Next Article in Journal
The Intestinal Efflux Transporter Inhibition Activity of Xanthones from Mangosteen Pericarp: An In Silico, In Vitro and Ex Vivo Approach
Next Article in Special Issue
Pressure-Dependent Structure of Methanol–Water Mixtures up to 1.2 GPa: Neutron Diffraction Experiments and Molecular Dynamics Simulations
Previous Article in Journal
Therapeutic Effects of Gallic Acid in Regulating Senescence and Diabetes; an In Vitro Study
Previous Article in Special Issue
Strong Affinity of Triazolium-Appended Dipyrromethenes (TADs) for BF4
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Importance of Strain (Preorganization) in Beryllium Bonds

1
Instituto de Química Médica (CSIC), Juan de la Cierva, 3, E-28006 Madrid, Spain
2
Instituto de Química-Física “Rocasolano” (CSIC), Serrano, 119, E-28006 Madrid, Spain
3
Departamento de Química, Facultad de Ciencias, Módulo 13, and Institute of Advanced Chemical Sciences (IadChem), Campus de Excelencia UAM-CSIC, Universidad Autónoma de Madrid, Cantoblanco, E-28049 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Submission received: 23 November 2020 / Revised: 7 December 2020 / Accepted: 8 December 2020 / Published: 11 December 2020
(This article belongs to the Special Issue Non-Covalent Interaction in Solids and Large Clusters)

Abstract

:
In order to explore the angular strain role on the ability of Be to form strong beryllium bonds, a theoretical study of the complexes of four beryllium derivatives of ortho closo-carboranes with eight molecules (CO, N2, NCH, CNH, OH2, SH2, NH3, and PH3) acting as Lewis bases has been carried out at the G4 computational level. The results for these complexes, which contain besides Be other electron-deficient elements, such as B, have been compared with the analogous ones formed by three beryllium salts (BeCl2, CO3Be and SO4Be) with the same set of Lewis bases. The results show the presence of large and positive values of the electrostatic potential associated to the beryllium atoms in the isolated four beryllium derivatives of ortho-carboranes, evidencing an intrinsically strong acidic nature. In addition, the LUMO orbital in these systems is also associated to the beryllium atom. These features led to short intermolecular distances and large dissociation energies in the complexes of the beryllium derivatives of ortho-carboranes with the Lewis bases. Notably, as a consequence of the special framework provided by the ortho-carboranes, some of these dissociation energies are larger than the corresponding beryllium bonds in the already strongly bound SO4Be complexes, in particular for N2 and CO bases. The localized molecular orbital energy decomposition analysis (LMOEDA) shows that among the attractive terms associated with the dissociation energy, the electrostatic term is the most important one, except for the complexes with the two previously mentioned weakest bases (N2 and CO), where the polarization term dominates. Hence, these results contribute to further confirm the importance of bending on the beryllium environment leading to strong interactions through the formation of beryllium bonds.

Graphical Abstract

1. Introduction

In the last years, the importance of non-covalent interactions (NCI) in a number of fields has been highlighted. Parallel to these activities, new types of NCIs have been described and named using the Periodic Table column of the atom of the Lewis acid involved in the interaction [1,2]. Thus, aerogen [3], halogen [4,5], chalcogen [6,7,8,9], tetrel [10,11], triel [12,13], spodium [14], regium or coinage-metal [15,16,17,18], alkaline-earth [19,20], and alkaline bonds [21] have been described, which are associated with atoms of columns 18-11 and 2-1, respectively. Beryllium bonds [22,23] are a subfamily of the alkaline-earth NCIs that show very strong binding energies and profound changes in the geometric and electronic properties of the systems involved.
Icosahedral closo-heteroboranes are a family of compounds with 12 atoms forming the central framework. The parent compound is the very stable dianion (B12H12)2−. Derivatives with two carbon atoms in the skeleton are neutral and named closo-carboranes. They have been used from potential scaffolds in the design of new drugs, as a catalyst and for energy storage [24,25]. Among the different closo-carboranes with metallic replacement in the cage, in 1971, Popp and Hawthorne [26] described the single case in the literature with a beryllium atom, 3-beryl-1,2-dicarbacloso-dodecaborane(12). The experimental evidences indicated that this molecule forms complexes with diethylether or trimethylamine in the presence of these bases (Figure 1).
Strained systems have been shown to react more easily than other related non-strained systems. In some cases, this effect could be due to a pre-organization of the system to evolve more easily. For instance, the hydrogen bond donor ability of C(sp3)-H groups increases significantly when they are located in three membered rings [27], the hydrolysis of amides have lower barriers when they are highly twisted [28] and strained beryllium derivatives become excellent H2 acceptors [29,30].
In this article, we study the complexes formed by the particular set of beryl-1,2-dicarbacloso-dodecaboranes, where the metallic replacement introduces a typical electron deficient atom, Be, in a rather strained environment, involving also other electron-deficient atoms, such as B. More precisely, the structure, stability, and bonding characteristics of the complexes between four beryl-dicarbacloso-dodecaborane isomers and eight different Lewis bases (CO, N2, NCH, CNH, OH2, SH2, NH3, and PH3) have been studied at G4 computational level. These compounds are shown in Figure 2, where four different isomers, AD, corresponding to the different position of the C-C linkage within the cage. In addition, the complexes formed between the same set of bases and Cl2Be, CO3Be, and SO4Be salts, where Be is not attached to electron-deficient elements and has different degrees of strain, have been calculated and compared with the results obtained for the carboranes.

