; B2+–B13+ [68] (see also Reference [69]); multi-charged clusters Bn with n = 2–13 [70]; set of small-sized neutral Bn and anionic Bn boron clusters with n = 5–13 [31]; B7, B10 and B13 [71]; B12 and B13 for neutrals and cations [72]; B12+ and B13+ [73]; B12 and B13+ [74]; isomers of boron 13-clusters [75]; cationic, neutral and anionic charge states of B13 [76]; planar or quasi-planar structures of B13, B13+ and B13 [77]; isomers of planar boron cluster B13 [78]; small cationic clusters Bn+ with n = 2–14 [79]; Bn with n = 2–14 [80,81]; Bn clusters for n ≤ 14 [82]; Bn with n = 5–14 [83]; boron clusters in the 10- to 15-atom size range [84]; B16 and B162– [85]; B19 [86]; neutral and anionic B20 isomers [87]; B23+ [88]; isomers of 24-atom boron cluster [89]; B25 [90]; B27 [91]; Bn with n = 26–29 in both neutral and anionic states [32]; B32 [92]; B36 and B36 [93,94]; B41 and B42 [95]; neutral Bn clusters with n = 31–50 [33]; and boron cluster-families: spheres, double-rings and quasi-planars containing up to n = 122 atoms [96].

1.3. Applications

The unique behaviors of clusters of elemental boron have been identified: they react readily with metal surfaces; bond covalently to metal atoms; and cover surfaces with a boron-enriched hard, smooth and corrosion-resistant layer, which can be called nano-boron coating. Some parameters, especially feeding gas concentration, substrate temperature and input power, were optimized [97] to prepare high-pure boron (94%B) coating films by plasma-assisted CVD (Chemical Vapor Deposition). However, due to its thermal resilience, elemental boron is a difficult material to work with. To overcome this problem, Becker perfected a technique for generating clusters of elemental boron in plasma [98]. Unique tribological applications and scalability of boron-rich materials are likely to emerge from the combination of high mechanical strength, chemical stability, exceptional hardness and toughness, wear resistance, strong binding to substrates, low density and other promising physical–chemical properties [99].
The durability of boron coatings in sliding friction has been mentioned [100]. In the search ways to enhance the surface hardness of aluminum, the equilibrium structure, stability, elastic properties and formation dynamics of a boron-enriched surface were studied by using DFT [101]. Nano-indentation simulations suggested that the presence of boron nanostructures in the subsurface region significantly enhances the mechanical hardness of aluminum surfaces.
Boron is the only practical solid material with both volumetric and gravimetric energy densities substantially greater than those of hydrocarbons. That is why boron powder is attractive as a fuel or a supplement in propellants and explosives and potential source of secondary energy generation as well. One study [102] aimed to obtain the energy from elemental boron burning as solid fuel, which is synthesized from boron minerals. By experimental investigation [103] of the combustion characteristics of boron nanoparticles in the post-flame region, a two-stage combustion phenomenon was observed. The extended combustion model for single boron particles of sizes relevant for ramjet chambers was introduced [104] and validated [105]; it comprised a consistent formulation of the heat and mass transfer processes in the boron particles’ environment. A review on boron powders given in Reference [106] serves as the basis for research on the tendency of nanosized boron particles to group in an oscillating flow and its effect on the combustion process, flame characteristics and pollutants’ emission. Nanoboron can be considered as a superior rocket fuel because nano-particles have almost fluid-like properties. To optimize the reactive surface for combustion, nanoparticle size could be shrunk to clusters consisting of several atoms each. One review paper [107] encompassed the status and challenges in the synthesis process of boron nanoparticles, their dispersion and stability of in liquid hydrocarbon fuels, ignition and combustion characteristics of boron loaded liquid fuel, particle combustion and characterization of post-combustion products. The combustion characteristics of nanofluid fuels containing additions of boron and iron particles together were investigated in Reference [108]. Furthermore, mechanical milling was used to prepare a boron-based composite powder containing 5wt.% nanoiron to behave as a catalyst of boron oxidation [109]. The energy density of reactive metal fuel containing Ti, Al and B nanopowders was optimized [110] by varying the Ti:Al:B ratio.
As for information about oxidation of boron clusters, the cross-sections for ionic products formed in reactions of B1–13+ with oxygen were measured under single collision conditions [111] and three main reaction mechanisms found to be important: oxidative fragmentation, collision induced dissociation and boron atom abstraction. Cross-sections for oxidation reactions of CO2 with boron cluster ions B1–14+ were reported as a function of collision energy [112]. At least in some cases, oxidation causes structural rearrangement of the boron clusters. To rationally design and explore a future energy source based on the highly exothermic oxidation of boron, DFT was used to characterize small boron clusters with 0–3 oxygen atoms and, in total, up to 10 atoms [63].
Thin coatings made from the 10B isotopically enriched nanoboron providing the highest possible concentration of neutron capture centers can greatly simplify the problem of protection against thermal neutron irradiation [113,114,115], as well as neutron-fluence nanosensors [116,117]. Boron 2D metallic crystal is a prospective electromagnetic shielding nanomaterial as well [118], so nanoboron can combine neutron and electromagnetic shielding properties.
Reference [119] presented the concept that an elongated planar boron cluster can serve as a “tank tread” at the sub-nanometer scale, a novel propulsion system for potential nanomachines. Ferromagnetism in all-boron planar clusters, e.g., in B34, has been revealed theoretically [120]. They can be assembled to construct all-boron ferromagnetic monolayers, in which ferromagnetism–paramagnetism and semiconductor–metal transitions are expected to occur around 500 K, indicating their potential applications in nanoelectronic and spintronic devices at room temperature.
The recent review [121] on borophene potential applications discusses in detail its other utilizations, such as alkali metal ion and Li–S batteries, hydrogen storage, supercapacitors, sensor and catalytic in hydrogen evolution, oxygen reduction and evolution and CO2 electroreduction reaction.
The hierarchy of clusters relative stability/formation probability mainly (together with peculiarities characteristic of formation kinetics) is determined by the cluster-specific binding energy—its binding energy per atom. This work is focused on theoretical estimation of this key energy parameter for boron small planar clusters in frames of recently modified phenomenological diatomic molecular model.

2. Methodology

Specific binding energy or binding energy per chemical formula unit of the substance clustered form serves for important factor determining relative stabilities and, consequently, affects the relative concentrations of clusters with different numbers of formula units synthesized during a formation process (of course, the mentioned concentrations are influenced by the process kinetics as well). Here, we intend to calculate specific binding energies for boron small clusters in (quasi)planar structures, starting from the old diatomic model [122] of bounded multi-atomic structures.
The diatomic model is based on the saturation property of interatomic bonding. In its initial approximation, when binding energy is the sum of energies of pair interactions between only neighboring in the structure atoms, the microscopic theory of expansion allows for the quite correct estimation of the thermal expansion coefficient for crystals [123]. Despite its simplicity, the diatomic model has been successfully used to calculate some other anharmonic effects in solids as well [124].
As for the clusters binding energy, to the best our knowledge, there are no reports on its calculations within diatomic model, unless our previous estimates of small all-boron (quasi)planar clusters relative stabilities [125,126,127,128] and also their dipole moments [129,130]. These results have been summarized in the mini-reviews [131,132]. Furthermore, the problem was specially analyzed [133] for the three most abundant clusters, namely B11, B12 and B13, in different charge states, while taking into account ionization processes.
As is mentioned above, boron (quasi)planar clusters are of special interest, as they can form a borophene–monatomic boron layer with unique properties. The B–B bond length value obtained for boron finite planar clusters in quasi-classical approximation was used for an input parameter in quasi-classical calculations of electron energy band surfaces and DoS (Density-of-State) for a flat boron sheet with a perfect (i.e., without any type of holes) triangular network [134]. It is expected to have metallic properties. The Fermi curve of the boron flat sheet is found consist of 6 parts of 3 closed curves well-approximated by ellipses and then representing the quadric energy dispersion of the conduction electrons. The effective mass of electrons at the Fermi level is found to be too small compared with the free electron mass and highly anisotropic. The low effective mass of conduction electrons indicates their high mobility and, hence, high conductivity of the boron sheet.
Recently, we introduced [135] the most general formulation of the diatomic model allowing analytical calculation of the clusters binding energy. It is based on the following assumptions:
Cluster binding energy is the sum of energies of pair interactions between nearest neighboring in its structure atoms;
Assuming that relative deviations of bond lengths in the multi-atomic cluster structure from their values in corresponding diatomic molecules are small, pair interaction energies between neighboring atoms are approximated by their quadratic functions;
Valence-electron-density-redistribution-related corrections to the bond energies in the cluster can be expressed through effective static charges localized on pairs of nearest neighboring atoms;
Interatomic vibrations related corrections can be approximated by ground-state vibrational energies of corresponding diatomic molecules.
In the simplest but of practical interest special case, when all the bond lengths can be assumed to be almost equal each to other, a , cluster symmetry does not lead to any constrain (a relation to be satisfied by bond lengths), and most of other characteristics of valence bonding are equal each to other as well: a 0 is the bond length corresponding to diatomic molecule, E 0 is the diatomic molecule binding/dissociation energy, ω is the cyclic frequency of relative interatomic vibrations and M is the reduced mass of the pair of atoms with masses μ 1 and μ 2 constituting the bond:
1 M = 1 μ 1 + 1 μ 2
In case of identical atoms, we have the following:
μ 1 = μ 2 μ
and
M = μ 2
However, effective atomic charge numbers Z i 1 and Z i 2 characterizing electrostatic correction to the valence bonding energy remain different. The point is that the static charges localized on the pair of nearest-neighboring atoms differ for pairs placed at the center and periphery of the cluster, which is a finite structure of atoms. Here, the i index is as follows:
i = 1 , 2 , 3 , , k
where the numbers are different types of chemical bonds presented in the cluster, and k denotes their total number. If N i is the number of bonds of i -type, then we have the following:
N = i = 1 i = k N i
which is the number of bonds in whole the cluster.
When we introduce the following parameter,
Z = i = 1 i = k N i Z i 1 Z i 2
equilibrium bonds length, a , and corresponding cluster binding energy, E (in Gauss units), are as follows:
a = a 0 + 2 e 2 Z μ ω 2 a 0 2 N
and
E = ( E 0 ω 2 ) N e 2 Z a 0 + e 4 Z 2 μ ω 2 a 0 4 N
respectively.
From the data available [39] for diboron molecule, a 0 ≈ 1.590 Å, E 0 ≈ 3.02 eV and ω ≈ 1051.3 cm−1 and boron atomic mass weighted for stable isotopes (10B and 11B) natural abundance μ ≈ 10.811 amu, we get the following formula for numerical calculations:
E [ eV ] 2.995   N 9.053   Z + 0.7273   Z 2 N
The specific binding energy for cluster of n atoms is calculated as E / n .
To know parameter Z , one needs the estimates of effective atomic charges in the cluster. Below, they are found based on the assumption that the effective number of outer valence shell electrons localized on a given atomic site is proportional to its coordination number. Every neutral boron atom contains only 1 electron in the outer valence shell (2p-state). Then their total number in the all-boron cluster Bn is given as follows:
ν = { n        B n 0 n 1    B n + n + 1    B n
Moreover, if C j denotes the coordination number of j -site, then we have the following:
Z j 1 ν C j l = 1 l = n C l
The numerical values of these effective charge numbers and a summary of them for parameter Z for the ground-state structures of boron planar clusters with n = 2 15 atoms in three charge states are shown below in Table 1.

