Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Heats of formation on the way from B2H6 to B20H16: thermochemical consequences of multicenter bonding in ab initio and DFT methods

Jindřich Fanfrlík *a, Jan Řezáč a, Drahomír Hnyk b and Josef Holub b
aInstitute of Organic Chemistry and Biochemistry of the Czech Academy of Sciences, 160 00 Prague, Czech Republic. E-mail: fanfrlik@uochb.cas.cz
bInstitute of Inorganic Chemistry of the Czech Academy of Sciences, Husinec-Řež 250 68, Czech Republic

Received 10th September 2024 , Accepted 4th November 2024

First published on 5th November 2024


Abstract

The objective of this study is to evaluate the effectiveness of various computational methods in reproducing the experimental heats of formation of boron hydrides using the atomization energy approach. The results have demonstrated that the empirical dispersion combined with the BJ damping function provided too large intramolecular dispersion energies, thereby compromising the accuracy of the outcomes produced by the DFT-D3 methods. Additionally, the CCSD(T) method has reproduced the experimental values only when combined with a basis set optimized for an accurate description of the core-valence correlation effect. Furthermore, the number of multicenter bonds present in the molecules under examination has also reflected their stability, as indicated by the heats of formation. Finally, a five-center two-electron (5c–2e) bond has emerged in B5H9, by applying the intrinsic bond orbital (IBO) method.


1 Introduction

Boron hydrides represent a vast chemical space and an astonishing variety of stable structures of various shapes. The shapes are dictated by the formation of three-center two-electron (3c–2e) bonds,1 leading to 3D aromaticity.2,3 Therefore, they differ dramatically from aliphatic hydrocarbons, which only exhibit two-center two-electron (2c–2e) bonds. In aromatic hydrocarbons, the 2c–2e bonds are complemented by π electron clouds, which form the background of 2D aromaticity.

The gas-phase enthalpy of formation (ΔfH°) of a chemical substance is a fundamental observable in chemical sciences. In addition to its use in thermochemistry, ΔfH° has been actively employed in the development, parameterization and evaluation of the methods of electronic structure theory, in particular density functional theory (DFT) and semiempirical quantum mechanical (SQM) methods. For instance, the most prevalent SQM techniques, including the PM6 and PM7 methods, have been parameterized against ΔfH°. Such calibrations are typically, although not exclusively, based on experimental ΔfH° data.4,5 It is noteworthy that even high-level computational methods exhibit greater discrepancies in the reproduction of the ΔfH° of boron hydrides than that of hydrocarbons.6,7 For instance, the mean absolute errors (MAEs) of the B3LYP-D3, ωB97X-D3 and DLPNO-CCSD(T) methods for a comprehensive set of 113 organic molecules have been documented to be 6.0, 3.2 and 1.6 kcal mol−1, respectively.8 For boron hydrides, the reported MAEs for B3LYP-D2, G3, CCSD(T) and G4 were 13.3, 10.0, 6.6 and 3.6 kcal mol−1, respectively.6 This limitation of higher-level computational methods represents a significant obstacle to the advancement of approximate techniques for the accurate representation of boron hydrides. It is thus evident that the benchmark computational methods capable of reproducing the experimental ΔfH° of boron hydrides are a prerequisite for the further development of more reliable approximate methods. The lack of computational methods applicable to molecular systems containing boron atoms poses a significant challenge when dealing with extended molecular systems. This limitation is frequently addressed through the implementation of simplistic solutions, such as the substitution of boron atoms with carbon atoms in protein–ligand docking of boron-containing compounds.9,10

In this study, we have evaluated the efficacy of computational methods in reproducing the experimental ΔfH° of boron hydrides using the atomization energy approach and proposed a novel benchmark methodology for these calculations, which can be employed for generating reference data for the development of more approximate methods. Furthermore, we sought to gain deeper insights into the bonding of boron hydrides by applying the intrinsic bond orbitals (IBOs) bonding analysis to a series of binary boron hydrides and also to electron-precise (with a suffucient number of electrons to form conventional electron-pair bonds between the linked atoms) boron compounds.

