Next Article in Journal
Hepatic Oleate Regulates Insulin-like Growth Factor-Binding Protein 1 Partially through the mTORC1-FGF21 Axis during High-Carbohydrate Feeding
Next Article in Special Issue
Cyclic Dimers of 4-n-Propyloxybenzoic Acid with Hydrogen Bonds in the Gaseous State
Previous Article in Journal
Regulation of P-Glycoprotein in the Brain
Previous Article in Special Issue
Molecular Structure, Vibrational Spectrum and Conformational Properties of 4-(4-Tritylphenoxy)phthalonitrile-Precursor for Synthesis of Phthalocyanines with Bulky Substituent
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nature of Beryllium, Magnesium, and Zinc Bonds in Carbene⋯MX2 (M = Be, Mg, Zn; X = H, Br) Dimers Revealed by the IQA, ETS-NOCV and LED Methods

1
Faculty of Chemistry, Jagiellonian University, Gronostajowa 2, 30-387 Cracow, Poland
2
Faculty of Chemistry, Nicolaus Copernicus University in Toruń, Gagarina 7, 87-100 Toruń, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(23), 14668; https://doi.org/10.3390/ijms232314668
Submission received: 24 October 2022 / Revised: 21 November 2022 / Accepted: 22 November 2022 / Published: 24 November 2022

Abstract

:
The nature of beryllium–, magnesium– and zinc–carbene bonds in the cyclopropenylidene⋯MX 2 (M = Be, Mg, Zn; X = H, Br) and imidazol-2-ylidene⋯MBr 2 dimers is investigated by the joint use of the topological QTAIM-based IQA decomposition scheme, the molecular orbital-based ETS-NOCV charge and energy decomposition method, and the LED energy decomposition approach based on the state-of-the-art DLPNO-CCSD(T) method. All these methods show that the C⋯M bond strengthens according to the following order: Zn < Mg < < Be. Electrostatics is proved to be the dominant bond component, whereas the orbital component is far less important. It is shown that QTAIM/IQA underestimates electrostatic contribution for zinc bonds with respect to both ETS-NOCV and LED schemes. The σ carbene→MX 2 donation appears to be much more important than the MX 2 carbene back-donation of π symmetry. The substitution of hydrogen atoms by bromine (X in MX 2 ) strengthens the metal–carbene bond in all cases. The physical origin of rotational barriers has been unveiled by the ETS-NOCV approach.

Graphical Abstract

1. Introduction

Carbenes are undoubtedly one of the most important groups of organic compounds [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18]. They are characterized by having only two valence carbon atoms, thanks to which it generally forms two single covalent bonds to the substituents, CR 1 R 2 . As a consequence, the carbon atom retains two unbound electrons, which in a singlet state occupy the same p orbital, creating a lone electron pair, which is the cause of nucleophilic properties of carbenes. In other words, singlet carbenes are good Lewis bases. It is therefore not surprising that carbenes participate in various kinds of intermolecular interactions, such as hydrogen bonds [19,20,21,22,23,24,25,26,27], lithium bonds [16,18,28,29,30], beryllium bonds [16,18,31,32,33,34,35], magnesium bonds [16,18,34,35,36,37,38], triel bonds [18,39,40,41,42], tetrel bonds [18,43,44,45], pnictogen bonds [18,46,47,48], chalcogen bonds [18,49], halogen bonds [18,50,51,52,53] (in particular to the iodine atom [50,51]), and aerogen bonds [54]. However, carbenes in particular are known for their great ease of bonding with transition metal atoms [8,12,16,36,55,56,57,58,59,60,61] to form various adducts, which is the goal of numerous studies in organometallic chemistry. In this case, the N-heterocyclic carbenes (NHCs), especially those derived from imidazol-2-ylidene, are of particular importance [4,8,9,11,12,15,16,17,18,27]. In turn, within the group of the NHC–M complexes, those involving zinc are of increasing practical importance [35,36,38,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78], due in part to the easy availability and relatively low cost of precursors of such complexes. Some practical applications of NHC–Zn complexes in organometallic catalysis have recently been extensively described [62,63,64,65]. It is worth mentioning, however, that apart from the lone electron pair which gives singlet carbenes nucleophilic properties, singlet carbenes also feature electrophilic properties [7,79,80,81,82,83,84] resulting from the presence of an empty p orbital perpendicular to the plane formed by CR 1 R 2 .
Very recently, one of us (M. J.) has investigated [35] the beryllium [85,86,87,88,89,90,91,92,93], magnesium [85,92,93,94,95,96] and zinc (spodium) [78,97,98,99,100] bonds in a large group of dimers formed by carbenes ((NH 2 ) 2 C, imidazol-2-ylidene, imidazolidin-2-ylidene, tetrahydropyrymid-2-ylidene, and cyclopropenylidene) or carbodiphosphoranes ((PH 3 ) 2 C and (NH 3 ) 2 C) and MX 2 (M = Be, Mg, Zn and X = H, F, Cl, Br, Me) molecules. This was the first study of beryllium, magnesium, and zinc bonds of the C⋯M type. Moreover, in the vast majority of cases, in earlier systems with the beryllium or magnesium bond, the Lewis base was a small molecule involving either atom being good electron donors [85,86,87,90,92,93,95] or π -electron bonds [88]. On the other hand, theoretical research on zinc bonds is rather sporadic [35,78,97,100].
Among many important findings was the demonstration that carbene⋯MX 2 dimers are characterized by a very high charge transfer from carbene to the MX 2 molecule [35]. Namely, it was from two to almost four times greater than in the case of HOH⋯OH 2 and HOH⋯NH 3 dimers. This result showed that the carbene⋯MX 2 (M = Be, Mg, Zn; X = H, F, Cl, Br, Me) dimers are undoubtedly systems in which the charge transfer effect plays a large role, especially in the presence of very polarizable bromine atoms as X. This, in turn, proves the strong interorbital interactions between the carbene molecule and MX 2 . Importantly, it has been shown that in terms of the electron density and the total electronic energy density values computed at the bond critical point [101] of C⋯M the C⋯Zn zinc bond is similar to the C⋯Be beryllium bond. Treating the value of electron density as a measure of bond strength [101], this result has shown that the zinc bonds should be of similar strength to the beryllium bonds, while the magnesium bonds should be significantly weaker than them [35]. Moreover, together with positive values of the Laplacian of the electron density, negative values of the total electronic energy density obtained for the beryllium and zinc bonds have indicated a significant degree of electron sharing, in turn reflecting a high degree of covalence [102].
A somewhat surprisingly different picture of the similarity of these three types of bonding has been obtained on the basis of the delocalization index, δ (C,M), which describes the exchange of the electrons in the basins of atoms C and M [103,104,105,106]. Namely, the δ (C,M) values for the zinc bond were large, while those for the beryllium and magnesium bond were much lower and similar to each other [35]. Thus, taking into account the relationship of δ (A,B) to the exchange energy [107] and therefore treating δ (A,B) as a measure of the covalent character of an A–B bond [108], this result has shown that the zinc bond is much more covalent than the beryllium and magnesium bond. On the other hand, the NCI-based (the abbreviation NCI is derived from the Non-Covalent Interactions index [109,110]) analysis has shown [35] that the zinc bond is the strongest; however, the beryllium bond is only slightly weaker.
In light of the previously obtained results [35] briefly recalled here, it is necessary to investigate comparatively the physical nature of the beryllium, magnesium and TM-type zinc bonds, not only qualitatively but also quantitatively. This is the main goal of this article, and this nature will be analyzed using complementary energy decomposition methods, IQA (i.e., the Interacting Quantum Atoms approach [111,112]), which provides a local picture of bonding and ETS-NOCV (i.e., the combination of the Extended Transition State (ETS) method [113] with the Natural Orbitals for Chemical Valence (NOCV) method [114,115,116,117,118]), which is more suited for the description of bonding between molecular fragments. Results will also be compared with the state-of-the-art DLPNO-CCSD(T) (i.e., the Domain-based Localized Pair-Natural Orbital Singles and Doubles Coupled Cluster with perturbative Triples) [119,120,121,122,123] calculations and the Local Energy Decomposition (LED) scheme [124,125].