2. Computational Methods

The geometry of the isolated molecules and complexes was optimized with the Gaussian-4 (G4) [31] formalism at the B3LYP/6-31G (2df,p) level of theory. Frequency calculations at B3LYP/6-31G (2df,p) level were used to confirm that the complexes correspond to energy minima. Thermochemical corrections were also computed with the G4 formalism and the Gaussian-16 software [32].
The dissociation enthalpies of the complexes were calculated as the sum of the values of the isolated monomers minus that of the complex calculated at G4 level. The deformation energy was obtained as the G4 electronic energy difference between the molecules in the geometry of the complex and in their minima conformation, as isolated molecules.
The molecular electrostatic potential of the isolated molecules was calculated at B3LYP/6-31(2df,p) level, analyzed on the 0.001 au electron density isosurface with the Multiwfn program [33], and represented with the Jmol program [34].
The electron density distribution was analyzed within the quantum theory of Atoms in Molecules (QTAIM) methodology [35,36] and the AIMAll program [37]. The critical points of the electron density correspond to maxima (3,–3) associated to the location of the atom, two types of saddle points, (3,–1) and (3,+1), associated with the presence of bond and ring critical points, respectively, and minima (3,+3), associated to cage critical points.
The localized molecular orbital energy decomposition (LMOEDA) method [38] at B3LYP/6-31G(2df,p) level was carried out in order to analyze the importance of the different energy terms in the interaction energy. Based on this method, the interaction energy can be decomposed in a repulsive term and several attractive terms (electrostatic, exchange, polarization and dispersion energies). The electrostatic term describes the classical Coulomb interaction. The repulsion and exchange terms are associated to Pauli’s exclusion principle. The polarization term is caused by the relaxation of the electronic distribution in the complex. The dispersion term arises from the electron correlation. These calculations were carried out with the GAMESS program [39].

3. Results and Discussion

3.1. Isolated Be-Containing Lewis Acids

In order to understand and rationalize how the different dodecaboranes may interact with the selected set of Lewis bases, we will firstly analyze the isolated dodecaboranes. The four isomers with general formula B9H11C2Be considered in this article show an icosahedral shape (Figure 3). Only isomer A has been described in the literature [26]. The relative enthalpy of the four isomers is listed in Table 1. The most stable isomer D is the one with the maximum distance between the carbon and beryllium atoms, whereas isomer A, which exhibits the shortest distance between those atoms, lies 18 kJ·mol−1 higher in energy (see Table 1). The other two isomers, C and B, show relative energies of 29 and 35 kJ·mol−1, respectively. Thus, no clear relationship is observed between the relative stability of the isomers and the distribution of the atoms in the icosahedral structure of the molecule.
The electrostatic potential on the 0.001 au electron density isosurface of the molecules (see Table 1) shows its maximum value, Vs, max, in the proximity of the beryllium atom (see Figure 4a for isomer A, as a suitable example. Figure S1 of the Supporting Information shows the same information for isomers BD). The largest value is obtained for isomer A (669 kJ·mol−1), in which the Be atom is directly linked to both carbon atoms. The following value (626 kJ·mol−1) corresponds to isomer B, in which Be is directly bonded to only one of the two carbon atoms. The electrostatic potential for isomers C and D is decreasingly smaller. However, for isomer C, this potential is higher than for isomer D, because in the former, two of the B atoms attached to Be are bonded to the C atoms, whereas in isomer D, only one is attached to a C atom. For comparative purposes, the values corresponding to Cl2Be, CO3Be, and SO4Be have also been included in the same table. The values obtained for SO4Be (938 kJ·mol−1) and CO3Be (907 kJ·mol−1) are larger than those obtained for the B9H11C2Be isomers, reflecting the strong electron withdrawing power of the sulfate and carbonate groups, while in the case of Cl2Be (141 kJ·mol−1), the value of Vs, max is very small as compared to the rest of the beryllium derivatives considered here.
As regards to the molecular orbitals, the LUMO orbital is mostly associated to the beryllium atom in all B9H11C2Be isomers (see Figure 4b for isomer A and Figure S2 for isomers AD). It is interesting to note that the energy of the LUMO for all these compounds is negative (see Table S1 of the Supporting Information), in agreement with the electron deficient character of Be. Interestingly, these energies, in absolute value, follow the trend A > B > C > D, which is the same as the trend observed for the values of the Vs, max of MESP (see Table 1). In fact, there is a reasonably good linear correlation between both sets of values (see Figure S3). Thus, from both an electrostatic and orbital perspective, the most acidic regions of these molecules are associated with the location of the beryllium atom, and the intrinsic acidities of these compounds should follow A > B > C > D, which implies that the most stable isomer is the least reactive and the second most stable isomer is the most reactive one. We will see later on that the interaction energies between these four isomers and the different Lewis bases considered in this study follow exactly the same trend, A > B > C > D.
As a measure of the geometrical strain of the beryllium atom in each of the B9H11C2Be isomers, the distance between this atom and the plane defined by the five atoms attached to it has been calculated. Ideally, a beryllium atom in the same plane to those of the surrounding atoms should not have strain, but as the distance to the plane increases, the strain should increase too. More importantly, the calculated distance values (1.030, 1.014, 1.010, and 1.000 Å for isomers AD, respectively) have been found to correlate linearly (R2 = 0.99) with the calculated Vs, max, a clear indication that changes in the strain at the Be center are closely related with its electrophilicity.
The molecular graphs of the four isomers are depicted in Figure 5, and show the presence of only three bond critical points (BCPs) between the Be atom and the surrounding atoms for A, B, and D isomers and five BCPs in C. However, in the latter case, two of the BCPs are in close proximity of an RCP, which indicates the possibility of their disappearance upon some external perturbation. In any case, the Be-C and Be-B BCPs obtained for all the isomers show a small value of the electron density (between 0.079 and 0.075 au), as should be expected for bonds involving electron-deficient atoms, but with negative values of the energy density, which indicates that the interactions have some covalent character. The B-B and the C-C BCPs in these systems show electron densities between 0.123 and 0.109 au for the B-B linkages and around 0.18 au for the C-C ones. For both kinds of bonds, the total energy density is negative, as should be expected for covalent interactions.