3. Results and Discussion

Here, we report the specific binding energy that we calculated for boron small planar clusters only in their ground-state structures. Ground-state structural isomers were chosen based on two criteria leading to maximal binding energy: (1) maximum number of bonds and (2) highest symmetry. We calculated clusters containing up to 15 boron atoms, because, for bigger species, 2D (quasi)planar structures are challenged by 3D ring-like structures (actually fragments of nanotubes), which are characteristic for boron clusters at n 20 .
Specific binding energies calculated in diatomic model for neutral Bn0, cationic Bn+ and anionic Bn clusters with the number of atom n = 2 15 are listed in Table 1 and presented in Figure 2.
For comparison, available experimental data are presented in Figure 3. In Reference [130], clusters in boron vapor were generated by the thermal decomposition of electrode made of a boron-rich metal boride. Note that, in this way, initially clusters are formed in neutral state but then ionized by collisions with released in the chamber energetic electrons and accelerated toward the mass-spectrometer. Boron cluster cations were generated [29] by laser evaporation of target compact made of boron dust with gold added to increase its stability. Here, a disproportionately intense peak related to traces of gold is erased because it masks that of B+18, as masses of gold Au atom and boron 18-atom cluster are almost undistinguishable. As for the boron cluster anions, they were produced [28] by laser vaporization from homogeneous pure boron target rods.
Again, for comparison, Figure 4 represents the theoretical specific binding energies of small boron clusters (in different, not only in planar, structures) depending on their size calculated by using QC methods [70]. One can note that such a typical curve looks similar to curves obtained in frames of diatomic model.
Thus, from the diatomic model, the theoretical equilibrium binding energies per B atom and B–B bond lengths are expected within ranges of 0.37–6.26 eV and 1.58–1.65 Å, respectively. For the most stabile neutral, positively and negatively charged species, B130, B12+ and B11 are predicted; their characteristics are 6.26, 6.15 and 6.23 eV and 1.59, 1.59 and 1.58 Å, respectively.
From experimental reports, the diboron molecule B2 dissociation energy is expected within the ranges of 2.58–3.06 [15] and 2.26–3.12 eV [34]. Theoretical values obtained by HF [37], MO [38], CI [40], quasi-classical [41,42] and PES [53] methods are 2.86, 2.71, 2.70–2.78, 2.80 and 2.70 eV, respectively. As diboron molecule contains two atoms, the corresponding specific binding energies equal to (2.58–3.06)/2 = 1.29–1.53 and (2.26–3.12)/2 = 1.13–1.56 for measured and 2.86/2 = 1.43, 2.71/2 ≈ 1.36, (2.70–2.78)/2 = 1.35–1.39, 2.80/2 = 1.40 and 2.70/2 = 1.35 eV for calculated dissociation energies. Calculations based on DFT [30] and QC [70] yielded 1.36 and 1.39 eV, respectively. The value of 1.48 eV obtained for cluster B20 from the diatomic model falls in both experimental ranges and seems only slightly overestimated if compared with previous theoretical ones.
The ground-state dissociation energy of diboron cation B2+ calculated with the CI approach [46] is 1.94 eV, which corresponds to 1.94/2 = 0.97 eV for specific binding energy, significantly exceeding 0.37 eV yielded by diatomic model for B2+. The same is true for the QC specific binding energy for B2+: 1.16 eV [70]. This discrepancy should be related not only to the diatomic model itself but mainly phenomenological estimation of static atomic charges used. The point is that, when calculating static atomic charges in diatomic species, the equal dividing of a single elemental charge between constituent atoms is too crude of an approximation, leading to the overestimated Coulomb repulsive energy. SCF CI [69] and calibrated hybrid DFT [82] approaches’ B2+→B1+ + B10 fragmentation energies are 1.47 and 1.96 eV, respectively. The corresponding specific binding energies are 1.47/2 ≈ 0.74 and 1.96/2 = 0.98 eV.
The QC specific binding energy of anionic cluster B2 is 2.24 eV [70]. As for the cationic cluster, this value significantly exceeds 0.37 eV, which is suggested by the diatomic model. The reason should be the same as for cationic isomer.
For the B3 cluster, the HF scaled dissociation energy is 8.59 eV [37], while the PES study showed that atomization energy for B3 would be in the range of 8.21–8.36 eV [53]. As the molecule contains three atoms, the corresponding specific binding energies are equal to 8.59/3 ≈ 2.86 and (8.21–8.36)/3 ≈ 2.74–2.79 eV, respectively. Calculations based on DFT [30] and QC [70] approaches yielded 2.76 and 2.82 eV, respectively. The diatomic model’svalue of 2.96 eV for cluster B30 again seems to be slightly overestimated if compared with other theoretical results.
The QC specific binding energy of B3+ and B3 clusters is 2.46 and 3.58 eV, respectively [70], while the diatomic model gives 1.96 eV for both clusterions.
A SCF calculation [51] predicted that the energy of fragmentation of the B3 cluster to produce diatomic B2 and atomic B is 4.96 eV. From the diatomic model, one obtains 3 × 2.96 – 2 × 1.48 = 5.92 eV.
The SCF CI B3+→B1+ + B20 and B3+→B2+ + B10 fragmentation energies are 1.45 and 2.00 eV, respectively [68], while the B3+→B1+ + B20 fragmentation energy calibrated in a hybrid DFT approach is 4.33 eV [82]. The diatomic model gives 3 × 1.96 − 2 × 0.37 = 5.14 and 3 × 1.96 − 2 × 1.48 = 2.92 eV for B3+→B1+ + B20 and B3+→B2+ + B10 reactions, respectively.
According to the study of the B4 cluster PES, its total atomization energy is expected in the range of 13.46–13.64 eV [53]. As this cluster consists of four atoms, the corresponding specific binding energy is (13.46–13.64)/4 ≈ 3.37–3.41 eV. The calculations based on DFT [30] and QC [70] approaches yielded 3.37 and 3.45 eV, respectively. The value of 3.97 eV obtained in the diatomic model for the neutral B40 cluster is higher but comparable. However, the CI analysis of B4 tetramer in rhombus geometry significantly underestimated its specific binding energy: 2.42 eV [54].
The QC specific binding energies of B4+ and B4 clusters are 3.28 and 3.76 eV, respectively [70]. The diatomic model gives quite close results of 3.31 and 3.13 eV for these cluster-ions.
SCF CI B4+→B1+ + B30, B4+→B3+ + B10 and B4+→B2+ + B20 fragmentation energies are 1.75, 2.80 and 3.20 eV, respectively [68]. Moreover, B4+→B1+ + B30 fragmentation energy calibrated in a hybrid DFT approach equals to 4.23 eV [82]. In the diatomic approach, their values are 4 × 3.31 − 3 × 2.96 = 4.36, 4 × 3.31 − 3 × 1.96 = 7.36 and 4 × 3.31 − 2 × 0.37 − 2 × 1.48 = 9.54, respectively.
The specific binding energies of the B50 cluster calculated in DFT [30] and two versions of QC approach [31,70] are 3.67, 3.68 and 3.76 eV, respectively. Our result is 4.54 eV.
The QC specific binding energy of B5+ is 3.82 eV [70], while we obtained 4.17 eV.
The specific binding energies of the B5 cluster calculated in two versions of the QC approach are 4.04 [31] and 4.11 eV [70], respectively, while the diatomic model yields the lower value of 3.77 eV.
The SCF CI B5+→B1+ + B40 fragmentation energy is 2.00 eV [68]. By calibrating this energy in a hybrid DFT approach, it was found 4.95 eV [82] to be in almost perfect agreement with the diatomic model result of 5 × 4.17 − 4 × 3.97 = 4.97 eV.
By using DFT [30], QC [31,70] and diatomic model (this work) methods, we calculated the specific binding energy for the B60 cluster as 3.79, 3.84, 3.88 and 4.94 eV, respectively. The QC approach [70] and diatomic model have yielded specific binding energy for B6+ cluster as 3.93 and 4.83 eV, respectively. According to two QC [31,70] and diatomic model calculations, the specific binding energy for B6 cluster is 4.16, 4.25 and 4.15 eV, respectively.
The SCF CI B6+→B5+ + B10 fragmentation energy is 2.20 eV [68], while, in a hybrid DFT approach, it is found to be 4.11 eV [82]. The diatomic value of 6 × 4.82 − 5 × 4.17 = 8.07 eV exceeds both of them.
DFT [30], QC [31,70], LMTO MD [71] and diatomic model specific binding energies for the B70 cluster are 4.07, 4.11, 4.17, 5.24 and 5.67 eV, respectively. For the B7+ cluster, the QC [70] and diatomic-model specific binding energies are 4.28 and 5.55 eV, respectively. Moreover, the QC [31,70] and diatomic-model specific binding energies for B7 cluster are 4.49, 4.48 and 5.07 eV, respectively.