2. Results and Discussion

As already mentioned in the Introduction section, one of us (M. J.) has recently investigated [35] beryllium, magnesium and zinc (spodium) bonds in a large group of dimers formed by either a carbene ((NH 2 ) 2 C, imidazol-2-ylidene, imidazolidin-2-ylidene, tetrahydropyrymid-2-ylidene, and cyclopropenylidene) or a carbodiphosphorane ((PH 3 ) 2 C and (NH 3 ) 2 C) and the MX 2 (M = Be, Mg, Zn and X = H, F, Cl, Br, Me) species. While the C⋯M contact has been shown to be certainly dominant in all of these dimers, some of them contain additional intermolecular interactions (either hydrogen bond or dihydrogen bond). Since the purpose of this article is to quantify the physical nature of the C⋯M bond, such dimers are not considered here. In other words, in this article, only those dimers in which the described interaction C⋯M is undoubtedly dominant will be considered. The structural characteristics of these dimers will be briefly discussed in the first subsection. Then, in the second and third subsections, the description of the C⋯M interaction obtained by the IQA and ETS-NOCV methods, respectively, will be presented. The correspondence of the results with the outcomes originating from the DLPNO-CCSD(T) method, and based on it, LED energy decomposition will be shown in the fourth subsection. Finally, in the fifth subsection, we will analyze the energy changes of dimers during the rotation of carbene with respect to the ZnBr 2 subunit.