It is noteworthy that boron hydrides have garnered significant interest in thermochemistry as a potential rocket fuel.11 There are experimental ΔfH° for eight boron-hydride compounds documented in the literature, as presented in Table 1. For the archetypal boron hydride, diborane B2H6, the JANAF tables provide a ΔfH° value of 9.8 kcal mol−1.12 This value is in close agreement with those determined by Gunn (8.6 kcal mol−1)13 and by Prosen et al.14 (7.5 kcal mol−1, derived from 6.7 kcal mol−1 for conversion to amorphous B and applying the 0.4 kcal mol−1 estimate for amorphous B15). Gunn and Green have reported a lower value of 5.8 kcal mol−1, determined by explosive decomposition.16 We use the average value of 7.9 kcal mol−1 for B2H6 in this study (see Table 1). Gunn and Green have determined the ΔfH° of B4H10, B5H9, B5H11 and B6H10 by explosion of mixtures with SbH3 (the ΔfH° of 13.8, 15.4, 22.2 and 19.6 kcal mol−1, respectively).17 Prosen et al. have determined a ΔfH° value of 15.0 kcal mol−1 for B5H9,14,15 and the JANAF tables give the ΔfH° value of 17.5 kcal mol−1 for B5H9.12 We utilize the average value of 16.0 kcal mol−1 for B5H9 (Table 1). Gunn and Kindsvater have measured the ΔfH° of B6H12, B10H14 and B10H16 calorimetrically by explosion and obtained the ΔfH° of 26.5, 4.4 and 34.8 kcal mol−1, respectively.18 Johnson et al.19 and Galchenko et al.20 have studied B10H14 by pyrolysis methods, and when combined with the heat of sublimation by Miller,21 similar ΔfH° values of 2.5 and 4.3 kcal mol−1 were obtained. However, the JANAF tables give a higher value of 11.3 kcal mol−1.12 Because of this discrepancy, we did not include B10H14 in the evaluation of calculation methods. In order to assess the impact of multicenter bonding on computational estimation more accurately, we also utilized eight electron-precise boron compounds (B2Cl4, B2F4, B2O3, B3N3H6, BCl3, BF3, HO-BO and BO3H3), for which experimental ΔfH° data are available (see Table 1).

Table 1 The experimental gas-phase enthalpy of formation (ΔfH°) of boron compounds in kcal mol−1. The molecular structures of boron hydrides are shown in Fig. 1
Molecule Exp. ΔfH° Avg. exp. ΔfH°
Boron hydrides
B2H6 9.8,12 8.6,13 7.5,14,15 5.8[thin space (1/6-em)]16 7.9 ± 1.7
B4H10 13.8[thin space (1/6-em)]17 13.8
B5H9 15.4,17 15.0,14 17.5[thin space (1/6-em)]12 16.0 ± 1.3
B5H11 22.2[thin space (1/6-em)]17 22.2
B6H10 19.6[thin space (1/6-em)]17 19.6
B6H12 26.5[thin space (1/6-em)]18 26.5
B10H14 11.3,12 4.4,18 2.5,19,21 4.3[thin space (1/6-em)]20,21 5.6 ± 3.9
B10H16 34.8[thin space (1/6-em)]18 34.8
Other boron compounds
B2Cl4 −116.9[thin space (1/6-em)]12 −116.9
B2F4 −342.2[thin space (1/6-em)]12 −342.2
B2O3 −199.8[thin space (1/6-em)]12 −199.8
B3N3H6 −121.9,12 −124.1[thin space (1/6-em)]22 −123.0 ± 1.6
BCl3 −96.3,12 −97.5[thin space (1/6-em)]23 −96.9 ± 0.9
BF3 −269.7,13 −271.6[thin space (1/6-em)]12 −270.7 ± 1.3
HO-BO −134.012 −134.0
BO3H3 −237.212 −237.2



image file: d4dt02589j-f1.tif
Fig. 1 Molecular structures of boron hydrides with experimentally determined ΔfH°.

2 Methods

All the compounds were optimized using the B3LYP/DZVP-DFT (the DZVP-DFT basis set24 is available through the basis set exchange repository, https://www.basissetexchange.org, as dgauss-DZVP) and MP2/cc-pVDZ levels and the LBFGS algorithm with strict optimization criteria, namely the energy change below 0.0006 kcal mol−1, the largest gradient component below 0.12 kcal mol−1 Å−1 and the root-mean-square gradient below 0.06 kcal mol−1 Å−1. Cuby425,26 was utilized for gradient optimizations. The Cuby4 program used Turbomole 7.327 for the MP2 and B3LYP calculations.