The B7+→B6+ + B10 fragmentation energy calibrated in a hybrid DFT approach is 6.05 eV [82], i.e., lower than that predicted by the diatomic model: 7 × 5.55 − 6 × 4.82 = 9.93 eV.
For the B80 cluster, the DFT [30], QC [31,70] and diatomic-model specific binding energies are 4.31, 4.33, 4.41 and 5.50 eV, respectively. As for the B8+ cluster, its QC [70] and diatomic-model specific binding energies are 4.42 and 5.53 eV, respectively. Moreover, the B8 cluster’s QC [31,70] and diatomic-model specific binding energies equal to 4.71, 4.71 and 4.88 eV, respectively.
The B8+→B7+ + B10 fragmentation energy calibrated in a hybrid DFT approach is 5.04 eV [82], which is in good agreement with this work’s calculation: 8 × 5.53 − 7 × 5.55 = 5.39 eV.
The DFT [30], QC [70] and diatomic-model specific binding energies of B90 cluster are 4.30, 4.39 and 5.00 eV, respectively. The QC [70] and diatomic-model specific binding energies of B9+ cluster are 4.46 and 5.25 eV, respectively. The QC [31,70] and diatomic-model specific binding energies of B9 cluster are 4.55, 4.71 and 4.25 eV, respectively.
The B9+→B8+ + B10 fragmentation energy calibrated in a hybrid DFT approach is 4.29 eV [82]. The diatomic model yields the fragmentation energy of 9 × 5.25 − 8 × 5.53 = 3.01 eV, which is surprisingly less than previous theoretical results.
The DFT [30], QC [70], LMTO MD [71] and diatomic-model specific binding energies of B100 cluster are 4.45, 4.56, 5.78 and 5.96 eV, respectively. The QC [70] and diatomic-model specific binding energies of the B10+ cluster are 4.57 and 5.84 eV, respectively. The QC [70] and diatomic-model specific binding energies of the B10 cluster are 4.79 and 5.24 eV, respectively.
As for the B10+→B9+ + B10 fragmentation energy calibrated in a hybrid DFT approach, it equals 5.58 eV [82], while the diatomic model gives 10 × 5.84 − 9 × 5.25 = 11.15 eV.
The DFT [30], QC [70] and diatomic-model specific binding energies of the B110 cluster are 4.52, 4.63 and 5.63 eV, respectively. QC [70] and diatomic-model specific binding energies of B11+ cluster are 4.69 and 5.45 eV, respectively. The QC [70] and diatomic-model specific binding energies of B11 cluster are 4.89 and 6.23 eV, respectively.
The B11+→B10+ + B10 fragmentation energy calibrated in a hybrid DFT approach is 5.18 eV [82]. In this case, the diatomic model value is significantly less: 11 × 5.45 − 10 × 5.84 = 1.55 eV.
In Reference [67], the Jahn–Teller distortion mechanism, which transforms the highlysymmetric icosahedral structure B120 into a quasi-planar disc-like structure with binding energy per atom as 4.60 eV, was proposed. The same value was suggested based on DFT calculations: 4.60 eV [30]. The QC specific binding energy of B120 is 4.71 eV [70], while our diatomic model gives 5.77 eV. The QC [70] and diatomic-model specific binding energies of the B12+ and B12 charged cluster are 4.72 and 4.85, and 6.15 and 5.72 eV, respectively.
The B12+→B11+ + B10 fragmentation energy calibrated in a hybrid DFT approach is 4.64 eV [82], while 12 × 6.15 − 11 × 5.45 = 13.85 eV is obtained based on diatomic model.
The QC [70], LMTO MD [71] and diatomic-model specific binding energies of the B130 cluster are 4.67, 5.94 and 6.26 eV, respectively. The QC [70] and diatomic-model specific binding energies of the B13+ and B13 charged cluster are 4.73 and 4.91, and 6.03 and 6.05 eV, respectively.
The B13+→B12+ + B10 fragmentation energy calibrated in a hybrid DFT approach is 5.68 eV [82]. The diatomic model gives 13 × 6.03 − 12 × 6.15 = 4.59 eV.
Finally, the B14+→B13+ + B10 fragmentation energy calibrated in a hybrid DFT approach is 3.96 eV [82], which is in satisfactory agreement with diatomic model value of 14 × 5.87 − 13 × 6.03 = 3.79 eV.
The above comparisons between the diatomic model and the previous results on specific binding and fragmentation energies of small boron clusters are summarized in Table 2 and Table 3, respectively. Here, the term “specific fragmentation energy” implies the fragmentation energy per doubled difference between B–B bonds’ numbers in the initial cluster (i.e., before fragmenting) and its fragments. For example, specific binding energies (in diatomic model), numbers of B atoms and B–B bonds of clusters B4+, B2+ and B20 are 3.31, 0.37 and 1.48 eV; 4, 2 and 2; and 5, 1 and 1, respectively. Then, specific B4+→B2+ + B20 fragmentation energy is (4 × 3.31 − 2 × 0.37 − 2 × 1.48)/(2 × (5 − 1 − 1)) = 1.59 eV.
Theoretical E / n n curves obtained in the diatomic model in general features coincide with experimental ones. In particular, we can see maxima in the range B11–B13. At a number of atoms 8 , neutral clusters are predicted to be more stable than their charged isomers. Moreover, at a number of 9 , cations seem to be more stable than anions. However, at a higher number of atoms in boron planar clusters, there is no common trend of relative stability in the dependence of the charge state. As for the specific bind energy averaged by charge states, it saturates. Discrepancies in details seem to be related to the different kinetics of experimental generation processes of small boron clusters, as well as assumptions of the diatomic method used in theoretical calculations.
As for quantitative agreement in specific binding and fragmentation energies and their dependences on charge states and number of atoms in small boron clusters with previously reported (in the most part theoretical) studies, it also seems satisfactory. However, the diatomic model frequently overestimates these energy characteristics. This could be related to the energy phenomenological parameter—dissociation energy of diboron molecule B2—used in constructing the diatomic binding energy curve, as it is measured with too significant an error. Of course, such an almost systematic deviation from previous studies partially should be related to their assumptions on cluster structures and applied methods of calculation or measurement errors.

4. Conclusions

In summary, a previously developed diatomic-type model for calculating clusters’ specific (per atom) binding energy was applied to obtain this key parameter for boron small planar clusters’(Bn, n = 2–15) ground-state isomers in different charge states. The main result of the conducted study was that, in any of the three considered (neutral, single-cationic and single-anionic) charge states, the formation of B11–B13 clusters is preferable. In general features, it is in good agreement with the available experimental data and previous theoretical reports.
As for quantitative agreement in specific binding and fragmentation energies, the diatomic model predicts the values in magnitude comparable with previous results, but it usually overestimates them. These deviations could be explained by different sets of assumptions made in theoretical calculations and/or too significant errors in experimental determination of boron clusters’ energy characteristics, one of which (B2 molecule dissociation energy), in particular, serves for key phenomenological parameter in diatomic model approach to calculating cluster specific binding energy.
Specific binding energies and B–B bond lengths of most stabile neutral, positively and negatively charged species are B130—6.26 eV and 1.59 Å, B12+—6.15 eV and 1.59 Å and B11—6.23 eV and 1.58 Å, respectively.
The success of the diatomic approach can serve for the basis for more detailed calculations of boron small clusters, including not only ground-state, but all the possible planar structural isomers, as well as competitive ring-like clusters. In this way, one can determine not only the specific binding energy but also other important characteristics, such as the cluster dipole moment, ionization potential and electron affinity, vibration, atomization, fragmentation energies, etc.
These results would be not only academic, but also practical interests, as boron (quasi)planar clusters serve for building blocks of borophene and other boron-based nanomaterials perspective for variety of technological applications, such as thin super-hard coatings, radiation shielding, solid fuel production, nanoelectronics, etc.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available via personal communication with proper reasons.