3.2. B9H11C2Be: Lewis Base Complexes

The structures of the most stable complexes formed by the interaction of the Lewis bases under consideration with the beryllium atoms of the four B9H11C2Be isomers AD have been located. As an illustration, we show in Figure 6 the molecular graphs of four suitable examples, but a complete information on the structures of these complexes is shown in Table S2 of the Supporting Information.
The molecular graph of the complexes presents a unique BCP between the two interacting molecules. The BCPs are always located closer to the beryllium atoms than to the base, as expected due to the smaller size and more electropositive character of the beryllium. The electron density at this intermolecular BCP ranges from 0.038 to 0.080, but what is more important, for a given Lewis base, the value of the electron density at this BCP (see Table S3) follows the sequence A > B > C > D (except for the CO and CNH complexes), confirming the picture obtained from the energy of the LUMO and the values of the Vs, max, which indicate again that the most stable isomer is the least reactive and the less stable isomer A the most reactive one. Additionally, in all cases, the values of the Laplacian (between 0.09 and 0.57 au) are positive, consistent with a dominant electrostatic interaction.
From the values in Table 2, it can be seen that the intermolecular distances between the Be atom and the interacting atom of the Lewis Base are always very short, pointing to a rather strong interaction. For those cases where the basic site of the Lewis base belongs to the first row of the periodic table (C, N, and O), the interatomic distances are between 1.60 and 1.88 Å, while for the second-row basic sites (S and P), the distances range between 2.11 and 2.28 Å.
Exponential relationships are obtained between the electron densities at the base-Be BCP and the Be-base interatomic distances, as long as the data are grouped based on the nature of the basic center of the Lewis base (Figure S4a) in agreement with other interactions [40,41,42]. In the case of the Laplacian, the correlation with the interatomic distance is independent of the atom interacting with Be (Figure S4b). Importantly, the total energy density is negative for all the complexes with CNH, CO, NH3, SH2, and PH3, indicating a certain covalent character of the interaction. Nonetheless, for all complexes with N2 and OH2 and those of NCH with the beryllium carboranes, positive values of the energy density are obtained.
In order to gain some insight on these findings, we have used the LMOEDA energy partition method, which permits to analyze the contribution of the different energy terms to the dissociation energies (Table S4 and Figure S5). In the AD complexes, the most important term is the repulsion energy, which is overcome by the sum of the attractive terms (electrostatic, exchange, polarization, and dispersion). Among the attractive terms, the two most important ones are the electrostatic and the polarization terms. The electrostatic contribution is dominant and increases with the binding energy for stronger complexes, i.e., (SH2, PH3, NCH, OH2, CNH, and NH3), with a top of 51% for NH3 complexes. The opposite happens with the polarization, which is the most important term in the two weakest complexes (N2 and CO), with contributions between 44 and 42% to the total attractive interaction. The two less important attractive terms, exchange and dispersion, have contributions between 15–21% and 6–9% of the sum of the attractive terms, respectively. The LMOEDA partition results for CO3Be, SO4Be, and Cl2Be complexes follow similar trends, except for an increment of the contribution of the exchange energy in the Cl2Be (between 20 and 24%) and a reduction of the polarization term (between 26 and 36%).
From this analysis, it is easy to justify that complexes with N2 exhibit positive values of the energy density, because as mentioned above, the interaction is dominated by dispersion and polarization interactions. However, a positive energy density for water complexes is unclear.
It is also worth noting that the intermolecular Be-Lewis base distances for the B9H11C2Be-base complexes of isomer A are highly correlated (R2 > 0.99) with those of the other three isomers, BD (Figure S5). The quality of the correlation drops slightly when the values of the complexes with CO3Be and SO4Be are compared with those from the AD isomers (R2 between 0.98 and 0.95) and even more with the Cl2Be complexes (R2 between 0.94 and 0.92). Indeed, although the values in Table 2 clearly show that the distances for the B9H11C2Be-containing complexes are similar to those found for the complexes with CO3Be, SO4Be, and Cl2Be, the differences are not always of the same sign. In all cases, the longest intermolecular distances for a given Lewis base are obtained for the Cl2Be complexes, save for the complex with OH2, with shorter distance as compared to complexes with the B, C, and D isomers, likely due to the effect of secondary interactions. The peculiar behavior observed for the Cl2Be complexes is a direct consequence of the deformation changes upon complex formation. In general, whereas these deformations are small for the Be compounds considered, Cl2Be is a clear exception, because a significant change in the Cl-Be-Cl angle is observed. Indeed, the Cl-Be-Cl angle in the complexes ranges between 135° and 146°, whereas the isolated structure is linear the isolated structure is linear.
The dissociation enthalpies of the complexes are gathered in Table 3, and the values of ΔG are listed in Table S6 of the SI. The dissociation enthalpies range between 6.1 kJ·mol−1 for Cl2Be:N2 to 256.1 kJ·mol−1 for the SO4Be:NH3 complex. For a given Lewis base, the smallest values are always obtained in the BeCl2 complexes, again a consequence, as we will see later, of the large deformation effects in these particular complexes, while the largest ones are found in the complexes with SO4Be or with isomer A. In the isomers of B9H11C2Be, the dissociation enthalpies decrease as the distance between the Be and carbon atoms increases; thus, the largest values for a given base are obtained in the complex with isomer A, while the smallest in those with isomer D. In fact, excellent correlations (R2 > 0.99) are obtained between the strain parameter values calculated for the isolated B9H11C2Be molecules and the dissociation enthalpy for each Lewis base, with the exception of N2 complexes. In all systems studied here for a given Lewis acid, the smallest dissociation enthalpy is always found in the N2 complexes, while the largest one is found in the NH3 complexes.
As for the intermolecular distances, the dissociation enthalpies for the AD complexes are highly correlated among each other (R2 > 0.99, see Figure S7). Again, the correlations are slightly worse when the dissociation enthalpies for the AD complexes are compared with those of the CO3Be- and SO4Be-containing complexes (R2 between 0.98 and 0.92) and even worse when dealing with Cl2Be complexes (R2 between 0.93 and 0.90).
Interestingly, even though the largest values of the Vs, max (Table 1) in the isolated beryllium derivatives correspond to SO4Be followed by CO3Be, the dissociation energies in several cases of isomer A are larger than those for SO4Be and CO3Be, an indication that using only MESP as a binding energy predictor is not enough for these complexes. Although the electrostatic interactions are a strong component of the stabilization of the complexes, other interaction terms, such as polarization and dispersion, play an important role in defining the correct stability trends. Another important issue to be considered is the deformation penalty when the complex is formed. The values obtained are listed in Table 4. Clearly, the complexes of Cl2Be have larger distortion energies (between 19 and 55 kJ·mol−1), which is one of the factors that contributes to the relatively low stabilization enthalpies calculated for such complexes. For the remaining complexes, the deformation energy values are smaller than 13 kJ·mol−1.
The dissociation energy of the complexes has been analyzed using a model proposed originally by Legon and Miller (Equation (1)) [43] that relates these energies with a nucleophilic parameter characterizing the bases, Nb, and an electrophilic parameter characterizing the Lewis acids, Ea [19,44,45]:
D e = c ·   N b · E a
where the c constant has a value of 1.0 kJ·mol−1 to maintain the units of the equation.
In a first attempt, the 56 complexes have been used to obtain 7 Ea’s (because in our analysis we include the four AD isomers of the beryllium ortho-carboranes and the three Be salts) and 8 Nb’s (corresponding to the 8 Lewis base considered) with Equation (2). The 7 LA and 8 LB have been fitted:
D e = c · i = 1 8 x i · N b i · j = 1 7 x j · E a j
where the values of xi and xj are 1.0 when the corresponding Lewis base or Lewis acid is present in the complex, and 0.0 otherwise.
The fitted values of Nb and Ea are listed in Table 5. The results with all the complexes (Model 1) clearly show that the complexes with Cl2Be have the largest deviations between the calculated and fitted values (Figure 7). Removing the Cl2Be complexes (Model 2), the statistics improve (larger R2 and smaller SD). These results indicate that the electrophilicities of the bases follow the order: SO4Be > A > B > CO3Be > C > D > Cl2Be.
The Nb values for the LB in this set of complexes are larger than those derived for complexes of simple beryllium derivatives [19], but they appear in the same order as the ones found here: NH3 > OH2 > NCH > PH3 > CO (only those cases common to the two studies are indicated).