2.1. Structural Characteristics of the Considered Dimers

To meet the aforementioned condition, in this article only the cyclopropenylidene⋯MX 2 (M = Be, Mg, Zn; X = H, Br) and, additionally, imidazol-2-ylidene⋯MBr 2 dimers will be studied (see Figures S1 and S2). It is worth noting that in the former case, the lack of additional intermolecular interactions (i.e., apart from the described C⋯M) is ensured by the perpendicular orientation of the cyclopropenylidene and MX 2 molecules (see the representative cyclopropenylidene⋯ZnBr 2 dimer in Figure 1).
In the latter case, however, although the flatness of the entire dimer admits the possibility of the presence of N–H⋯Br hydrogen bonds, these interactions should be relatively weaker than C⋯M due to the large distance H⋯Br and the almost parallel orientation of the N–H and Zn–Br bonds relative to each other (see the representative imidazol-2-ylidene⋯ZnBr 2 dimer in Figure 1). A thorough analysis of these intermolecular N-H⋯Br hydrogen bonds will also be performed. The values of some fundamental parameters characterizing the considered carbene⋯MX 2 dimers are presented in Table 1.
As can be seen from Figure 1 and Table 1 (see the α XMX angle), the characteristic feature of the considered dimers is the clearly bent structure of the MX 2 molecule. It is worth noting that this bend is similar in the case of MX 2 molecules with beryllium or zinc, while it is clearly smaller in the case of the magnesium atom, which forms more ionic and thus more resistant M–X bonds. Thus, the value of the angle α XMX suggests the similarity between the Be and Zn atoms. It should be noted that the bending of the MX 2 molecule is a characteristic effect of the presence of a beryllium or magnesium bond in dimers with different Lewis bases [85,86,88,92]. Moreover, Martín-Sómer et al. [86] have shown that in the case of H 3 N⋯BeH 2 n X n (X = F, Cl, Br; n 2) dimers, the nonlinearity of the BeH 2 n X n molecule is due to the decrease in LUMO energy. It is also worth noting that for the bond length M–X the relation Be<Zn<Mg applies so that for a given X the magnesium atom forms the longest M–X bond. It should also be emphasized that the formation of the C⋯M bond leads to significant elongation of the M–X bond (see Δ d MX in Table 1) and a greater opening of the L–C–L angle in the carbene molecule (see Δ α LCL in Table 1). For the same MX 2 molecule, both of these effects are clearly greater in imidazol-2-ylidene than in cyclopropenylidene, suggesting that the C⋯M interaction should be stronger in the dimers of imidazol-2-ylidene. This is confirmed by the computed dissociation energies [35]. For imidazol-2-ylidene, these energies are in the 41.2–48.6 kcal/mol range, while for cyclopropenylidene, only 28.2–35.4 kcal/mol (for the same set of MX 2 molecules, i.e., MBr 2 ), similar outcomes are valid when considering other XC and DLPNO-CCSD(T) method, Table S1. The greatest dissociation energy value characterizes the dimers with Be, while the smallest D 0 are noted in the dimers containing Zn, which can only partially be explained by the shortest distance C⋯M in the case when M = Be. This is because in the case of Mg and Zn, the distance C⋯Mg is longer than that of C⋯Zn, but the magnesium bond is stronger than the zinc bond. For example, for the cyclopropenylidene⋯MH 2 dimers, the C⋯M lengths for Be, Mg, and Zn, respectively, are 1.743, 2.268, and 2.121 Å, while the dissociation energies are 29.1, 20.9, and 15.2 kcal/mol. This result shows that the dissociation energy is not completely dependent on the distance C⋯M. Such a trend has already been observed in literature [35].
It has already been mentioned in Introduction that the carbene⋯MX 2 dimers considered here are characterized by a very high charge transfer effect from the Lewis base, i.e., the carbene molecule, to the Lewis acid, i.e., the MX 2 molecule, and this effect is particularly large in the presence of highly polarizable bromine atoms. Indeed, in the case of the imidazol-2-ylidene⋯BeBr 2 dimer, the value of the charge transfer (based on chemically reliable [128,129,130] Hirshfeld atomic charges [126,127]) is as high as −0.464 au. It is enough to mention that the charge transfer determined on the same level of theory in the case of the water dimer is only −0.098 au. This comparison suggests the presence of strong interorbital interactions between the carbene molecule and the MX 2 unit. Their significance, however, can be quantified by ETS-NOCV and LED methods.