Conflicts of Interest

The author declares no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Boustani, I. Towards novel boron nanostructural materials. Chem. Modell. 2011, 8, 1–44. [Google Scholar]
  2. Chkhartishvili, L. Micro- and nano-structured boron. In Boron: Compounds, Production and Application; Perkins, G.L., Ed.; Nova Sci. Publ.: New York, NY, USA, 2011; Chapter 6; pp. 221–294. [Google Scholar]
  3. Chkhartishvili, L. Nanoboron (An overview). Nano Studies 2011, 3, 227–314. [Google Scholar]
  4. Becker, R.; Chkhartishvili, L.; Martin, P. Boron, the new graphene? Vac. Technol. Coat. 2015, 16, 38–44. [Google Scholar]
  5. Chkhartishvili, L. All-boron nanostructures. In CRC Concise Encyclopedia of Nanotechnology; Kharisov, B.I., Kharissova, O.V., Ortiz–Mendez, U., Eds.; CRC Press: Boca Raton, FL, USA, 2016; Chapter 7; pp. 53–69. [Google Scholar]
  6. Li, D.; Gao, J.; Cheng, P.; He, J.; Yin, Y.; Hu, Y.; Chen, L.; Cheng, Y.; Zhao, J. 2D boron sheets: Structure, growth, and electronic and thermal transport properties. Adv. Funct. Mater. 2019, 30, 1904349. [Google Scholar] [CrossRef]
  7. Tian, Y.; Guo, Z.; Zhang, T.; Lin, H.; Li, Z.; Chen, J.; Deng, S.; Liu, F. Inorganic boron-based nanostructures: Synthesis, optoelectronic properties, and prospective applications. Nanomaterials 2019, 9, 538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Boustani, I. Molecular Modelling and Synthesis of Nanomaterials. Applications in Carbon- and Boron-Based Nanotechnology; Springer Nature: Cham, Switzerland, 2020. [Google Scholar]
  9. Matsuda, I.; Wu, K. (Eds.) 2D Boron: Boraphene, Borophene, Boronene; Springer Nature: Cham, Switzerland, 2021. [Google Scholar]
  10. Alexandrova, A.N.; Boldyrev, A.I.; Zhai, H.-J.; Wang, L.-S. All-boron aromatic clusters as potential new inorganic ligands and building blocks in chemistry. Coord. Chem. Rev. 2006, 250, 2811–2866. [Google Scholar] [CrossRef]
  11. Li, W.-L.; Chen, Q.; Tian, W.-J.; Bai, H.; Zhao, Y.-F.; Hu, H.-S.; Li, J.; Zhai, H.-J.; Li, S.-D.; Wang, L.-S. The B35 cluster with a double-hexagonal vacancy: A new and more flexible structural motif for borophene. J. Am. Chem. Soc. 2014, 136, 12257–12260. [Google Scholar] [CrossRef]
  12. Kiran, B.; Bulusu, S.; Zhai, H.-J.; Yoo, S.; Zeng, X.C.; Wang, L.-S. Planar-to-tubular structural transition in boron clusters: B20 as the embryo of single-walled boron nanotubes. Proc. Natl. Acad. Sci. USA 2005, 102, 961–964. [Google Scholar] [CrossRef] [Green Version]
  13. Boustani, I. Structural transitions and properties of boron nanoclusters. In 17th International Symposium on Boron, Borides and Related Materials; Istanbul Tech. Univ.: Istanbul, Turkey, 2011; p. 49. [Google Scholar]
  14. Arvanitidis, A.G.; Tai, T.B.; Nguyen, M.T.; Ceulemans, A. Quantum rules for planar boron nanoclusters. Phys. Chem. Chem. Phys. 2014, 16, 18311–18318. [Google Scholar] [CrossRef]
  15. Verhaegen, G.; Drowart, J. Mass spectrometric determination of the heat of sublimation of boron and the dissociation energy of B2. J. Chem. Phys. 1962, 37, 1367–1368. [Google Scholar] [CrossRef]
  16. Mar, R.W.; Bedford, R.G. The sublimation of boron. High. Temp. Sci. 1976, 8, 365–376. [Google Scholar]
  17. Goltz, D.M.; Chakrabarti, C.L.; Sturgeon, R.E.; Hughes, D.M.; Gregoire, D.C. Investigation of the vaporization and atomization of boron in a graphite furnace using digital imaging techniques. Appl. Spectrosc. 1995, 49, 1006–1016. [Google Scholar] [CrossRef]
  18. Becker, R.C. Method for Generating a Boron Vapor. U.S. Patent # 5861630, 19 January 1999. [Google Scholar]
  19. Boustani, I.; Becker, R. Boron clusters, single- and multiwalled nanotubes: Theoretical prediction and experimental observation. In Proc. 9th Ann. Nanotechnol. Conf. & Trade Show; Nano Sci. & Technol. Inst.: Boston, MA, USA, 2006; MO 60.802. [Google Scholar]
  20. Medvedovski, E. Preparation of boron nitride-based coatings through thermal diffusion process. Adv. Appl. Ceram. 2018, 117, 221–230. [Google Scholar] [CrossRef]
  21. Unsal, H.; Grasso, S.; Kovalcikova, A.; Hanzel, O.; Tatarkova, M.; Dlouhy, I.; Tatarko, P. In-situ graphene platelets formation and its suppression during reactive spark plasma sintering of boron carbide/titanium diboride composites. J. Eur. Ceram. Soc. 2021, 41, 6281–6289. [Google Scholar] [CrossRef]
  22. Feldman, C.; Moorjani, K.; Blum, N.A. Mass-spectrometry, optical absorption, and electrical properties of boron amorphous films. In Boron–Obtaining, Structure, and Properties; Tsagareishvili, G.V., Tavadze, F.N., Eds.; Nauka: Moscow, Russia, 1974; pp. 130–138. (In Russian) [Google Scholar]
  23. Kutelia, E.R.; Dekanosidze, R.N.; Maisuradze, N.I.; Dzigrashvili, T.A.; Petrov, V.I. Electron microscopic investigation of the structure of ultrafine elementary boron. In Abs. 8th Int. Symp. Boron, Borides, Carbides, Nitrides & Rel. Comp.; Tsagareishvili, G.V., Ed.; Metsniereba: Tbilisi, Georgia, 1984; pp. 12–13. [Google Scholar]
  24. Kervalishvili, P.J.; Kutelia, E.R.; Dzigrashvili, T.A.; Dekanosidze, R.N.; Petrov, V.I. Electron-microscopic study of amorphous boron structure. Phys. Solid State 1985, 27, 1414–1418. (In Russian) [Google Scholar]
  25. Dzigrashvili, T.A.; Kutelia, E.R. On structural relaxation of small particles. Bull. Acad. Sci. Georgian SSR 1988, 129, 53–56. (In Russian) [Google Scholar]
  26. La Placa, S.J.; Roland, P.A.; Wynne, J.J. Boron clusters (Bn, n = 2–52) produced by laser ablation of hexagonal boron nitride. Chem. Phys. Lett. 1992, 190, 163–168. [Google Scholar] [CrossRef]
  27. Liu, C.H.; Peng, W.; Sheng, L.M. Carbon and boron nanoparticles by pulsed-laser vaporization of boron carbide in liquids. Carbon 2001, 39, 144–147. [Google Scholar] [CrossRef]
  28. Xu, S.-J.; Nilles, J.M.; Radisic, D.; Zheng, W.-J.; Stokes, S.; Bowen, K.H.; Becker, R.C.; Boustani, I. Boron cluster anions containing multiple B12 icosahedra. Chem. Phys. Lett. 2003, 379, 282–286. [Google Scholar] [CrossRef]
  29. Oger, E. Strukturaufklärung durch Mobilitätsmessungen an Massenselektierten Clusterionen in der Gasphase (Dissertation); KIT Sci. Publ.: Karlsruher, Germany, 2010. (In German) [Google Scholar]
  30. Atis, M.; Ozdogan, C.; Guvenc, Z.B. Structure and energetic of Bn (n = 2–12) clusters: Electronic structure calculations. Int. J. Quantum Chem. 2007, 107, 729–744. [Google Scholar] [CrossRef]
  31. Tai, T.B.; Grant, D.J.; Nguyen, M.T.; Dixon, D.A. Thermochemistry and electronic structure of small boron clusters (Bn, n = 5−13) and their anions. J. Phys. Chem. A 2010, 114, 994–1007. [Google Scholar] [CrossRef] [PubMed]
  32. Tai, T.B.; Nguyen, M.T. Electronic structure and photoelectron spectra of Bn with n = 26–29: An overview of structural characteristics and growth mechanism of boron clusters. Phys. Chem. Chem. Phys. 2015, 17, 13672–13679. [Google Scholar] [CrossRef] [PubMed]