4. Conclusions

A theoretical study of four beryllium derivatives of ortho-carboranes, one of them known experimentally, was carried out at G4 ab initio computational level. The isolated molecules show a positive region of the electrostatic potential close to the beryllium atom and the LUMO orbital in the same region. Consistently, the Be site is the acidic center when these compounds interact with eight molecules (CO, N2, NCH, CNH, OH2, SH2, NH3, and PH3) acting as Lewis bases. In our survey, we have also included complexes of these same bases with three additional beryllium compounds with a different degree of strain (SO4Be, CO3Be, and Cl2Be). In all cases, strong complexes are formed between the beryllium derivatives and the Lewis bases reaching dissociation energies of 241 kJ mol−1, demonstrating the importance of the strain to lead to strong beryllium bonds. Interestingly, the interaction enthalpies of some of the beryllium ortho-carboranes are larger than the ones calculated for complexes with SO4Be. The strain around the beryllium atom has been shown to linearly correlate with the value of the positive region of the electrostatic potential and the dissociation energy values. The LMOEDA energy partition shows that among the attractive terms associated to the dissociation energy, the electrostatic one is the dominant one, being up to 51% of the sum of attractive terms except for the complexes with the two weakest bases (N2 and CO) where the polarization is dominant. The latter cases are specifically those for which ortho-carboranes led to even stronger beryllium bonds than previously reported values.