The ΔfH° values were computed using the atomization energy approach as described in ref. 28. We utilized the Cuby4 framework.25,26 The Cuby4 program used Turbomole 7.327 for the CCSD(T), MP2, B3LYP and BLYP calculations. For the ωB97X calculations, Cuby4 utilized ORCA 5.0.3.29 Harmonic vibrational calculations providing the ZPVE and other thermodynamic contributions necessary for the evaluation of ΔfH° were performed at the same level as gradient optimizations, namely at the B3LYP/DZVP-DFT and MP2/cc-pVDZ levels. The energy calculations of the studied molecules and their constituent atoms were performed at the all-electron CCSD(T) and various DFT levels. The DFT calculations were carried out in combination with the def2-QZVP basis set. The impact of empirical dispersion (D3)30,31 with Becke–Johnson (BJ) and zero (0) damping functions on DFT was tested. Unless otherwise stated, the BJ damping function was used as the default option. In the case of CCSD(T), we utilized the cc-pVQZ basis set. In the method abbreviated as CCSD(T)′, we used cc-pwCVQZ32 for boron atoms and cc-pVQZ for all the other elements (H, F, Cl, N, and O) and abbreviated it as CCSD(T)′. In the case of larger molecules, namely B8H12, B9H15, B10H14, B10H16 and B20H16, a composite approach was employed to compute CCSD(T). The CCSD was evaluated with cc-pwCVQZ and cc-pVQZ basis sets, while the perturbative triples (T) were evaluated using the smaller cc-pwCVTZ and cc-pVTZ basis sets.

The intrinsic bond orbitals (IBOs) method was used to link the quantitative SCF wave functions to a qualitative chemical picture.33 In other words, a projection operator projected IBO orbitals from the canonical ones. The IBOview program34 was utilized for this purpose. The corresponding input files for the latter were generated at the B3LYP/def2-TZVP level using the Turbomole 7.327 program package.

3 Results and discussion

3.1 Bonding properties

As far as the electronic structure is concerned, the compounds listed in Table 1 have been investigated in terms of the IBOs.33 The bonding in the electron-precise boron compounds is strictly dominated by 2c–2e bonds only. In the case of cyclic borazine, the 2c–2e bonds are complemented by pseudo 3c–2e bonds, which form the classical π electron clouds (the corresponding expansion coefficients, ECs: N 1.51, B 0.23 and B 0.23). Bonding properties in borazine are of 3c–2e-like in terms of the multicenter bonding approach. As a consequence, these two 3c–2e-like bonds form “π-bonds” (see Fig. 2, the violet vs. pink regions visualize the different phases). However, the wave function of this kind changes its sign in contrast to genuine 3c–2e bond appeared in the polyhdral boron clusters in which these orbitals have exclusively one sign.
image file: d4dt02589j-f2.tif
Fig. 2 Examples of bonding as revealed in electron-precise boron compounds by the application of the IBO computational approach. Color coding: pink – 2c–2e, violet – 3c–2e.

In the case of boron hydrides, there used to be a concept of two kinds of 3c–2e bonds, namely central and open. The structures of B2H6, B4H10, B5H11 and B6H10 were well described using open three-center bonds.35 However, the bonding schemes in B5H9 and B10H14 required a resonance hybrid of four valence structures and 24 valence structures, respectively, when open 3c–2e bonds were omitted. We have made efforts to unify these concepts by analyzing the bonding properties using the above approach, which has given rise to not only three- but also four- and even five-center bonding. The four-center bonding concept was found during the bonding analysis of a series of heteroboranes with the EB11 (E = S, Se, Te) closo-icosahedral core.36–38 The results have demonstrated that all the boron hydrides with bridging and terminal hydrogens have 3c–2e BHB and 2c–2e BH bonds, respectively. They are exemplified for the case of B2H6. The rest of the bonds in the remaining seven cases consist of a mixture of 3c–2e and 4c–2e bonds. In addition, the bond description of B5H9 even appears to have a 5c–2e bond (see Fig. 3), a motif that has been earlier observed by Lipscomb by Extended Huckel Theory (EHT).39 The corresponding ECs in this IBO have been computed as B 0.79, B 0.74, B 0.16, B 0.16 and B 0.11. B4H10 serves as an example of the 4c-2e bonding scheme (EC: B 0.80, B 0.80, B 0.18 and B 0.18). The most commonly used B10H14 has such bonding schemes with the ECs of B 0.74, B 0.71, B 0.22 and B 0.22. Interestingly, a classical 2c–2e B–B bond has been identified in the conjuncto B10H16 system (Fig. 3).


image file: d4dt02589j-f3.tif
Fig. 3 Examples of bonding as revealed in boron hydrides by the application of the IBO computational approach. Color coding: pink – 2c–2e, violet – 3c–2e, yellow – 4c–2e, green – 5c–2e.