  33. Wu, X.; Sai, L.; Zhou, S.; Zhou, P.; Chen, M.; Springborg, M.; Zhao, J. Competition between tubular, planar and cage geometries: A complete picture of structural evolution of Bn (n = 31–50) clusters. Phys. Chem. Chem. Phys. 2020, 22, 12959–12966. [Google Scholar] [CrossRef] [PubMed]
  34. Vedeneev, V.I.; Gurvich, L.V.; Kondrat’ev, V.N.; Medvedev, V.L.; Frankevich, E.L. Chemical Bonds Breaking Energy. Ionization Potentials and Electron Affinity (Handbook); Acad. Sci. USSR: Moscow, Russia, 1962. (In Russian) [Google Scholar]
  35. Graham, W.R.M.; Weltner, W. B atoms, B2 and H2BO molecules: ESR and optical spectra at 4 K. J. Chem. Phys. 1976, 65, 1516–1521. [Google Scholar] [CrossRef]
  36. Dupuis, M.; Liu, B. The ground electronic state of B2. J. Chem. Phys. 1978, 68, 2902–2910. [Google Scholar] [CrossRef]
  37. Martin, J.M.L.; Francois, J.P.; Gijbels, R. Ab initio study of boron, nitrogen, and boron–nitrogen clusters. I. Isomers and thermochemistry of B3, B2N, BN2, and N3. J. Chem. Phys. 1989, 90, 6469–6485. [Google Scholar] [CrossRef]
  38. Deutsch, P.W.; Curtiss, L.A.; Pople, J.A. Boron dimer: Dissociation energy and ionization potentials. Chem. Phys. Lett. 1990, 174, 33–36. [Google Scholar] [CrossRef]
  39. Huber, K.P.; Herzberg, H. Molecular Spectra and Molecular Structure. IV. Constants of Diatomic Molecules; van Nostrand Reinhold Co.: New York, NY, USA, 1979. [Google Scholar]
  40. Langhoff, S.R.; Bauschlicher, C.W. Theoretical study of the spectroscopy of B2. J. Chem. Phys. 1991, 95, 5882–5888. [Google Scholar] [CrossRef]
  41. Chkhartishvili, L.; Lezhava, D.; Tsagareishvili, O.; Gulua, D. Ground-state parameters of diatomic molecules B2, BC, BN, and BO. Proc. Georgian Police Acad. 1999, 1, 195–300. (In Russian) [Google Scholar]
  42. Chkhartishvili, L.; Lezhava, D.; Tsagareishvili, O. Quasi-classical determination of electronic energies and vibration frequencies in boron compounds. J. Solid State Chem. 2000, 154, 148–152. [Google Scholar] [CrossRef]
  43. Chkhartishvili, L. On quasi-classical estimations of boron nanotubes ground-state parameters. J. Phys. Conf. Ser. 2009, 176, 012013. [Google Scholar] [CrossRef]
  44. Chkhartishvili, L. Molar binding energy of the boron nanosystems. In Proc. 4th Int. Boron Symp.; Konuk, A., Kurama, H., Ak, H., Iphar, M., Eds.; Osmangazi Univ.–TMMOB: Ankara, Turkey, 2009; pp. 153–160. [Google Scholar]
  45. Chkhartishvili, L. Nanotubular boron: Ground-state estimates. In New Developments in Materials Science; Chikoidze, E., Tchelidze, T., Eds.; Nova Sci. Publ.: New York, NY, USA, 2013; Chapter 8; pp. 67–80. [Google Scholar]
  46. Bruna, P.J.; Wright, J.S. Strongly bound multiply excited states of B2+ and B2. J. Chem. Phys. 1989, 91, 1126–1136. [Google Scholar] [CrossRef]
  47. Mierzwa, G.; Gordon, A.J.; Berski, S. The nature of the triple B≡B, double B=B, single B–B, and one-electron B∙B boron–boron bonds from the topological analysis of Electron Localization Function (ELF) perspective. J. Mol. Str. 2020, 1221, 128530. [Google Scholar] [CrossRef]
  48. Boustani, I. New convex and spherical structures of bare boron clusters. J. Solid State Chem. 1997, 133, 182–189. [Google Scholar] [CrossRef]
  49. Boustani, I. New quasi-planar surfaces of bare boron. Surf. Sci. 1997, 370, 355–363. [Google Scholar] [CrossRef]
  50. Boustani, I.; Quandt, A. Boron in ab initio calculations. Comput. Mater. Sci. 1998, 11, 132–137. [Google Scholar] [CrossRef]
  51. Hernandez, R.; Simons, J. Electronic energies, geometries, and vibrational frequencies of the ground and low-lying excited states of the boron trimer. J. Chem. Phys. 1991, 94, 2961–2967. [Google Scholar] [CrossRef] [Green Version]
  52. Zhai, H.-J.; Wang, L.-S.; Alexandrova, A.N.; Boldyrev, A.I.; Zakrzewski, V.G. Photoelectron spectroscopy and ab initio study of B3 and B4 anions and their neutrals. J. Chem. Phys. A 2003, 107, 9319–9328. [Google Scholar] [CrossRef]
  53. Martin, J.M.L.; Francois, J.P.; Gijbels, R. Potential energy surface of B4 and total atomization energies of B2, B3, and B4. Chem. Phys. Lett. 1992, 189, 529–536. [Google Scholar] [CrossRef]
  54. Koutecky, J.; Pacchioni, G.; Jeung, G.H.; Hass, E.C. Comparative study of tetramers built from Ia, IIa, IIIa, and IVa atoms. Surf. Sci. 1985, 156, 650–669. [Google Scholar] [CrossRef]
  55. Li, Q.S.; Jin, H.W. Structure and stability of B5, B5+, and B5 clusters. J. Phys. Chem. A 2002, 106, 7042–7047. [Google Scholar] [CrossRef]
  56. Zhai, H.-J.; Wang, L.-S.; Alexandrova, A.N.; Boldyrev, A.I. Electronic structure and chemical bonding of B5 and B5 by photoelectron spectroscopy and ab initio calculations. J. Chem. Phys. 2002, 117, 7917–7924. [Google Scholar] [CrossRef]
  57. Alexandrova, A.N.; Boldyrev, A.I.; Zhai, H.-J.; Wang, L.-S.; Steiner, E.; Fowler, P.W. Structure and bonding in B6 and B6: Planarity and antiaromaticity. J. Phys. Chem. A 2003, 107, 1359–1369. [Google Scholar] [CrossRef]
  58. Alexandrova, A.N.; Boldyrev, A.I.; Zhai, H.-J.; Wang, L.-S. Electronic structure, isomerism, and chemical bonding in B7 and B7. J. Phys. Chem. A 2004, 108, 3509–3517. [Google Scholar] [CrossRef]
  59. Hanley, L.; Anderson, S.L. Production and collision-induced dissociation of small boron cluster ions. J. Phys. Chem. 1987, 91, 5161–5163. [Google Scholar] [CrossRef]
  60. Ray, A.K.; Howard, I.A.; Kanal, K.M. Structure and binding in small neutral and cationic boron clusters. Phys. Rev. B 1992, 45, 14247–14255. [Google Scholar] [CrossRef]
  61. Kato, H.; Tanaka, E. Stabilities of small Ben and Bn clusters (4 ≤ n ≤ 8) by vibrational analysis. J. Comput. Chem. 1991, 12, 1097–1109. [Google Scholar] [CrossRef]
  62. Zhai, H.-J.; Alexandrova, A.N.; Birch, K.A.; Boldyrev, A.I.; Wang, L.-S. Hepta- and octacoordinate boron in molecular wheels of eight- and nine-atom boron clusters: Observation and confirmation. Angew. Chem. Int. Ed. 2003, 42, 6004–6008. [Google Scholar] [CrossRef]
  63. Drummond, M.L.; Meunier, V.; Sumpter, B.G. Structure and stability of small boron and boron oxide clusters. J. Phys. Chem. A 2007, 111, 6539–6551. [Google Scholar] [CrossRef]
  64. Kato, H.; Yamashita, K.; Morokuma, K. Ab initio MO study of neutral and cationic boron clusters. Chem. Phys. Lett. 1992, 190, 361–366. [Google Scholar] [CrossRef]
  65. Bambakidis, G.; Wagner, R.P. Electronic structure and binding energy of the icosahedral boron cluster B12. J. Phys. Chem. Solids 1981, 42, 1023–1025. [Google Scholar] [CrossRef]
  66. Kawai, R.; Weare, J.H. Instability of the B12 icosahedral cluster: Rearrangement to a lower energy structure. J. Chem. Phys. 1991, 95, 1151–1159. [Google Scholar] [CrossRef]
  67. Bhattacharyya, P.; Boustani, I.; Shukla, A. First principles study of structural and optical properties of B12 isomers. arXiv 2018, arXiv:1802.01072. [Google Scholar] [CrossRef] [Green Version]
  68. Hanley, L.; Whitten, J.L.; Anderson, S.L. Collision-induced dissociation and ab initio studies of boron cluster ions: Determination of structures and stabilities. J. Phys. Chem. 1988, 92, 5803–5812. [Google Scholar] [CrossRef]