Supplementary Materials

The following are available online, Figure S1. Molecular electrostatic potential (MESP) on the 0.001 au electron density, Figure S2. LUMO of the B9H11C2Be isomers, Figure S3. Linear correlation between the Vs, max of the MESP and the absolute value of the energy of the LUMO, Figure S4. Electron density properties (ρBCP and ∇2ρBCP) at the intermolecular BCPs (au) vs. the interatomic distance (Å), Figure S5. Graphical representation of the LMOEDA energy terms (kJ mol-1), Figure. S6. Intermolecular distances in A complexes vs. B-D complexes (Å), Figure S7. Dissociation enthalpies of A complexes vs. B-D complexes (kJ mol−1), Table S1. Energy (a.u.) of the LUMO of the B9H11C2Be isomers, Table S2. Molecular graphs and optimized Cartesian coordinates (Å) of the complexes, Table S3. Electron density (au) at the BCP between the Be atom and the Lewis base, Table S4. LMOEDA energy terms (kJ mol−1), Table S5. Dissociation ΔG energies (kJ mol−1).

Author Contributions

I.A. performed the calculations. I.A., J.E., J.M.O.-E., M.Y., O.M., M.M.M.-C. participated in the discussion of the results and contributed to the writing of this paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministerio de Ciencia, Innovación y Universidades of Spain (PGC2018-094644-B-C21, PGC2018-094644-B-C22, and PID2019-110091GB-I00 (MICINN)) and Comunidad de Madrid (P2018/EMT-4329 AIRTEC-CM).