Furthermore, the number of multicenter bonds present in the molecules under study also reflects their stability, as indicated by the ΔfH°. It is evident that an increase in the number of multicenter bonds results in enhanced system stability with respect to this thermochemical parameter. Decaborane(10), B10H14, a fundamental component of boron chemistry, is distinguished by its exceptional stability among binary boron hydrides, which is attributable to its nine multicenter bonds (six 3c–2e BBB and three 4c–2e BBBB). In contrast, B5H11 and B6H12 have only two multicenter bonds (one 3c–2e BBB and one 4c–2e BBBB) and three multicenter bonds (two 3c–2e BBB and one 4c–2e BBBB), respectively, which is indicative of a lack of stability, consistent with the observed heats of formation.

3.2 Thermochemical properties

Whereas in the previous section, we evaluated the studied molecules in terms of the nature of the bonds, in this section, we are going to focus on the energetic contributions of these motifs when reproducing experimental ΔfH° as precisely as possible. First, we evaluated the correctness of the optimized geometries employed for the ΔfH° calculations. Both methods utilized for the optimizations, namely MP2/cc-pVDZ and B3LYP/DZVP-DFT, were in good agreement with the experimental structures of B2H6, B4H10, B5H11 and B6H12 determined by electron diffraction,40–42 with the MAEs in the bond lengths being 0.015 and 0.020 Å, respectively (see Table S1). We have also tried a larger basis set for MP2 calculations, specifically cc-pVTZ, but the obtained geometries had a larger MAE of 0.023 Å.

In the case of electron-precise boron compounds, the ΔfH° computed by all the tested methods was in good agreement with the experimental values, with the Pearson coefficient R being higher than 0.995 (see Table 2 and Fig. 4). The MAEs of the ΔfH° of the DFT-based methods were between 2.9 and 8.3 kcal mol−1, with the best performance among the DFT methods achieved by ωB97X-D3-0//B3LYP. The empirical dispersion D3 did not dramatically change the MAE in any case. CCSD(T)//MP2 and CCSD(T)′//MP2, with the MAE for the ΔfH° being 2.2 and 1.9 kcal mol−1, respectively, approached the experimental accuracy. A comparison of the MAE obtained for the individual molecules across all the evaluated computational methods, as presented in Table 3, revealed that B3N3H6 was the most challenging electron-precise boron compound for the computational reproduction of ΔfH°. This is consistent with the established difficulty of B3N3H6 in this context.43 Surprisingly, another molecule that is known to be challenging, B2F4,44,45 had the MAE comparable to other electron-precise compounds.


image file: d4dt02589j-f4.tif
Fig. 4 The computed gas-phase enthalpy of formation (ΔfH°) for (A) boron hydrides (B2H6, B4H10, B5H9, B5H11, B6H10, B6H12 and B10H16) and (B) other boron compounds (B2Cl4, B2F4, B2O3, B3N3H6, BCl3, BF3, HO-BO and BO3H3) plotted against the experimental ΔfH°. All in kcal mol−1.
Table 2 The mean absolute error (MAE) and Pearson correlation coefficient (R) of the computed gas-phase enthalpy of formation (ΔfH°) for boron hydrides (B2H6, B4H10, B5H9, B5H11, B6H10, B6H12 and B10H16) and other boron compounds (B2Cl4, B2F4, B2O3, B3N3H6, BCl3, BF3, HO-BO and BO3H3) in kcal mol−1. The CCSD(T)//MP2 approach entails the utilization of CCSD(T) for the evaluation of energy, whereas MP2 is employed for the determination of gradients and the calculation of other thermodynamic contributions, which are essential for the evaluation of ΔfH°
Method MAE R MAE R
  Boron hydrides Other boron compounds
CCSD(T)//MP2 12.1 0.71 2.2 0.999
CCSD(T)′//MP2 1.4 0.99 1.9 0.999
B3LYP 3.2 0.98 4.6 0.998
B3LYP-D3//B3LYP 23.0 −0.12 3.4 0.997
B3LYP-D3-0//B3LYP 11.9 0.91 4.2 0.998
BLYP//B3LYP 14.8 0.98 7.2 0.996
BLYP-D3//B3LYP 9.8 0.58 8.3 0.996
BLYP-D3-0//B3LYP 4.0 0.98 7.7 0.997
ωB97X-D3BJ//B3LYP 23.5 −0.58 5.1 0.998
ωB97X-D3-0//B3LYP 17.1 −0.06 2.9 0.999
ωB97X-V//B3LYP 19.7 −0.26 4.9 0.999