2.2. IQA-Based Results

As already mentioned in the Methods section, the IQA approach is a unique method that enables the determination of the interaction energy of any two atoms and the decomposition of this energy into individual terms, according to Equations (1) and (2). The results of such decomposition carried out for the contact C⋯M are shown in Table 2.
A certain inconvenience of IQA (which is also a characteristic feature of energy decomposition methods) is that the components of the interaction energy are generally much larger than the interaction energy itself. This is obviously due to the close cancellation of the negative and largest E neen and the positive sum of E nn and E ee , C . It is obvious that the rapid growth of the relevant components with the change Be→Mg→Zn results from both the increase in the atomic number of M, i.e., the charge of the atomic nucleus, and the increase in the number of electrons. However, it is much more important that according to IQA, the C⋯M (M = Be, Mg, Zn) interaction in all the dimers with the participation of cyclopropenylidene is stabilizing, which is indicated by negative values of the obtained interaction energies. For the same X, the binding effect increases in the Zn→Mg→Be direction and is slightly stronger when X = Br, so that the strongest C⋯M bond is for BeBr 2 (−0.3644 a.u.).
It is interesting to look at the values of the exchange-correlation energy ( E ee , xc ) as well as the percentage share of this component in the total interaction energy (% E ee , xc ). Of course, independently of X, E ee , xc increases (i.e., becomes more negative) in the direction Mg→Be→Zn, with the corresponding values being slightly greater for Br than for H. Importantly, the values for Zn are significantly greater than those for Be or Mg (e.g., E ee , xc is −0.1091 a.u. in cyclopropenylidene⋯ZnH 2 and only −0.0458 a.u. and −0.0307 a.u. in cyclopropenylidene⋯BeH 2 and cyclopropenylidene⋯MgH 2 , respectively). The contribution of negative electrostatic energy ( E elst ) in the case of the cyclopropenylidene dimers quickly decreases in the Be→Mg→Zn series, as a consequence of which the percentage of the exchange-correlation energy quickly increases from ca. 12–13% for the beryllium bond, to ca. 16% for the magnesium bond and up to ca. 63–66% for the zinc bond. This result shows that the IQA method suggests that the zinc bond is highly covalent, whereas the beryllium and magnesium bonds are electrostatic in nature. Of course, such a result was to be expected, because, as already mentioned in the Introduction section, the same trend has already been obtained earlier for the delocalization index δ (C,M) [35], which, as is known [107,108], correlates very well with the exchange-correlation energy.
Interestingly, IQA has given quite small interaction energies for the imidazol-2-ylidene⋯MBr 2 dimers, which do not match the significant dissociation energies for these complexes; Table 1. Even more surprisingly, a positive interaction energy value (0.0379 a.u.) has been obtained for the imidazol-2-ylidene⋯MgBr 2 dimer. This destabilizing effect of C⋯Mg (0.0379 a.u.) is due to a significant destabilizing electrostatic contribution (0.0777 a.u.), which exceeds the negative exchange-correlation energy (−0.0397, i.e., ca. 51% only). In the case of imidazol-2-ylidene⋯ZnBr 2 , E elst is also positive (0.1190 a.u.; and therefore the purely Coulombic contribution gives no bonding), but unlike for imidazol-2-ylidene⋯MgBr 2 , this contribution is excessively overshadowed (over 110%) by E ee , xc (−0.1317 a.u.), which gives stabilizing C⋯Zn bond, which could be considered as anti-electrostatic and covalent.
These diatomic IQA data do not explain the stability of the imidazol-2-ylidene⋯MgBr 2 species. To this end, we have also summed the interaction energies corresponding to the pairs of atoms in the groups representing the {carbene}⋯{MX 2 } fragmentation (i.e., all the combinations where one of the atoms is in the {carbene} moiety and the other one in {MX 2 }). In such treatment, which in fact ignores many-body effects, all the E int values are negative, as expected. Furthermore, the IQA fragment-based data (see Table 3) indicate domination of E elst over E xc for the Be- and Mg-based systems ( ca. 51–58% of E int ), with the Zn-based species as the only exceptions and the share of electrostatics in their total stabilization reaching ca. 16–21% only.
For the imidazol-2-ylidene complexes, among all the {carbene}⋯{MX 2 } atomic pairs, the most stabilizing according to IQA are N⋯M interactions, providing from ca. −150 kcal/mol in the case of Zn via. ca. −220 kcal/mol (Mg) to ca. −250 kcal/mol (Be), followed by C⋯Br (from ca. −43 kcal/mol to ca. −39 kcal/mol) and H⋯Br (from ca. −48 kcal/mol to ca. −32 kcal/mol), Table S2. Surprisingly, the C⋯M interactions are not among the most stabilizing ones, ranging from +20 kcal/mol (Mg) to −40 kcal/mol (Be). In the case of cyclopropenylidene complexes, the C⋯M interactions are definitely the most important, ranging from ca. −100 to over −200 kcal/mol, Tables S3 and S4.
In the fragment-based treatment, a quite different picture of the C⋯M bond nature has been obtained. Namely, for cyclopropenylidene⋯MX 2 , the exchange-correlation energy component became more prominent (ca. 45% for Be and Mg, and up to 80% in the case of Zn; Table 3). Concurrently, in the fragment-based approach, all the MX 2 complexes with imidazol-2-ylidene have negative E int values, as well as their electrostatic and exchange-correlation components; Table 3. Moreover, the character of these interactions changed in the case of Be and Mg as the electrostatic energy became dominant.
Clearly, the fragment-based approach seems to be much more credible for the description of the carbene⋯metal bonds in both cyclopropenylidene and imidazol-2-ylidene complexes. However, the lack of the C⋯M interaction among the most important contributions is alarming, as much as the counterintuitive positive E int value for the imidazol-2-ylidene⋯MgBr 2 dimer (Table 2). The most probable explanation is a quite significant positive QTAIM charge on the metal atoms (e.g., the QTAIM-based atomic charge of Mg is as much as 1.687 a.u. compared to 0.616 a.u. only obtained from the Hirshfeld method) [35]. This, coupled with the incorrect sign of charges on the carbene carbon atom (already stressed in the first work concerning these systems [35]), puts the reliability of IQA electrostatics in question. Since the IQA method clearly depends on the charge distribution between QTAIM atomic basins and therefore provides variable outcomes for different charge density partitionings, we will switch to ETS-NOCV, which estimates E elst in a more reliable way through the quasi-classical approach based on self-consistently converged fragment electron densities; see Methods.

2.3. ETS-NOCV-Based Results

In this section, we will present the results obtained by means of the molecular orbital-based ETS-NOCV method. We have used singlet {carbene} and {MX 2 } fragments which have appeared to be more suitable than triplets since they deliver the smallest orbital interaction contribution. The obtained values of the ETS-NOCV energy terms (see Equation (5)) are listed in Table 4.
Contrary to the atomic resolution-based IQA results, all ETS-NOCV interaction energies ( Δ E int ) are negative and correspond better to the bond dissociation energies ( D 0 values in Table 1). Bonds formed by beryllium are the strongest, followed by energetically similar magnesium and zinc bonds. Systems with Br atoms are bonded noticeably stronger than with H atoms, and complexes with imidazol-2-ylidene are stabilized more than their cyclopropenylidene counterparts. The carbene⋯metal bonds are dominated by electrostatics ( Δ E elst , Table 4), with its percentage contribution to the total stabilization being 67–71% in the case of Mg and Zn dimers and 56–61% in systems with Be. This is similar to other reports of bonding in transition metal complexes of imidazol-2-ylidene-based NHCs [12,56,57,58,59,61] and other model carbenes [55,58,61]. For example, it turned out that the electrostatic contribution to C⋯M bonds in imidazol-2-ylidene complexes with Cu, Ag and Au halides accounts for more than 65% of the binding energy [56]. This contribution was also very similar (ca. 2/3 of the interaction energy) in the various NHC–TM complexes studied by Tonner et al. [58]. It is worth mentioning at this point the excellent review on the NHC–M bond by Jacobsen et al. [12], in which it was found that NHC–M bonds are mainly electrostatic in origin with a percentage contribution from 60% to 80%.
Orbital interaction is the second strongest energy term, reaching up to 40% of total stabilization in beryllium systems and 25–30% in the case of Mg and Zn systems, Table 4. The carbene→metal donation (the LP(C) σ (MX) transfer in MO language) dominates by far (72–87%) over the metal→carbene back-donation in π symmetry, with 16–20% contribution in the case of systems with X = H and 7–12% for X = Br. The clearly lower back-donation effect is visible in the latter case, resulting from the electron-withdrawing properties of Br atoms. It should be pointed out that in many transition metal complexes, the π -back-bonding may be far more pronounced and can reach even 40% of the overall orbital interaction energy [60]. Finally, it should be commented that the geometry distortion term appeared to be the most pronounced for Be-species, Table S5. Also noteworthy, weak (up to −1.7 kcal/mol, i.e., ca. 3% of the orbital interaction energy) side N–H⋯Br hydrogen bonds have been identified in the imidazol-2-ylidene⋯MBr 2 complexes, Figure 2.
Such a negligible share of N–H⋯Br hydrogen bonds really confirms the correct choice of systems with the dominant C⋯M interaction. As ETS-NOCV is only capable of separating the Δ E orb contribution, it is necessary to support it with complementary IQA results, which indicate a strong electrostatic component ( E elst H Br represents 95% of E int H Br amounting to −45.5 kcal/mol). Such pronounced interaction energies are common for a two-atomic approach. For example, for the hydrogen bond in water, the Δ E int O H value amounts to ca. −100 kcal/mol [111].