Acknowledgments

Our thanks are also given to the CTI (CSIC) for their continued computational support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alkorta, I.; Elguero, J.; Frontera, A. Not Only Hydrogen Bonds: Other Noncovalent Interactions. Crystals 2020, 10, 180. [Google Scholar] [CrossRef] [Green Version]
  2. Cavallo, G.; Metrangolo, P.; Pilati, T.; Resnati, G.; Terraneo, G. Naming Interactions from the Electrophilic Site. Cryst. Growth Des. 2014, 14, 2697–2702. [Google Scholar] [CrossRef]
  3. Bauzá, A.; Frontera, A. Aerogen Bonding Interaction: A New Supramolecular Force? Angew. Chem. Int. Ed. 2015, 54, 7340–7343. [Google Scholar] [CrossRef] [PubMed]
  4. Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. The Halogen Bond. Chem. Rev. 2016, 116, 2478–2601. [Google Scholar] [CrossRef] [Green Version]
  5. Desiraju, G.R.; Ho, P.S.; Kloo, L.; Legon, A.C.; Marquardt, R.; Metrangolo, P.; Politzer, P.; Resnati, G.; Rissanen, K. Definition of the halogen bond (IUPAC Recommendations 2013). Pure Appl. Chem. 2013, 85, 1711. [Google Scholar] [CrossRef]
  6. Wang, W.; Ji, B.; Zhang, Y. Chalcogen Bond: A Sister Noncovalent Bond to Halogen Bond. J. Phys. Chem. A 2009, 113, 8132–8135. [Google Scholar] [CrossRef]
  7. Sanz, P.; Yáñez, M.; Mó, O. Competition between X···H···Y Intramolecular Hydrogen Bonds and X····Y (X = O, S, and Y = Se, Te) Chalcogen-Chalcogen Interactions. J. Phys. Chem. A 2002, 106, 4661–4668. [Google Scholar] [CrossRef]
  8. Alkorta, I.; Rozas, I.; Elguero, J. Molecular Complexes between Silicon Derivatives and Electron-Rich Groups. J. Phys. Chem. A 2001, 105, 743–749. [Google Scholar] [CrossRef]
  9. Aakeroy, C.B.; Bryce, D.L.; Desiraju, G.R.; Frontera, A.; Legon, A.C.; Nicotra, F.; Rissanen, K.; Scheiner, S.; Terraneo, G.; Metrangolo, P.; et al. Definition of the chalcogen bond (IUPAC Recommendations 2019). Pure Appl. Chem. 2019, 91, 1889. [Google Scholar] [CrossRef]
  10. Grabowski, S.J. Tetrel bond–σ-hole bond as a preliminary stage of the SN2 reaction. Phys. Chem. Chem. Phys. 2014, 16, 1824–1834. [Google Scholar] [CrossRef]
  11. Bauzá, A.; Mooibroek, T.J.; Frontera, A. Tetrel-Bonding Interaction: Rediscovered Supramolecular Force? Angew. Chem. Int. Ed. 2013, 52, 12317–12321. [Google Scholar] [CrossRef] [PubMed]
  12. Grabowski, S.J. Boron and other Triel Lewis Acid Centers: From Hypovalency to Hypervalency. ChemPhysChem 2014, 15, 2985–2993. [Google Scholar] [CrossRef] [PubMed]
  13. Alkorta, I.; Elguero, J.; Del Bene, J.E.; Mó, O.; Yáñez, M. New Insights into Factors Influencing B-N Bonding in X:BH3−nFn and X:BH3−nCln for X = N2, HCN, LiCN, H2CNH, NF3, NH3 and n = 0–3: The Importance of Deformation. Chem. Eur. J. 2010, 16, 11897–11905. [Google Scholar] [CrossRef] [PubMed]
  14. Bauzá, A.; Alkorta, I.; Elguero, J.; Mooibroek, T.J.; Frontera, A. Spodium bonds: Noncovalent interactions involving group 12 of elements. Angew. Chem. Int. Ed. 2020, 59, 17482–17487. [Google Scholar] [CrossRef]
  15. Legon, A.C.; Walker, N.R. What’s in a name? ‘Coinage-metal’ non-covalent bonds and their definition. Phys. Chem. Chem. Phys. 2018, 20, 19332–19338. [Google Scholar] [CrossRef] [Green Version]
  16. Sánchez-Sanz, G.; Trujillo, C.; Alkorta, I.; Elguero, J. Understanding Regium Bonds and their Competition with Hydrogen Bonds in Au2:HX Complexes. ChemPhysChem 2019, 20, 1572–1580. [Google Scholar] [CrossRef]
  17. Zierkiewicz, W.; Michalczyk, M.; Scheiner, S. Regium bonds between Mn clusters (M = Cu, Ag, Au and n = 2–6) and nucleophiles NH3 and HCN. Phys. Chem. Chem. Phys. 2018, 20, 22498–22509. [Google Scholar] [CrossRef] [Green Version]
  18. Stenlid, J.H.; Brinck, T. Extending the σ-Hole Concept to Metals: An Electrostatic Interpretation of the Effects of Nanostructure in Gold and Platinum Catalysis. J. Am. Chem. Soc. 2017, 139, 11012–11015. [Google Scholar] [CrossRef]
  19. Alkorta, I.; Legon, A.C. Non-Covalent Interactions Involving Alkaline-Earth Atoms and Lewis Bases B: An ab Initio Investigation of Beryllium and Magnesium Bonds, B···MR2 (M = Be or Mg, and R = H, F or CH3). Inorganics 2019, 7, 35. [Google Scholar] [CrossRef] [Green Version]
  20. Mó, O.; Montero-Campillo, M.M.; Alkorta, I.; Elguero, J.; Yáñez, M. Ternary Complexes Stabilized by Chalcogen and Alkaline-Earth Bonds: Crucial Role of Cooperativity and Secondary Noncovalent Interactions. Chem. Eur. J. 2019, 25, 11688–11695. [Google Scholar] [CrossRef]
  21. Alkorta, I.; Hill, J.G.; Legon, A.C. An ab initio investigation of alkali–metal non-covalent bonds B⋯LiR and B⋯NaR (R = F, H or CH3) formed with simple Lewis bases B: The relative inductive effects of F, H and CH3. Phys. Chem. Chem. Phys. 2020, 22, 16421–16430. [Google Scholar] [CrossRef] [PubMed]
  22. Yáñez, M.; Sanz, P.; Mó, O.; Alkorta, I.; Elguero, J. Beryllium Bonds, Do They Exist? J. Chem. Theor. Comput. 2009, 5, 2763–2771. [Google Scholar] [CrossRef] [PubMed]
  23. Montero-Campillo, M.M.; Mó, O.; Yáñez, M.; Alkorta, I.; Elguero, J. Chapter three—The beryllium bond. In Advances in Inorganic Chemistry; van Eldik, R., Puchta, R., Eds.; Academic Press: Cambridge, MA, USA, 2019; Volume 73, pp. 73–121. [Google Scholar]