Table 3 The mean absolute error (MAE) of the computed gas-phase enthalpy of formation (ΔfH°) divided by molecular weight (M). ΔfH° M−1 in kcal g−1
Molecule MAE Molecule MAE
Boron hydrides Other boron compounds
B2H6 0.17 B2Cl4 0.04
B4H10 0.18 B2F4 0.05
B5H9 0.20 B2O3 0.07
B5H11 0.18 B3N3H6 0.15
B6H10 0.16 BCl3 0.04
B6H12 0.17 BF3 0.03
B10H16 0.21 HO-BO 0.09
BO3H3 0.04


In the case of boron hydrides, all compounds were computed with a similar average error (Table 3), but the performance of the tested methods exhibited notable discrepancies. DFT-D3 methods combined with the BJ damping function had a large MAE in the range of 9.8 to 23.5 kcal mol−1 and Pearson correlation in the range of −0.58 to 0.58. The D3 dispersion was excessively attractive, resulting in too low ΔfH°, Table 2 and Fig. 4. Using the zero-damping function instead of BJ improved the results, with MAE in the range of 4.0 to 17.1 kcal mol−1 and Pearson correlation in the range of −0.06 to 0.98. In agreement with our results, Becke has recently found that BJ damping may give too large (in absolute value) intramolecular dispersion energies in some alkaline clusters, in particular in Li8 and Na8 clusters. It has been shown that it is due to the too-small effective van der Wals interatomic separation (RvdW,ij) that is linearly related to a critical interatomic separations (Rc,ij).46

CCSD(T)//MP2 had surprisingly large MAE for boron hydrides of 12.1 kcal mol−1. In contrast, CCSD(T)′//MP2 had a small MAE of 1.4 kcal mol−1, being thus close to the experimental accuracy also for this class of compounds. In analogy with our observations, the incorporation of core-correlation functions, as exemplified by the cc-pwCVnZ basis sets, also led to an improvement in the computational description of ΔfH° in complexes comprising alkali and alkaline earth metals.47 We propose CCSD(T)′//MP2 as the benchmark method for the computational evaluation of ΔfH° data for the boron compounds for which no experimental determinations are available or in cases where there are discrepancies between several experimental measurements. Consequently, we have applied this method to B8H12, B9H15, B10H14 and B20H16 (see Table 4 and Fig. 1, 5). The computed ΔfH° of −2.5 kcal mol−1 for B10H14 thus preferred, among several experimental values, the smallest one of 2.5 kcal mol−1, reported by Prosen et al.19 Although the X-ray crystal structures of B8H12 and B9H15 are known,48,49 no experimental ΔfH° values are available. Nevertheless, these ΔfH° values have been estimated empirically and calculated ab initio many times in the literature.6,50–53 The ranges of the theoretical values reported for B8H12 and B9H15 are considerable, spanning from 11.1 to 42.0 and from 26.5 to 49.4 kcal mol−1, respectively. The CCSD(T)′//MP2 gives values of 21.5 and 14.7 kcal mol−1, respectively. The B20H16 molecule is the first synthesized closo macropolyhedron.54 Whereas the crystal structure of B20H16 is known,55 there are no experimental or computational estimates of the ΔfH° of B20H16 in the literature. The CCSD(T)′//MP2 method yields a ΔfH° value of −33.8 kcal mol−1 for B20H16, which is a highly favorable result even when compared to that of B10H14. In accordance with these findings, the pyrolysis of B2H6 at lower temperatures results in the formation of B10H14, while higher temperatures lead to the production of B20H16. This confirms that B20H16 exhibits relatively higher thermodynamic stability.56 Additionally, in accordance with these findings, the IBO analysis has shown besides sixteen 2c–2e terminal B–H bonds also thirty multicenter bonds. Specifically, they are fourteen 3c–2e B–B–B and sixteen 4c–2e B–B–B–B bonds, which is approximately three times the number observed in the previously discussed B10H14.


image file: d4dt02589j-f5.tif
Fig. 5 Molecular diagrams of the B8H12, B9H15 and B20H16 boron hydrides.
Table 4 The computed gas-phase enthalpy of formation (ΔfH°) in kcal mol−1
  ΔfH°
Molecule CCSD(T)′//MP2
B8H12 21.5
B9H15 14.7
B10H14 −2.5
B20H16 −33.8


4 Conclusions

This paper investigates the accuracy of various computational methods for reproducing the available experimental ΔfH° data for boron compounds. Our findings demonstrate that CCSD(T) yields reliable ΔfH° values for boron hydrides only in conjunction with a correlation-consistent basis set optimized for a reliable description of core-valence correlation effects (cc-pwCVQZ) on boron atoms. Additionally, we have demonstrated that empirical dispersion can impair the results of DFT-based methods, particularly when employed in conjunction with the BJ damping function.