  69. Hanley, L.; Whitten, J.L.; Anderson, S.L. Collision-induced dissociation and ab initio studies of boron cluster ions: Determination of structures and stabilities [Erratum to document cited in CA109(18):156723t]. J. Phys. Chem. 1990, 94, 2218. [Google Scholar] [CrossRef]
  70. Akman, N.; Tas, M.; Ozdogan, C.; Boustani, I. Ionization energies, Coulomb explosion, fragmentation, geometric, and electronic structures of multicharged boron clusters Bn (n = 2–13). Phys. Rev. B 2011, 84, 075463. [Google Scholar] [CrossRef]
  71. Cao, P.-l.; Zhao, W.; Li, B.-X.; Song, B.; Zhou, X.-Y. A full-potential linear-muffin-tin-orbital molecular-dynamics study of B7, B10 and B13 clusters. J. Phys. Cond. Matter 2001, 13, 5065–5076. [Google Scholar] [CrossRef]
  72. Kato, H.; Yamashita, K.; Morokuma, K. Ab initio study of neutral and cationic B12 and B13 clusters. Bull. Chem. Soc. Jpn. 1993, 66, 3358–3361. [Google Scholar] [CrossRef]
  73. Gu, F.L.; Yang, X.; Tang, A.-C.; Jiao, H.; von Schleyer, P.R. Structure and stability of B13+ clusters. J. Comput. Chem. 1998, 19, 203–214. [Google Scholar] [CrossRef]
  74. Kiran, B.; Kumar, G.G.; Nguyen, M.T.; Kandalam, A.K.; Jena, P. Origin of the unusual stability of B12 and B13+ clusters. Inorg. Chem. 2009, 48, 9965–9967. [Google Scholar] [CrossRef]
  75. Kawai, R.; Weare, J.H. Anomalous stability of B13+ clusters. Chem. Phys. Lett. 1992, 191, 311–314. [Google Scholar] [CrossRef]
  76. Fowler, J.E.; Ugalde, J.M. The curiously stable B13+ cluster and its neutral and anionic counterparts: The advantages of planarity. J. Phys. Chem. A 2000, 104, 397–403. [Google Scholar] [CrossRef]
  77. Aihara, J.-I. B13+ is highly aromatic. J. Phys. Chem. A 2001, 105, 5486–5489. [Google Scholar] [CrossRef]
  78. Shinde, R.; Tayade, M. Optical absorption in B13cluster: A Time-dependent density functional approach. AIP Conf. Proc. 2012, G–163, 1–2. [Google Scholar]
  79. Boustani, I. Systematic LSD investigation on cationic boron clusters: Bn+ (n = 2–14). Int. J. Quant. Chem. 1994, 52, 1081–1111. [Google Scholar] [CrossRef]
  80. Boustani, I. A comparative study of ab initio SCF–CI and DFT. Example of small boron clusters. Chem. Phys. Lett. 1995, 233, 273–278. [Google Scholar] [CrossRef]
  81. Boustani, I. Structure and stability of small boron clusters. A density functional theoretical study. Chem. Phys. Lett. 1995, 240, 135–140. [Google Scholar] [CrossRef]
  82. Ricca, A.; Bauschlicher, C.W., Jr. The structure and stability of Bn+ clusters. Chem. Phys. 1996, 208, 233–242. [Google Scholar] [CrossRef]
  83. Boustani, I. Systematic ab initio investigation of bare boron clusters: Determination of the geometry and electronic structures of Bn (n = 2–14). Phys. Rev. B 1997, 55, 16426–16438. [Google Scholar] [CrossRef]
  84. Zhai, H.-J.; Kiran, B.; Li, J.; Wang, L.-S. Hydrocarbon analogues of boron clusters–Planarity, aromaticity and antiaromaticity. Nat. Mater. 2003, 2, 827–833. [Google Scholar] [CrossRef]
  85. Sergeeva, A.P.; Zubarev, D.Y.; Zhai, H.-J.; Boldyrev, A.I.; Wang, L.-S. A photoelectron spectroscopic and theoretical study of B16 and B162−: An all-boron naphthalene. J. Am. Chem. Soc. 2008, 130, 7244–7246. [Google Scholar] [CrossRef] [PubMed]
  86. Huang, W.; Sergeeva, A.P.; Zhai, H.-J.; Averkiev, B.B.; Wang, L.-S.; Boldyrev, A.I. A concentric planar doubly π-aromatic B19 cluster. Nat. Chem. 2010, 2, 202–206. [Google Scholar] [CrossRef] [PubMed]
  87. An, W.; Bulusu, S.; Gao, Y.; Zeng, X.C. Relative stability of planar versus double-ring tubular isomers of neutral and anionic boron cluster B20 and B20. J. Chem. Phys. 2006, 124, 154310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Oger, E.; Crawford, N.R.M.; Kelting, R.; Weis, P.; Kappes, M.M.; Ahlrichs, R. Boron cluster cations: Transition from planar to cylindrical structures. Angew. Chem. Int. Ed. 2007, 46, 8503–8506. [Google Scholar] [CrossRef]
  89. Chacko, S.; Kanhere, D.G.; Boustani, I. Ab initio density functional investigation of B24 clusters: Rings, tubes, planes, and cages. Phys. Rev. B 2003, 68, 035414. [Google Scholar] [CrossRef] [Green Version]
  90. Piazza, Z.A.; Popov, I.A.; Li, W.-L.; Pal, R.; Zeng, X.C.; Boldyrev, A.I.; Wang, L.-S. A photoelectron spectroscopy and ab initio study of the structures and chemical bonding of the B25 cluster. J. Chem. Phys. 2014, 141, 034303. [Google Scholar] [CrossRef] [Green Version]
  91. Li, W.-L.; Pal, R.; Piazza, Z.A.; Zeng, X.C.; Wang, L.-S. B27: Appearance of the smallest planar boron cluster containing a hexagonal vacancy. J. Chem. Phys. 2015, 142, 204305. [Google Scholar] [CrossRef] [Green Version]
  92. Boustani, I.; Rubio, A.; Alonso, J.A. Ab initio study of B32 clusters: Competition between spherical, quasiplanar and tubular isomers. Chem. Phys. Lett. 1999, 311, 21–28. [Google Scholar] [CrossRef]
  93. Chen, Q.; Wei, G.-F.; Tian, W.-J.; Bai, H.; Liu, Z.-P.; Zhai, H.-J.; Li, S.-D. Quasi-planar aromatic B36 and B36 clusters: All-boron analogues of coronene. Phys. Chem. Chem. Phys. 2014, 16, 18282–18287. [Google Scholar] [CrossRef]
  94. Piazza, Z.A.; Hu, H.-S.; Li, W.-L.; Zhao, Y.-F.; Li, J.; Wang, L.-S. Planar hexagonal B36 as a potential basis for extended single-atom layer boron sheets. Nat. Commun. 2014, 5, 3113. [Google Scholar] [CrossRef]
  95. Bai, H.; Chen, T.-T.; Chen, Q.; Zhao, X.-Y.; Zhang, Y.-Y.; Chen, W.-J.; Li, W.-L.; Cheung, L.F.; Bai, B.; Cavanagh, J.; et al. Planar B41 and B42 clusters with double-hexagonal vacancies. Nanoscale 2019, 11, 23286–23295. [Google Scholar] [CrossRef] [PubMed]
  96. Mukhopadhyay, S.; He, H.; Pandey, R.; Yap, Y.K.; Boustani, I. Novel spherical boron clusters and structural transition from 2D quasi-planar structures to 3D double-rings. J. Phys. Conf. Ser. 2009, 176, 012028. [Google Scholar] [CrossRef] [Green Version]
  97. Oyaidzu, M.; Yoshikawa, A.; Kodama, H.; Oya, Y.; Sagara, A.; Noda, N.; Okuno, K. Preparation of pure boron coating film and its characterization by XPS and TDS. Appl. Surf. Sci. 2005, 244, 240–243. [Google Scholar] [CrossRef]
  98. Becker, R.C.; Bunker, S.N. Method for depositing boron-rich coatings. Patent Appl. # US 2005/0208218 A1, 22 September 2005. [Google Scholar]
  99. Becker, R.; Chkhartishvili, L.; Martin, P. Tribological applications for boron. Vac. Technol. Coat. 2015, 16, 36–41. [Google Scholar]
  100. Korbut, E.V.; Labunets, V.F.; Radko, O.V.; Zagrebelnyi, V.V.; Yakobchuk, O.E. Improve of the high-speed steel wear resistance using boron coating. In Proc. 5th Int. Conf. High Mat Tech; IPMS: Kyiv, Ukraine, 2015; p. 189. [Google Scholar]
  101. Zhu, Z.; Kwon, D.-G.; Kwon, Y.-K.; Tomanek, D. Enhancing mechanical toughness of aluminum surfaces by nano-boron implantation: An ab initio study. Chem. Phys. Lett. 2015, 620, 25–28. [Google Scholar] [CrossRef]
  102. Demirbas, A. Energy from boron and non-nuclear metallic fuels. Energy Sources A 2008, 30, 1108–1113. [Google Scholar] [CrossRef]