  24. Stockmann, P.; Gozzi, M.; Kuhnert, R.; Sárosi, M.B.; Hey-Hawkins, E. New keys for old locks: Carborane-containing drugs as platforms for mechanism-based therapies. Chem. Soc. Rev. 2019, 48, 3497–3512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Fisher, S.P.; Tomich, A.W.; Lovera, S.O.; Kleinsasser, J.F.; Guo, J.; Asay, M.J.; Nelson, H.M.; Lavallo, V. Nonclassical Applications of closo-Carborane Anions: From Main Group Chemistry and Catalysis to Energy Storage. Chem. Rev. 2019, 119, 8262–8290. [Google Scholar] [CrossRef]
  26. Hawthorne, M.F.; Popp, G. 3-Trimethylamino-3-beryl-1,2-dicarba-closo-dodecaborane(12) and related species. Inorg. Chem. 1971, 10, 391–393. [Google Scholar] [CrossRef]
  27. Alkorta, I.; Campillo, N.; Rozas, I.; Elguero, J. Ring Strain and Hydrogen Bond Acidity. J. Org. Chem. 1998, 63, 7759–7763. [Google Scholar] [CrossRef]
  28. Mujika, J.I.; Mercero, J.M.; Lopez, X. Water-Promoted Hydrolysis of a Highly Twisted Amide: Rate Acceleration Caused by the Twist of the Amide Bond. J. Am. Chem. Soc. 2005, 127, 4445–4453. [Google Scholar] [CrossRef]
  29. Alkorta, I.; Montero-Campillo, M.M.; Elguero, J.; Yáñez, M.; Mó, O. Complexes between neutral oxyacid beryllium salts and dihydrogen: A possible way for hydrogen storage? Dalton Trans. 2018, 47, 12516–12520. [Google Scholar] [CrossRef]
  30. Alkorta, I.; Montero-Campillo, M.M.; Elguero, J.; Yáñez, M.; Mó, O. Complexes between H2 and neutral oxyacid beryllium derivatives. The role of angular strain. Mol. Phys. 2018, 1–9. [Google Scholar] [CrossRef]
  31. Curtiss, L.A.; Redfern, P.C.; Raghavachari, K. Gaussian-4 theory. J. Chem. Phys. 2007, 126, 084108. [Google Scholar] [CrossRef]
  32. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Rev. A.03; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  33. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  34. Stearns, C.A.; Kohl, F.J. Mass spectrometric determination of the dissociation energies of AlC2, Al2C2, and AlAuC2. J. Phys. Chem. 1973, 77, 136–138. [Google Scholar] [CrossRef]
  35. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Clarendon Press: Oxford, UK, 1990. [Google Scholar]
  36. Popelier, P.L.A. Atoms in Molecules. An introduction; Prentice Hall: Harlow, UK, 2000. [Google Scholar]
  37. Keith, T.A. AIMAll; Version 17.11.14 B; T.K Gristmill Software: Overland Park, KS, USA, 2017. [Google Scholar]
  38. Su, P.; Li, H. Energy decomposition analysis of covalent bonds and intermolecular interactions. J. Chem. Phys. 2009, 131, 014102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Schmidt, M.W.; Baldridge, K.K.; Boatz, J.A.; Elbert, S.T.; Gordon, M.S.; Jensen, J.H.; Koseki, S.; Matsunaga, N.; Nguyen, K.A.; Su, S.J.; et al. Gamess; Iowa State University: Ames, IA, USA, 2013. [Google Scholar]
  40. Mata, I.; Alkorta, I.; Molins, E.; Espinosa, E. Universal Features of the Electron Density Distribution in Hydrogen-Bonding Regions: A Comprehensive Study Involving H···X (X = H, C, N, O, F, S, Cl, π) Interactions. Chem. Eur. J. 2010, 16, 2442–2452. [Google Scholar] [CrossRef] [PubMed]
  41. Alkorta, I.; Elguero, J. Discrimination of hydrogen-bonded complexes with axial chirality. J. Chem. Phys. 2002, 117, 6463–6468. [Google Scholar] [CrossRef] [Green Version]
  42. Alkorta, I.; Solimannejad, M.; Provasi, P.; Elguero, J. Theoretical Study of Complexes and Fluoride Cation Transfer between N2F+ and Electron Donors. J. Phys. Chem. A 2007, 111, 7154–7161. [Google Scholar] [CrossRef]
  43. Legon, A.C.; Millen, D.J. Hydrogen bonding as a probe of electron densities: Limiting gas-phase nucleophilicities and electrophilicities of B and HX. J. Am. Chem. Soc. 1987, 109, 356–358. [Google Scholar] [CrossRef]
  44. Alkorta, I.; Legon, A.C. Nucleophilicities of Lewis Bases B and Electrophilicities of Lewis Acids A Determined from the Dissociation Energies of Complexes B⋯A Involving Hydrogen Bonds, Tetrel Bonds, Pnictogen Bonds, Chalcogen Bonds and Halogen Bonds. Molecules 2017, 22, 1786. [Google Scholar] [CrossRef] [Green Version]
  45. Alkorta, I.; Elguero, J.; Yáñez, M.; Mó, O.; Montero-Campillo, M.M. Relativistic Effects on NMR Parameters of Halogen-Bonded Complexes. Molecules 2019, 24, 4399. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The two complexes described for the 3-beryl-l,2-dicarbacloso-dodecaborane(12) [26].
Figure 1. The two complexes described for the 3-beryl-l,2-dicarbacloso-dodecaborane(12) [26].
Molecules 25 05876 g001
Figure 2. Schematic representation of the complexes between the four beryl-dicarbacloso-dodecaborane isomers (AD) and the Lewis Bases considered in the article.
Figure 2. Schematic representation of the complexes between the four beryl-dicarbacloso-dodecaborane isomers (AD) and the Lewis Bases considered in the article.
Molecules 25 05876 g002
Figure 3. Optimized structure of the four B9H11C2Be isomers considered in this article with the IUPAC nomenclature.
Figure 3. Optimized structure of the four B9H11C2Be isomers considered in this article with the IUPAC nomenclature.
Molecules 25 05876 g003aMolecules 25 05876 g003b
Figure 4. (a) Molecular electrostatic potential (MESP) on the 0.001 au electron density of B9H11C2Be (a). Regions with MESP > 0.1 au are indicated with blue color and MESP < −0.015 au with red color. The location of the Vs, max is indicated with a black sphere; (b) shows the LUMO orbital for the same species.