The IBO method has been utilized to establish a correlation between quantitative SCF wave functions and a qualitative chemical representation. In addition to confirming the existence of 2c–2e and 3c–2e bonds, our findings have also revealed the existence of 4c–2e and 5c–2e bonds. The 5c–2e bond, a previously reported configuration of a five-centered σ molecular orbital, has been identified in B5H9.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

J.F., D.H. and J.H. acknowledge the support of the Czech Science Foundation, project number 23-05083S. J.Ř. acknowledges the support of the Czech Science Foundation, project number 22-17063S.

References

  1. P. Melichar, D. Hnyk and J. Fanfrlik, A systematic examination of classical and multi-center bonding in heteroborane clusters, Phys. Chem. Chem. Phys., 2018, 20, 4666–4675 RSC.
  2. J. Poater, C. Vinas, I. Bennour, S. Escayola, M. Sola and F. Teixidor, Too Persistent to Give Up: Aromaticity in Boron Clusters Survives Radical Structural Changes, J. Am. Chem. Soc., 2020, 142, 9396–9407 CrossRef CAS PubMed.
  3. J. Poater, S. Escayola, A. Poater, F. Teixidor, H. Ottosson, C. Vinas and M. Sola, Single-Not Double-3D-Aromaticity in an Oxidized Closo Icosahedral Dodecaiodo-Dodecaborate Cluster, J. Am. Chem. Soc., 2023, 145, 22527–22538 CrossRef CAS.
  4. J. Stewart, Optimization of parameters for semiempirical methods V: Modification of NDDO approximations and application to 70 elements, J. Mol. Model., 2007, 13, 1173–1213 CrossRef CAS.
  5. J. Stewart, Optimization of parameters for semiempirical methods VI: more modifications to the NDDO approximations and re-optimization of parameters, J. Mol. Model., 2013, 19, 1–32 CrossRef CAS.
  6. D. Sethio, L. M. L. Daku, H. Hagemann and E. Kraka, Quantitative Assessment of B-B-B, B-Hb-B, and B-Ht Bonds: From BH3 to B12H122-, ChemPhysChem, 2019, 20, 1967–1977 CrossRef CAS PubMed.
  7. M. T. Nguyen, M. H. Matus and D. A. Dixon, Heats of formation of boron hydride anions and dianions and their ammonium salts y with y=1-2, Inorg. Chem., 2007, 46, 7561–7570 CrossRef CAS.
  8. Y. Minenkov, H. Wang, Z. Wang, S. M. Sarathy and L. Cavallo, Heats of Formation of Medium-Sized Organic Compounds from Contemporary Electronic Structure Methods, J. Chem. Theory Comput., 2017, 13, 3537–3560 CrossRef CAS.
  9. G. Minibaeva, A. Ivanova and P. Polishchuk, EasyDock: customizable and scalable docking tool, J. Cheminf., 2023, 15, 102 CAS.
  10. M. Calvaresi and F. Zerbetto, In Silico Carborane Docking to Proteins and Potential Drug Targets, J. Chem. Inf. Model., 2011, 51, 1882–1896 CrossRef CAS PubMed.
  11. C. Yu and S. Bauer, Thermochemistry of the boranes, J. Phys. Chem. Ref. Data, 1998, 27, 807–835 CrossRef CAS.
  12. M. Chase, C. Davies, J. Downey, D. Frurip, R. McDonald and A. Syverud, NIST-JANAF Thermochemical Tables - 4th Edition, J. Phys. Chem. Ref. Data, 1985, 14, 1–1951 CrossRef CAS.
  13. S. Gunn, Heats of Reaction of Boron Trifluoride with HF.3.75H2O and of Diborane with Trimethylamine. Correlation of Thermochemical Data for Some Boron Compounds, J. Phys. Chem., 1965, 69, 1010–1015 CrossRef CAS.
  14. E. Prosen, W. Johnson and F. Pergiel, Heats of Formation of Diborane and Pentaborane, J. Res. Natl. Bur. Stand., 1958, 61, 247–250 CrossRef CAS.
  15. F. Rossini, D. Wagman, W. Evans, S. Levine and I. Jaffe, Selected Values of Chemical Thermodynamic Properties, Circular of the National Bureau of Standards, 1952, vol. 500, year Search PubMed.