  103. Young, G.; Sullivan, K.; Zachariah, M.R.; Yu, K. Combustion characteristics of boron nanoparticles. Combust. Flame 2009, 156, 322–333. [Google Scholar] [CrossRef]
  104. Hussmann, B.; Pfitzner, M. Extended combustion model for single boron particles–Part I: Theory. Combust. Flame 2010, 157, 803–821. [Google Scholar] [CrossRef]
  105. Hussmann, B.; Pfitzner, M. Extended combustion model for single boron particles–Part II: Validation. Combust. Flame 2010, 157, 822–833. [Google Scholar] [CrossRef]
  106. Katoshevski, D.; Chkhartishvili, L. Manipulating grouping dynamics of nanoscale boron particles as basis for environmentally friendlier combustion and efficient filtration. In Nanotechnology in Environmental Science; Hussain, C.M., Mishra, A.K., Eds.; Wiley–VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2018; Chapter 13; pp. 413–442. [Google Scholar]
  107. Ojha, P.K.; Karmakar, S. Boron for liquid fuel engines–A review on synthesis, dispersion stability in liquid fuel, and combustion aspects. Prog. Aerospace Sci. 2018, 100, 18–45. [Google Scholar] [CrossRef]
  108. Gan, Y.; Lim, Y.S.; Qiao, L. Combustion of nanofluid fuels with the addition of boron and iron particles at dilute and dense concentrations. Combust. Flame 2012, 159, 1732–1740. [Google Scholar] [CrossRef]
  109. Chintersingh, K.-L.; Schoenitz, M.; Dreizin, E.L. Combustion of boron and boron–iron composite particles in different oxidizers. Combust. Flame 2018, 192, 44–58. [Google Scholar] [CrossRef]
  110. Epshteyn, A.; Weismiller, M.R.; Huba, Z.J.; Maling, E.L.; Chaimowitz, A.S. Optimization of a high-energy Ti–Al–B nanopowder fuel. Energy Fuels 2017, 31, 1811–1819. [Google Scholar] [CrossRef]
  111. Hanley, L.; Anderson, S.L. Oxidation of small boron cluster ions (B1–13+) by oxygen. J. Chem. Phys. 1988, 89, 2848–2860. [Google Scholar] [CrossRef]
  112. Ruatta, S.A.; Hintz, P.A.; Anderson, S.L. Boron cluster ion oxidation: Reactions with CO2, dissociation of boron cluster oxide (BnO+) ions, and sequential oxidation. J. Chem. Phys. 1991, 94, 2833–2847. [Google Scholar] [CrossRef]
  113. Chkhartishvili, L.; Tsagareishvili, O.; Gabunia, D. 10B-based materials for neutron-shielding. In Proc. 1st Int. Conf. “Modern Technologies and Methods of Inorganic Materials Science”; Khantadze, J., Chkhartishvili, L., Gabunia, D., Ramazashvili, D., Eds.; Meridian: Tbilisi, Georgia, 2012; pp. 188–202. [Google Scholar]
  114. Chkhartishvili, L. Interaction between neutron-radiation and boron-containing materials. In Radiation Synthesis of Materials and Compounds; Kharisov, B.I., Kharissova, O.V., Mendez, U.O., Eds.; CRC Press–Taylor & Francis Group: Boca Raton, FL, USA, 2013; Chapter 3; pp. 43–80. [Google Scholar]
  115. Chkhartishvili, L. Boron-contained nanostructured materials for neutron-shields. In Nanostructured Materials for the Detection of CBRN; Bonca, J., Kruchinin, S., Eds.; Springer Science: Dordrecht, The Netherlands, 2018; Chapter 11; pp. 133–154. [Google Scholar]
  116. Chkhartishvili, L. Neutron-fluence nanosensors based on boron-containing materials. In Nanomaterials for Environmental Protection; Kharisov, B.I., Kharissova, O.V., Rasika Dias, H.V., Eds.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2014; Chapter 26; pp. 445–449. [Google Scholar]
  117. Chkhartishvili, L.; Tsagareishvili, O.; Tavadze, G. Neutron detectors based on 10B-containing nanomaterials. In Nuclear Radiation Nanosensors and Nanosensory Systems; Kervalishvili, P.J., Yannakopoulos, P.H., Eds.; Springer Science: Dordrecht, The Netherlands, 2016; Chapter 12; pp. 187–196. [Google Scholar]
  118. Chkhartishvili, L.; Murusidze, I. Band structure of all-boron 2D metallic crystals as a prospective electromagnetic shielding material. In Proc. Int. Conf. “Fundamental and Applied Nano Electromagnetics”; Belarusian State Univ.: Minsk, Belarus, 2012; p. 11. [Google Scholar]
  119. Wang, Y.-J.; Zhao, X.-Y.; Chen, Q.; Zhai, H.-J.; Li, S.-D. B11: A moving subnanoscale tank tread. Nanoscale 2015, 7, 16054–16060. [Google Scholar] [CrossRef]
  120. He, C.-C.; Xu, S.-G.; Zhao, Y.-J.; Xu, H.; Yang, X.-B. All-boron planar ferromagnetic structures: From clusters to monolayers. Nanoscale 2021, 13, 9881–9887. [Google Scholar] [CrossRef]
  121. Wang, Z.-Q.; Lu, T.-Y.; Wang, H.-Q.; Feng, Y.P.; Zheng, J.-C. Review of borophene and its potential applications. Front. Phys. 2019, 14, 33403. [Google Scholar] [CrossRef] [Green Version]
  122. Fermi, E. Molecules and Crystals; Barth: Leipzig, Germany, 1938. [Google Scholar]
  123. Novikova, S.I. Thermal Expansion of Solids; Nauka: Moscow, Russia, 1974. (In Russian) [Google Scholar]
  124. Slutsker, A.I.; Gilyarov, V.L.; Luk’yanenko, A.S. Energy features of an adiabatically loaded anharmonic oscillator. Phys. Solid State 2006, 48, 1947–1953. [Google Scholar] [CrossRef]
  125. Chkhartishvili, L.; Becker, R.; Avci, R. Relative stability of boron quasi-planar clusters. In Proc. Int. Conf. “Advanced Materials and Technologies”; Darsavelidze, G., Guldamashvili, A., Chedia, R., Sichinava, A., Kadaria, M., Eds.; Universal: Tbilisi, Georgia, 2015; pp. 42–46. [Google Scholar]
  126. Chkhartishvili, L. Small elemental clusters in pair interaction approximation. In Proc. 4th Int. Conf. Exh. Adv. Nano Mater.; IAEMM: Montreal, QC, Canada, 2016; pp. 128–132. [Google Scholar]
  127. Chkhartishvili, L. Quasi-planar elemental clusters in pair interactions approximation. Open Phys. 2016, 14, 617–620. [Google Scholar] [CrossRef]
  128. Chkhartishvili, L. Planar clusters of identical atoms in equilibrium: 1. Diatomic model approach. Am. J. Nano Res. Appl. 2017, 5, 1–4. [Google Scholar]
  129. Chkhartishvili, L.; Becker, R. Effective atomic charges and dipole moment of small boron clusters. In Proc. 3rd Int. Conf. Exh. Adv. Nano Mater.; IAEMM: Ottawa, ON, Canada, 2015; pp. 130–147. [Google Scholar]
  130. Becker, R.; Chkhartishvili, L. Dipole moment of quasi-planar boron clusters. Nano Studies 2015, 11, 29–48. [Google Scholar]
  131. Chkhartishvili, L. Boron quasi-planar clusters. A mini-review on diatomic approach. In Proc. IEEE 7th Int. Conf. Nanomaterials: Applications & Properties (NAP–2017), Track: Nanomaterials for Electronics, Spintronics and Photonics; Sumy State Univ.: Sumy, Ukraine, 2017; Part 4; pp. 04NESP10-1–04NESP10-5. [Google Scholar]
  132. Chkhartishvili, L. Boron triangular sheet: Calculation of ground-state and electronic-structure parameters. In Proc. 7th Int. Conf. Exh. Adv. Nano Mater.; IAEMM: Montreal, QC, Canada, 2019; pp. 11–16. [Google Scholar]
  133. Chkhartishvili, L. Relative stability of planar clusters B11, B12, and B13 in neutral- and charged-states. Char. Appl. Nanomater. 2019, 2, 761-1–761-7. [Google Scholar] [CrossRef]
  134. Chkhartishvili, L.; Murusidze, I.; Becker, R. Electronic structure of boron flat holeless sheet. Condensed Matter 2019, 4, 28. [Google Scholar] [CrossRef] [Green Version]