Figure 4. (a) Molecular electrostatic potential (MESP) on the 0.001 au electron density of B9H11C2Be (a). Regions with MESP > 0.1 au are indicated with blue color and MESP < −0.015 au with red color. The location of the Vs, max is indicated with a black sphere; (b) shows the LUMO orbital for the same species.
Molecules 25 05876 g004
Figure 5. Molecular graph for the four isomers of the beryllium ortho-carboranes investigated. The location of the bond, ring, and cage critical points is indicated with green, red, and blue dots, respectively. AD the same as in Figure 2.
Figure 5. Molecular graph for the four isomers of the beryllium ortho-carboranes investigated. The location of the bond, ring, and cage critical points is indicated with green, red, and blue dots, respectively. AD the same as in Figure 2.
Molecules 25 05876 g005aMolecules 25 05876 g005b
Figure 6. Molecular graph of some complexes. The location of the bond, ring, and cage critical points is indicated with green, red, and blue dots, respectively. AD the same as in Figure 2.
Figure 6. Molecular graph of some complexes. The location of the bond, ring, and cage critical points is indicated with green, red, and blue dots, respectively. AD the same as in Figure 2.
Molecules 25 05876 g006
Figure 7. G4 Dissociation enthalpies vs. fitted values (kJ mol−1). (a) Model 1: All the complexes with the Cl2Be indicated with red squares. (b) Model 2: All the complexes except those involving Cl2Be.
Figure 7. G4 Dissociation enthalpies vs. fitted values (kJ mol−1). (a) Model 1: All the complexes with the Cl2Be indicated with red squares. (b) Model 2: All the complexes except those involving Cl2Be.
Molecules 25 05876 g007
Table 1. Relative enthalpies, ΔHrel (kJ·mol–1), of the four isomers of B9H11C2Be, with respect to isomer D and Vs,max (kJ·mol–1). The Vs,max values of CO3Be, SO4Be, and Cl2Be are also included.
Table 1. Relative enthalpies, ΔHrel (kJ·mol–1), of the four isomers of B9H11C2Be, with respect to isomer D and Vs,max (kJ·mol–1). The Vs,max values of CO3Be, SO4Be, and Cl2Be are also included.
ΔHrelVs, max
B9H11C2Be (A)17.7669.4
B9H11C2Be (B)34.7625.9
B9H11C2Be (C)29.2605.0
B9H11C2Be (D)0.0575.3
CO3Be 906.7
SO4Be 938.5
Cl2Be 141.0
Table 2. Intermolecular distance (Å) between the Be atom and the basic site of the Lewis bases.
Table 2. Intermolecular distance (Å) between the Be atom and the basic site of the Lewis bases.
Base\AcidABCDCO3BeSO4BeCl2Be
N21.6921.6901.6881.6911.7241.7291.884
CO1.7351.7291.7221.7251.7941.8021.874
SH22.1272.1252.1342.1372.1182.1132.250
PH32.1402.1382.1372.1402.1672.1642.278
NCH1.6601.6611.6631.6651.6711.6701.730
OH21.6411.6461.6591.6591.6031.5971.642
CNH1.7361.7321.7281.7301.7701.7711.818
NH31.7081.7121.7191.7201.6971.6911.739
Table 3. Dissociation enthalpies (kJ·mol−1) calculated at G4 computational level.
Table 3. Dissociation enthalpies (kJ·mol−1) calculated at G4 computational level.
Base\AcidABCDCO3BeSO4BeCl2Be
N2111.8108.683.857.593.8103.66.1
CO142.9140.6116.789.6117.1127.724.7
SH2160.9154.5124.296.7146.5160.645.2
PH3169.6163.4135.3107.6150.0165.345.2
NCH193.3184.0153.0125.3185.1201.071.1
OH2194.3184.8150.3123.3197.4212.097.3
CNH206.7198.9169.5141.1190.6206.577.5
NH3241.2231.6199.2171.9238.6256.1124.7
Table 4. Deformation energy (kJ·mol−1) due to the complex formation.
Table 4. Deformation energy (kJ·mol−1) due to the complex formation.
A:LB aB:LBC:LBD:LBCO3Be:LBSO4Be:LBCl2Be:LB
BaseBeD bLBBeDLBBeDLBBeDLBBeDLBBeDLBBeDLB
N22.30.02.40.02.30.02.50.01.90.10.20.128.40.1
CO3.00.23.00.12.80.13.10.12.90.60.90.834.60.3
SH24.20.34.20.33.90.24.20.23.20.51.90.538.40.4
PH34.97.54.86.84.46.04.75.84.18.72.89.741.15.4
NCH5.60.75.70.75.40.75.90.74.81.03.11.148.10.8
OH24.71.14.91.14.90.95.31.03.72.52.32.645.13.9
CNH5.51.65.41.45.21.25.61.25.12.13.32.448.01.5
NH37.10.07.20.07.00−.17.50−.15.30.14.00.154.80.3
a LB stands for Lewis Base. b BeD stands for Beryllium derivative.
Table 5. Statistical values and fitted parameters.
Table 5. Statistical values and fitted parameters.
Model 1Model 2
AllWithout BeCl2
R20.960.98
SD (kJ mol−1)9.95.9
Ea values
Cl2Be5.5
CO3Be13.813.2
SO4Be14.914.3
A14.614.0
B14.113.5
C11.711.2
D9.69.2
Nb values
N27.07.5
CO9.29.7
SH210.711.2
PH311.211.8
NCH13.213.8
OH213.714.1
CNH14.114.7
NH317.117.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Sample Availability: Samples of the compounds are not available from the authors.

Share and Cite

MDPI and ACS Style

Alkorta, I.; Elguero, J.; Oliva-Enrich, J.M.; Yáñez, M.; Mó, O.; Montero-Campillo, M.M. The Importance of Strain (Preorganization) in Beryllium Bonds. Molecules 2020, 25, 5876. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25245876

AMA Style

Alkorta I, Elguero J, Oliva-Enrich JM, Yáñez M, Mó O, Montero-Campillo MM. The Importance of Strain (Preorganization) in Beryllium Bonds. Molecules. 2020; 25(24):5876. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25245876

Chicago/Turabian Style

Alkorta, Ibon, José Elguero, Josep M. Oliva-Enrich, Manuel Yáñez, Otilia Mó, and M. Merced Montero-Campillo. 2020. "The Importance of Strain (Preorganization) in Beryllium Bonds" Molecules 25, no. 24: 5876. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25245876

Article Metrics

Back to TopTop