  16. S. Gunn and L. Green, Heats of Formation of Some Unstable Gaseous Hydrides, J. Phys. Chem., 1961, 65, 779–783 CrossRef CAS.
  17. S. Gunn and L. Green, Heats if Descomposition of Some Higher Boron Hydrides, J. Phys. Chem., 1961, 65, 2173–2175 CrossRef CAS.
  18. S. Gunn and J. Kindsvater, Heats Decomposition of Some More Boron Hydrides, J. Phys. Chem., 1966, 70, 1114–1119 CrossRef CAS.
  19. W. Johnson, M. Kilday and E. Prosen, Heat of Formation of Decaborane, J. Res. Natl. Bur. Stand. A Phys. Chem., 1960, 64, 521–525 CrossRef.
  20. G. Galchenko, S. Skuratov and B. Timofeev, Determination of Heat of Formation of Decaborane, Dokl. Akad. Nauk SSSR, 1962, 142, 1077 CAS.
  21. G. Miller, Vapor Pressure of Solid Decaborane, J. Phys. Chem., 1963, 67, 1363–1364 CrossRef CAS.
  22. M. Kilday, W. Johnson and E. Prosen, Heat of Combustion of Borazine, B3N3H6, J. Res. Natl. Bur. Stand. A Phys. Chem., 1961, A65, 101–104 CrossRef.
  23. W. Johnson, R. Miller and E. Prosen, Heat of Formation of Boron Trichloride, J. Res. Natl. Bur. Stand., 1959, 62, 213–217 CrossRef CAS.
  24. N. Godbout, D. Salahub, J. Andzelm and E. Wimmer, Optimization of Gaussian-type Basis-sets for Local Spin-density Functional Calculations. 1. Boron Through Neon, Optimization Technique and Validation, Can. J. Chem., 1992, 70, 560–571 CrossRef CAS.
  25. J. Řezáč, Cuby – ruby framework for computational chemistry, https://cuby4.molecular.cz.
  26. J. Rezac, Cuby: An Integrative Framework for Computational Chemistry, J. Comput. Chem., 2016, 37, 1230–1237 CrossRef CAS PubMed.
  27. R. Ahlrichs, M. Bar, M. Haser, H. Horn and C. Kolmel, Electronic-Structure Calculations on Workstation Computers - The Program System Turbomole, Chem. Phys. Lett., 1989, 162, 165–169 CrossRef CAS.
  28. J. W. Ochterski, Thermochemistry in Gaussian, 2000, Gaussian, Inc., 2000, pp. 1–19 Search PubMed.
  29. F. Neese, Software update: The ORCA program system-Version 5.0, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2022, 12, e1606 Search PubMed.
  30. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem. Phys., 2010, 132, 154104 CrossRef.
  31. S. Grimme, S. Ehrlich and L. Goerigk, Effect of the Damping Function in Dispersion Corrected Density Functional Theory, J. Comput. Chem., 2011, 32, 1456–1465 CrossRef CAS PubMed.
  32. K. Peterson and T. Dunning, 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 CrossRef CAS.
  33. G. Knizia, Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical Concepts, J. Chem. Theory Comput., 2013, 9, 4834–4843 CrossRef CAS.
  34. G. Knizia, IboView, https://www.iboview.org.
  35. W. Lipscomb, Boron Hydrides, Benjamin, WA, New York, 1963 Search PubMed.
  36. W. Keller, J. Ballmann, M. Sarosi, J. Fanfrlik and D. Hnyk, The Telluraboranes closo-TeB5Cl5 and closo-TeB11Cl11 with Exceptionally Long Body Diagonals: Synthetic and Bonding Motifs for Innovative Octahedral and Icosahedral Geometries, Angew. Chem., Int. Ed., 2023, 62, e202219018 CrossRef CAS PubMed.
  37. W. Keller, F. Lissner, J. Ballmann, J. Fanfrlik, D. Hnyk and T. Schleid, A Direct Route to closo-SBnCln Thiaboranes from Simple Electron-Precise Synthons: The Different Role of Chalcogen Bonding in SB5Cl5 and SB11Cl11 Crystals, Angew. Chem., Int. Ed., 2024, e202406751 CAS.
  38. W. Keller, M. Hofmann, H. Wadepohl, M. Enders, J. Fanfrlik and D. Hnyk, Chlorinated polyhedral selenaboranes revisited by joint experimental/computational efforts: the formation of closo-1-SeB9Cl9 and the crystal structure of closo-SeB11Cl11, Dalton Trans., 2023, 52, 16886–16893 RSC.