  135. Chkhartishvili, L. Nanoclusters binding energy in diatomic model. Int. J. Adv. Nano Comput. Anal. 2021, 1, 80–83. [Google Scholar]
Figure 1. Quasi-classically calculated interatomic potentials for boron-containing diatomic molecules B2, BC, BN and BO versus inter-nuclear distances [42]. Copyright (2000), with permission from Elsevier.
Figure 1. Quasi-classically calculated interatomic potentials for boron-containing diatomic molecules B2, BC, BN and BO versus inter-nuclear distances [42]. Copyright (2000), with permission from Elsevier.
Molecules 27 01469 g001
Figure 2. Specific binding energy of neutral (♦), positively (■) and negatively (▲) charged boron small planar clusters in dependence on number of atoms calculated in diatomic model.
Figure 2. Specific binding energy of neutral (♦), positively (■) and negatively (▲) charged boron small planar clusters in dependence on number of atoms calculated in diatomic model.
Molecules 27 01469 g002
Figure 3. Experimentally recorded mass spectra of (a) neutral [130], Copyright (2015), with permission from Authors; (b) positively [29], Copyright (2010), with permission from KIT Scientific Publishing; and (c) negatively charged [28] boron clusters, Copyright (2003), with permission from Elsevier.
Figure 3. Experimentally recorded mass spectra of (a) neutral [130], Copyright (2015), with permission from Authors; (b) positively [29], Copyright (2010), with permission from KIT Scientific Publishing; and (c) negatively charged [28] boron clusters, Copyright (2003), with permission from Elsevier.
Molecules 27 01469 g003
Figure 4. Binding energy per atom of neutral and ionized boron clusters as function of size [70], Copyright (2011), with permission from American Physical Society.
Figure 4. Binding energy per atom of neutral and ionized boron clusters as function of size [70], Copyright (2011), with permission from American Physical Society.
Molecules 27 01469 g004
Table 1. Specific binding energy of boron small planar clusters calculated in diatomic model.
Table 1. Specific binding energy of boron small planar clusters calculated in diatomic model.
Number
of Atoms
n
StructureNumber
of Bonds
N
Charge
State
Charge Numbers in Dependence
on Coordination Number
Parameter
Z
Specific
Binding
Energy
E / n ,   eV
Bonds
Length,
a , Å
12346
2 Molecules 27 01469 i0011B20 0 01.481.59
B2+ + 1 2 +0.25000.371.65
B2 1 2 +0.25000.371.65
3 Molecules 27 01469 i0023B30 0 02.961.59
B3+ + 1 3 +0.33331.961.62
B3 1 3 +0.33331.961.62
4 Molecules 27 01469 i0035B40 + 1 5 1 5 −0.12003.971.58
B4+ + 2 5 + 1 10 +0.17003.311.60
B4 0 1 2 +0.25003.131.60
5 Molecules 27 01469 i0047B50 + 2 7 1 14 3 7 −0.21944.541.58
B5+ + 3 7 + 1 7 1 7 −0.02044.171.59
B5 + 1 7 2 7 5 7 +0.20413.771.60
6 Molecules 27 01469 i0059B60 + 1 3 1 3 −0.33334.941.58
B6+ + 4 9 1 9 −0.25934.821.58
B6 + 2 9 5 9 +0.18524.151.60
7 Molecules 27 01469 i00612B70 + 1 8 3 4 −0.46885.671.58
B7+ + 1 4 1 2 −0.37505.551.58
B7 0 1 05.071.59
8 Molecules 27 01469 i00714B80 + 3 7 + 1 7 1 7 5 7 −0.28575.501.59
B8+ + 1 2 + 1 4 0 1 2 −0.31255.531.58
B8 + 5 14 + 1 28 2 7 13 14 +0.25894.881.60
9 Molecules 27 01469 i00816B90 + 7 16 + 5 32 1 8 11 16 +0.25005.001.59
B9+ + 1 2 + 1 4 0 1 2 05.251.59
B9 + 3 8 + 1 16 1 4 7 8 +1.00004.251.61
10 Molecules 27 01469 i00919B100 + 4 19 1 19 11 19 −0.38505.961.59
B10+ + 11 38 + 1 19 8 19 −0.24655.841.59
B10 + 5 38 3 19 14 19 +0.41275.241.60
11 Molecules 27 01469 i01021B110 + 10 21 + 9 42 1 21 4 7 +0.01025.631.59
B11+ + 11 21 + 2 7 + 1 21 3 7 +0.23815.451.59
B11 + 3 7 + 1 7 1 7 5 7 −0.71436.231.58
12 Molecules 27 01469 i01124B120 + 1 4 0 1 2 +0.18755.771.59
B12+ + 5 16 + 1 12 3 8 −0.31256.151.59
B12 + 3 16 1 12 5 8 +0.24615.721.59
13 Molecules 27 01469 i01226B130 + 1 4 0 1 2 −0.50006.261.59
B13+ + 4 13 + 1 13 5 13 −0.16576.031.59
B13 + 5 26 1 13 8 13 −0.20716.051.59
14 Molecules 27 01469 i01328B140 + 1 2 + 1 4 0 1 2 +0.56255.551.60
B14+ + 15 28 + 17 56 + 1 14 11 28 +0.05905.871.59
B14 + 13 28 + 11 56 1 14 17 28 +1.61514.871.61
15 Molecules 27 01469 i01430B150 + 1 2 0 1 2 +0.75005.461.60
B15+ + 8 15 + 1 15 2 5 +0.41335.661.59
B15 + 7 15 1 15 3 5 +1.41335.061.60
Table 2. Comparison of the literature data on specific binding energies (in eV) of boron small clusters with values calculated in diatomic model.
Table 2. Comparison of the literature data on specific binding energies (in eV) of boron small clusters with values calculated in diatomic model.
ClusterLiterature DataDiatomic Model
B201.29–1.53 [15]
1.36 [30]
1.13–1.56 [34]
1.43 [37]
1.36 [38]
1.35–1.39 [40]
1.40 [41,42]
1.35 [53]
1.39 [70]
1.48
B2+0.97 [46]
0.74 [69]
1.16 [70]
0.98 [82]
0.37
B22.24 [70]0.37
B32.76 [30]
2.86 [37]
2.74–2.79 [53]
2.82 [70]
2.96
B3+2.46 [70]1.96
B33.58 [70]1.96
B43.37 [30]
3.37–3.41 [53]
2.42 [54]
3.45 [70]
3.97
B4+3.28 [70]3.31
B43.76 [70]3.13
B503.67 [30]
3.68 [31]
3.76 [70]
4.54
B5+3.82 [70]4.17
B54.04 [31]
4.11 [70]
3.77
B603.79 [30]
3.84 [31]
3.88 [70]
4.94
B6+3.93 [70]4.83
B64.16 [31]
4.25 [70]
4.15
B704.07 [30]
4.11 [31]
4.17 [70]
5.24 [71]
5.67
B7+4.28 [70]5.55
B74.49 [31]
4.48 [70]
5.07
B804.31 [30]
4.33 [31]
4.41 [70]
5.50
B8+4.42 [70]5.53
B84.71 [31]
4.71 [70]
4.88
B904.30 [30]
4.39 [70]
5.00
B9+4.46 [70]5.25
B94.55 [31]
4.71 [70]
4.25
B1004.45 [30]
4.56 [70]
5.78 [71]
5.96
B10+4.57 [70]5.84
B104.79 [70]5.24
B1104.52 [30]
4.63 [70]
5.63
B11+4.69 [70]5.45
B114.89 [70]6.23
B1204.60 [30]
4.60 [67]
4.71 [70]
5.77
B12+4.72 [70]6.15
B124.85 [70]5.72
B1304.67 [70]
5.94 [71]
6.26
B13+4.73 [70]6.03
B134.91 [70]6.05
Table 3. Comparison of the literature data on specific fragmentation energies (in eV) of boron small clusters with values calculated in diatomic model.
Table 3. Comparison of the literature data on specific fragmentation energies (in eV) of boron small clusters with values calculated in diatomic model.
ChanelLiterature DataDiatomic Model
B30→B20 + B101.24 [51]1.48
B3+→B20 + B1+0.36 [68]
1.08 [82]
1.29
B3+→B2+ + B100.50 [68]0.73
B4+→B30 + B1+0.44 [68]
1.06 [82]
1.09
B4+→B3+ + B100.70 [68]1.84
B4+→B20 + B2+0.80 [68]2.39
B5+→B40 + B1+0.50 [68]
1.24 [82]
1.24
B6+→B5+ + B100.55 [68]
1.03 [82]
2.02
B7+→B6+ + B101.01 [82]1.66
B8+→B7+ + B101.26 [82]1.35
B9+→B8+ + B101.07 [82]0.75
B10+→B9+ + B100.93 [82]1.86
B11+→B10+ + B101.30 [82]0.39
B12+→B11+ + B100.77 [82]2.31
B13+→B12+ + B101.42 [82]1.15
B14+→B13+ + B100.99 [82]0.95
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chkhartishvili, L. Relative Stability of Boron Planar Clusters in Diatomic Molecular Model. Molecules 2022, 27, 1469. https://doi.org/10.3390/molecules27051469

AMA Style

Chkhartishvili L. Relative Stability of Boron Planar Clusters in Diatomic Molecular Model. Molecules. 2022; 27(5):1469. https://doi.org/10.3390/molecules27051469

Chicago/Turabian Style

Chkhartishvili, Levan. 2022. "Relative Stability of Boron Planar Clusters in Diatomic Molecular Model" Molecules 27, no. 5: 1469. https://doi.org/10.3390/molecules27051469

Article Metrics

Back to TopTop