2.4. LED-Based Results

Local Energy Decomposition [124,125] allows for partitioning of the binding energy at the DLPNO-CCSD(T) [120,121,122] level of theory, often considered as the `gold standard’ for calculating interaction energies. The same fragments as before, i.e., {carbene} and {MX 2 }, have been chosen. The binding energies obtained at the DLPNO-CCSD(T) level of theory are consistent with the DFT-based energies, that is, the beryllium bonds are the strongest, followed by magnesium and zinc (spodium) bonds; compare Table 1 and Table 5.
Local Energy Decomposition partitioning confirms the importance of electrostatics in metal–carbene bonding, with E elst ref 90% of total stabilization for all compounds, Table 5. Dispersion contribution of ca. 2% of the total interaction energy proved to be very small. The charge-transfer contributions, bottom of the Table 5, confirm qualitatively the picture obtained by ETS-NOCV, i.e., the carbene → MX 2 donation being far more important than MX 2 carbene back–donation.

2.5. Flat vs. Perpendicular Structure of the Carben⋯MX 2 Dimer

Looking at Figure 1, showing the cyclopropenylidene ⋯ZnBr 2 and imidazol-2-ylidene⋯ZnBr 2 dimers, the fundamental difference between them in the orientation of the ZnBr 2 molecule relative to the carbene plane is clearly visible. Namely, the ZnBr 2 and carbene units are perpendicular to each other in cyclopropenylidene⋯ZnBr 2 , while imidazol-2-ylidene⋯ZnBr 2 is planar. Taking into account the fact that the cause of the twisting of the metal-unit in relation to the carbene plane was described in the literature only very sporadically [78], and taking advantage of the relatively high structural simplicity of the studied systems, it was tempting to investigate the cause of the structural differences regarding the twisting of the MX 2 and carbene planes in relation to each other. For this purpose, we have performed additional calculations for the twisted structures of both dimers, i.e., cyclopropenylidene⋯ZnBr 2 and imidazol-2-ylidene⋯ZnBr 2 , in which the twist angle (denoted by θ ) L-C-Zn-Br (L is either C or N) was changed every 10 from 0 to 90 . Figure 3 (left) shows the energy profile related to the change in θ from 0 to 90 .
It is clearly visible that although the equilibrium geometry of the cyclopropenylidene ⋯ZnBr 2 dimer corresponds to the perpendicular structure (i.e., with θ = 90 ), the planar structure ( θ = 0 ) is energetically less stable by only 1 kcal/mol so that the rotation of the ZnBr 2 unit around the C⋯Zn bond axis is practically free, Figure 3. Dimer imidazol-2-ylidene⋯ZnBr 2 presents a completely different situation, namely the flat structure of this dimer is more stable than the perpendicular structure by about 6.5 kcal/mol which proves hindered rotation of ZnBr 2 due to the presence of adjacent N–H bonds, Figure 1. Planarization clearly favors the formation of residual hydrogen bonds N–H⋯Br giving rise to Zn⋯C bond contraction by ca. 0.023 Å; Figure 3, right.
According to the Walsh rule [131], a molecule adopts the conformation that most ensures stabilization of HOMO. A Walsh diagram showing the dependence of the orbital energies of the three highest occupied orbitals, i.e., HOMO, HOMO-1 and HOMO-2, on the twist angle θ is shown in Figure 4.
It is clearly visible that in the case of imi-ZnBr 2 , i.e., the imidazol-2-yliden⋯ZnBr 2 complex, the 90 0 rotation leads to a greater stabilization of all the orbitals considered, and thus also of HOMO. Thus, based on Walsh’s rule, the planar structure should be more stable for this complex than the perpendicular structure, which is indeed true (see Figure 1 and Figure 3). By the way, it is worth noting that both structures are characterized by reversed positions of HOMO and HOMO-1. In the case of the cyclopropenylidene⋯ZnBr 2 dimer, the 0 90 rotation leads to the stabilization of HOMO, but this stabilization is so minimal that it is doubtful to see this stabilization as the cause of the perpendicular structure of this dimer (Figure 1). Nevertheless, small changes in orbital energies may perhaps explain a small change in total energy (Figure 3) and almost unhindered rotation around the C⋯Zn axis.
However, in order to make a more reliable assessment of structural changes, ETS-NOCV contributions along rotation angle have been determined and depicted in Figure 5.
It can be seen that the significant shortening (0.024 Å) of the C⋯Zn distance upon planarization of the imidazol-2-ylidene⋯ZnBr 2 dimer (see Figure 3, right) is due to an increase in the orbital interaction term and, to a larger extent, the electrostatic contributions, which causes a drop of Δ E int by ca. 4.7 kcal/mol. The former, i.e., the orbital interactions, results mainly from the amplification of the σ (Zn–C) component and, to a lesser extent, the π (Zn–C) component, Figure 5. Interestingly, the effect of twisting (0 90 ) the cyclopropenylidene⋯ZnBr 2 dimer also causes a similar shortening (0.025 Å) of the distance C⋯Zn (Figure 3, right). Nevertheless, the overall C⋯Zn interaction energy ( Δ E int ) changes only slightly by ca. 1.3 kcal/mol, which is consistent with a facile rotation. Such small variations in C⋯Zn bond interaction energies are caused by modifications of both orbital interaction and electrostatic contributions, Figure 5. Finally, it is worth mentioning that rotation around the C⋯Zn axis in the cyclopropenylidene⋯ZnBr 2 dimer has practically no effect on the Zn–Br bond length, while in the imidazol-2-ylidene⋯ZnBr 2 dimer the transition from perpendicular to planar structure slightly lengthens the bond by 0.006 Å. This effect can be attributed to a larger LP(C) σ * (Zn–Br) donation.