  39. W. Lipscomb, The Boranes and Thei Relatives, Nobel Lecture, December 11, 1976.
  40. K. Hedberg and V. Schomaker, A Reinvestigation of the Structures of Diborane and Ethane by Electron Diffraction, J. Am. Chem. Soc., 1951, 73, 1482–1487 CrossRef CAS.
  41. P. Brain, C. Morrison, S. Parsons and D. Rankin, Tetraborane(10), B4H10: Structures in gaseous and crystalline phases, J. Chem. Soc., Dalton Trans., 1996, 4589–4596 RSC.
  42. P. Brain, D. Hnyk, D. Rankin, M. Buhl and P. Schleyer, The Molecular-Structures of Pentaborane(11), B5H11, and Hexaborane(12), B6H12, in the Gas-Phase as Determined by Electron-Diffraction and Ab-Initio Calculations, Polyhedron, 1994, 13, 1453–1466 CrossRef CAS.
  43. M. H. Matus, K. D. Anderson, D. M. Camaioni, S. T. Autrey and D. A. Dixon, Reliable predictions of the thermochemistry of boron-nitrogen hydrogen storage compounds: BxNxHy, x=2, 3, J. Phys. Chem. A, 2007, 111, 4411–4421 CrossRef CAS PubMed.
  44. B. Chan, A. J. Trevitt, S. J. Blanksby and L. Radom, Comment on the Ionization Energy of B2F4, J. Phys. Chem. A, 2012, 116, 9214–9215 CrossRef CAS.
  45. Z. Li and K. Fan, B2F4 molecule: A challenge for theoretical calculations, J. Phys. Chem. A, 2002, 106, 6659–6664 CrossRef CAS.
  46. A. Becke, Density-functional theory vs density-functional fits, J. Chem. Phys., 2022, 156, 214101 CrossRef CAS PubMed.
  47. M. Sullivan, M. Iron, P. Redfern, J. Martin, L. Curtiss and L. Radom, Heats of formation of alkali metal and alkaline earth metal oxides and hydroxides: Surprisingly demanding targets for high-level ab initio procedures, J. Phys. Chem. A, 2003, 107, 5617–5630 CrossRef CAS.
  48. R. Dickerson, P. Wheatley, P. Howell and W. Lipscomb, Crystal and Molecular Structure of B9H15, J. Chem. Phys., 1957, 27, 200–209 CrossRef CAS.
  49. R. Enrione, F. Boer and W. Lipscomb, Octaborane(12), Inorg. Chem., 1964, 3, 1659–1666 CrossRef CAS.
  50. M. McKee, Estimation of Heats of Formation of Boron Hydrides From Ab Initio Energies, J. Phys. Chem., 1990, 94, 435–440 CrossRef CAS.
  51. D. Laurie and P. Perkins, Direct Calculation of the Heats of Formation of Some Boron Hydrides, Inorg. Chim. Acta, 1982, 63, 53–56 CrossRef CAS.
  52. C. Housecroft and K. Wade, Bond Length-Based Bond Enthalpies for Nido and Arachno Boranes BNHN+4 and BNHN+6, Inorg. Nucl. Chem. Lett., 1979, 15, 339–342 CrossRef CAS.
  53. W. Herndon and M. Ellzey, Resonance Eenergies and Delta-H(Atomization) for Boron Hydrides, Inorg. Nucl. Chem. Lett., 1980, 16, 361–366 CrossRef CAS.
  54. N. Miller and E. Muetterties, A New Boron Hydride, B20H16, J. Am. Chem. Soc., 1963, 85, 3506 CrossRef CAS.
  55. R. Dobrott, L. Friedman and W. Lipscomb, Molecular + Crystal Structure of B20H16, J. Chem. Phys., 1964, 40, 866–872 CrossRef CAS.
  56. D. Hnyk, J. Holub, T. Jelinek, J. Machacek and M. Londesborough, Revisitiong B20H16 by Means of a Joint Computational/experimental NMR Aproach, Collect. Czech. Chem. Commun., 2010, 75, 1115–1123 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: A comparison of experimental and computed bond lengths in B2H6, B4H10, B5H11 and B6H12, a summary of computed ΔfH° values and the coordinates for the optimized structures of the studied molecules are provided. See DOI: https://doi.org/10.1039/d4dt02589j

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.