3. Methods and Materials

Geometries of monomers and dimers were fully optimized on the ω B97X-D/6-311++G (2df,2p) level of theory, that is utilizing the ω B97X-D range-separated hybrid GGA exchange-correlation functional [132] of Density Functional Theory (DFT) [133,134,135,136] and the 6-311++G(2df,2p) basis set [137,138,139,140,141,142,143], which includes both polarization and diffuse functions. Mardirossian and Head-Gordon have shown [136] that ω B97X-D is one of the best exchange-correlation functionals among 200 for general purposes including intermolecular interactions. To increase the accuracy of the optimization procedure and numerical integration, cutoffs on forces and step size that are used to determine convergence were additionally tightened (0.000015 and 0.000010 for maximum force and its root mean square, respectively, and 0.000060 and 0.000040 for maximum displacement and its root mean square, respectively) and integration grid was increased to the (99, 590) one (UltraFine) having 99 radial shells and 590 angular points per shell. Vibration analysis showed that all the considered systems correspond to real minima on potential energy surfaces. Both geometry optimization and vibration analysis were performed by means of Gaussian 09 [144].
The Interacting Quantum Atoms (IQA) approach [111,112], which is based on Bader’s QTAIM (i.e., Quantum Theory of Atoms in Molecules) [101], allows the total energy of a system to be divided into mono- and polyatomic contributions. Of the many useful terms defined by the IQA method, this article will utilize the interatomic interaction energy:
E int E 1 E 2 = E nn E 1 E 2 + E ne E 1 E 2 + E en E 1 E 2 + E ee E 1 E 2 ( E 1 E 2 )
where E nn E 1 E 2 is the repulsion energy between nuclei of atoms E 1 and E 2 , V ne E 1 E 2 is the attraction energy between the nucleus of the atom E 1 and the electrons of the atom E 2 , E en E 1 E 2 is the attraction energy between electrons of the atom E 1 and the nucleus of the atom E 2 and E ee E 1 E 2 is the interatomic two-electron repulsion energy. The sum of the middle two terms gives the energy of the interatomic nucleus–electron attraction ( E neen E 1 E 2 ). Importantly, the interelectron repulsion energy can be further divided into a sum of the purely classical (Coulombic) contribution and the exchange-correlation (i.e., the non-classical term) energy:
E ee E 1 E 2 = E ee , C E 1 E 2 + E ee , xc E 1 E 2
Moreover, the sum of the first three terms in Equation (1) and the classical interelectron energy contribution gives the electrostatic energy, such that:
E int E 1 E 2 = E elst E 1 E 2 + E ee , xc E 1 E 2
As one can see, all the terms in the IQA equations have strictly defined physical meanings. It is also worth emphasizing that an extremely valuable feature of the IQA decomposition method is the fact that it does not require any reference system and that the interatomic interaction energy can be determined for any pair of E 1 and E 2 atoms, e.g., not necessarily connected with each other by a bond path. In this article, the interatomic interaction energy and its components will be determined for the C⋯M (M = Be, Mg, Zn) interaction. Moreover, for an estimation of the total interaction energy between carbene and metal moieties, pair atomic energies between all atoms in both fragments were summed, such as for moieties G and H :
E int GH = A G B H E int AB
The AIMAll program [145] was utilized for this purpose. MP2/6-311+G(d,p)-based [146,147] wave functions generated in Gaussian09 were used based on the aforementioned geometries.
The ETS-NOCV method [118] is a combination of the ETS Ziegler–Rauk method [113] with the Natural Orbitals for Chemical Valence (NOCV) method [114,115,116,117]. In the ETS method, the total interaction energy between fragments A and B (atoms, fragments and molecules) is decomposed according to the following equation:
Δ E total = Δ E dist + Δ E elst + Δ E Pauli + Δ E disp + Δ E orb = Δ E dist + Δ E int
where Δ E dist is the distortion energy that is needed to transform the structures of the isolated fragments A and B into the geometries they have in the created A–B system, Δ E elst is the energy of the electrostatic interaction of fragments A and B (in their final positions in the molecule, but frozen electron densities), Δ E Pauli is the Pauli repulsion energy resulting from the repulsion between electrons with same spins of fragments A and B, Δ E disp is the dispersion energy, and Δ E orb is the orbital interaction energy associated with the formation of the A–B system (e.g., an A–B bond) and results from the interaction between the occupied orbitals of the A subsystem with the virtual orbitals of the B subsystem and vice versa. This term also contains the polarization energy related to the electron density changes in each of the A and B subsystems and the interfragment charge transfer. The last four terms of Equation (5) form the interaction energy, Δ E int . On the other hand, the NOCV orbitals are eigenvectors of the differential charge and bond orders matrix, Δ P = P P 0 , where P is the density matrix of a molecule (e.g., A–B), and P 0 is the sum of the density matrices of the orthogonalized fragments A and B, P A 0 + P B 0 . The overriding feature of the NOCV orbitals representation is the ability to decompose the differential electron density, Δ ρ orb = ρ AB ρ A 0 ρ B 0 , into chemically meaningful Δ ρ orb k diagonal contributions:
Δ ρ orb = k = 1 M / 2 ν k [ ψ k 2 + ψ k 2 ] = k = 1 M / 2 Δ ρ orb k
where ν k are the eigenvalues obtained by diagonalizing the matrix Δ P , and M is the number of basis set functions. Visualization of contributions Δ ρ orb k enables the identification and characterization of the individual components ( σ , π , δ , etc.) of a bond. Importantly, under the ETS-NOCV method, it is possible to quantify the energies ( Δ E orb k ) corresponding to the individual Δ ρ orb k components:
Δ E orb = k = 1 M / 2 ν k [ F k , k TS + F k , k TS ] = k = 1 M / 2 Δ E orb k
Calculations under the ETS-NOCV method were performed on the BLYP-D3(BJ) [148,149,150]/ TZP-ZORA [151,152] level of theory using the ADF program [153]. Such a computational protocol proved to be adequate for bonding analysis both in our and other previous studies [154,155]. In this work, we will focus on the instantaneous interaction energy Δ E int , since total energy has already been studied in Ref. [35].
The Domain-based Localized Pair-Natural Orbital Singles and Doubles Coupled Cluster with perturbative Triples [120,121,122,123] (DLPNO-CCSD(T)) approach is a relatively new approximation of the Coupled Clusters method, which in turn is often referred to as a `golden standard’ for quantum-chemical calculations [142]. DLPNO-CCSD(T) itself recovers >95% of triples contribution and 99.8% of correlation energy, and thanks to its scaling being close to DFT, it is the first CCSD(T) approximation capable of calculating the entire protein [122]. It is performed by localizing the orbitals and generating the Projected Atomic Orbitals from occupied localized orbitals, followed by creation of Pair Natural Orbitals from selected PAOs [119,120].
Local Energy Decomposition [124,125] was created for the DLPNO-CCSD(T) approach in order to harness its accuracy in determining interaction energies and provide an energy decomposition scheme on the CCSD(T) level of theory. It decomposes the binding energy into the following quantities:
Δ E = Δ E geo prep + Δ E int
The first component is the energy needed to distort fragments A and B from their optimal geometries to their final state in the AB adduct. On the other hand, Δ E int is the interaction energy of said fragments, which can be further decomposed into reference (HF) energy, Δ E int HF , and a correlation contribution ( Δ E int C ):
Δ E int = Δ E int HF + Δ E int C
The Hartree–Fock part can in turn be expanded into:
Δ E int HF = Δ E el prep HF + Δ E elstat + Δ E exch
where Δ E el prep HF is the energy needed to distort the electronic structure of fragments into the state optimal for fragment interaction. Electrostatic interaction between monomers is denoted by Δ E elstat , and the HF exchange by Δ E exch . Correlation energy is decomposed into:
Δ E int C = Δ E no disp C + Δ E disp C + Δ E int C ( T )
Terms Δ E disp C and Δ E no disp C are correlation energy contributions associated with dispersion and non-dispersive corrections, respectively. The term denoted by Δ E int C ( T ) is the triples correction arising from CCSD(T).

4. Conclusions

This article presents a comparative study of a series of metal–carbene bonds between main- and transition group metals: beryllium, magnesium and zinc, and model carbenes: cyclopropenylidene (a regular carbene), and imidazol-2-ylidene (N-heterocyclic carbene, NHC). The physical nature of these carbene⋯MX 2 (X = H, Br) interactions has been described for the first time by the joint use of topological QTAIM-based IQA decomposition scheme, molecular orbital-based ETS-NOCV charge and energy decomposition method, as well as LED energy decomposition basing on the state-of-the-art DLPNO-CCSD(T) method.
All methods agree on the increasing bond strength in a series: Zn < Mg < < Be. For beryllium and magnesium bonds, electrostatics proved to be the most important bond component (55–90% of total stabilization), followed by stabilization due to electron sharing (10–45% of total stabilization), while dispersion interactions proved to be marginal (2–5%). QTAIM/IQA underestimates electrostatic contribution for zinc bonds with respect to both ETS-NOCV and LED schemes which, similarly to Be and Mg species, identify the following importance of Zn–carbene bond constituents: E elstat > E orb > E dispersion . Both ETS-NOCV and LED also provide a consistent picture of these carbene bonds on an orbital level, with the σ carbene→MX 2 donation strongly dominating over the MX 2 carbene back-donation of π symmetry. Substitution of hydrogen atoms by bromines (X in MX 2 ) strengthens the metal–carbene bond in all cases due to the amplification of electrostatic forces as well as donation and back-donation effects. The former is likely due to making metal atoms more electrophilic. Finally, the conformational preferences (planar vs twisted) are controlled by the nature of C–Zn bonding as unveiled by ETS-NOCV analyses performed along rotation coordinate pathways.

Supplementary Materials

Author Contributions

Conceptualization, M.M. and M.J.; methodology, F.S., M.M. and M.J.; software, F.S., M.M. and M.J.; validation, F.S., M.M. and M.J.; formal analysis, F.S., M.M. and M.J.; investigation, F.S., M.M. and M.J.; data curation, F.S., M.M. and M.J.; writing–original draft preparation, F.S., M.M. and M.J.; writing–review and editing, F.S., M.M. and M.J.; visualization, F.S. and M.J.; project administration, M.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data available from the authors on reasonable request.

Acknowledgments

This research was supported in part by PLGrid Infrastructure.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
QTAIMQuantum Theory of Atoms in Molecules
IQAthe Interacting Quantum Atoms approach
ETSthe Extended Transition State method
NOCVthe Natural Orbitals for Chemical Valence method
DLPNODomain-based Localized Pair-Natural Orbital
CCSD(T)the coupled cluster (CC) single-double-triple method
LEDthe Local Energy Decomposition scheme

References

  1. Kirmse, W. Carbene Chemistry; Academic Press: Cambridge, MA, USA, 1964. [Google Scholar]
  2. Hubert, A.J. Catalysis in C1 Chemistry; Springer: Dordrecht, The Netherlands, 1983. [Google Scholar]
  3. Schubert, U. Advances in Metal Carbene Chemistry; Springer: Dordrecht, The Netherlands, 1989. [Google Scholar]
  4. Herrmann, W.A.; Köcher, C. N-Heterocyclic Carbenes. Angew. Chem. Int. Ed. Engl. 1997, 36, 2162–2187. [Google Scholar] [CrossRef]
  5. Bourissou, D.; Guerret, O.; Gabbaï, F.P.; Bertrand, G. Stable Carbenes. Chem. Rev. 2000, 100, 39–91. [Google Scholar] [CrossRef] [PubMed]
  6. Bertrande, G. Carbene Chemistry: From Fleeting Intermediates to Powerful Reagents; FontisMedia S.A.: Lausanne, Switzerland; Marcel Dekker, Inc.: New York, NY, USA, 2002. [Google Scholar]
  7. Moss, R.A.; Platz, M.S.; Jones, M., Jr. (Eds.) Reactive Intermediate Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2005. [Google Scholar]
  8. Scott, N.M.; Nolan, S.P. Stabilization of Organometallic Species Achieved by the Use of N-Heterocyclic Carbene (NHC) Ligands. Eur. J. Inorg. Chem. 2005, 2005, 1815–1828. [Google Scholar] [CrossRef]
  9. Nolan, S.P. N-Heterocyclic Carbenes in Synthesis; Wiley-VCH: Weinheim, Germany, 2006. [Google Scholar]
  10. Carey, F.A.; Sundberg, R.J. Carbenes, Part B: Reactions and Synthesis. Advanced Organic Chemistry; Springer: New York, NY, USA, 2007. [Google Scholar]
  11. Díez-González, S.; Nolan, S.P. Stereoelectronic parameters associated with N-heterocyclic carbene (NHC) ligands: A quest for understanding. Coord. Chem. Rev. 2007, 251, 874–883. [Google Scholar] [CrossRef]
  12. Jacobsen, H.; Correa, A.; Poater, A.; Costabile, C.; Cavallo, L. Understanding the M–(NHC) (NHC = N-heterocyclic carbene) bond. Coord. Chem. Rev. 2009, 253, 687–703. [Google Scholar] [CrossRef]
  13. de Frémont, P.; Marion, N.; Nolan, S.P. Carbenes: Synthesis, properties, and organometallic chemistry. Coord. Chem. Rev. 2009, 253, 862–892. [Google Scholar] [CrossRef]
  14. Moss, R.A.; Doyle, M.P. Contemporary Carbene Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2013. [Google Scholar]
  15. Nelson, D.J.; Nolan, S.P. Quantifying and understanding the electronic properties of N-heterocyclic carbenes. Chem. Soc. Rev. 2013, 42, 6723–6753. [Google Scholar] [CrossRef]
  16. Bellemin-Laponnaz, S.; Dagorne, S. Group 1 and 2 and Early Transition Metal Complexes Bearing N-Heterocyclic Carbene Ligands: Coordination Chemistry, Reactivity, and Applications. Chem. Rev. 2014, 114, 8747–8774. [Google Scholar] [CrossRef]
  17. Hopkinson, M.N.; Richter, C.; Schedler, M.; Glorius, F. An overview of N-heterocyclic carbenes. Nature 2014, 510, 485–496. [Google Scholar] [CrossRef]
  18. Nesterov, V.; Reiter, D.; Bag, P.; Frisch, P.; Holzner, R.; Porzelt, A.; Inoue, S. NHCs in Main Group Chemistry. Chem. Rev. 2018, 118, 9678–9842. [Google Scholar] [CrossRef]
  19. Pople, J.A.; Raghavachari, K.; Frisch, M.J.; Binkley, J.S.; Schleyer, P.v.R. Comprehensive Theoretical Study of Isomers and Rearrangement Barriers of Even Electron Polyatomic Molecules HmABHn (A, B = C, N, O, and F). J. Am. Chem. Soc. 1983, 105, 6389–6398. [Google Scholar] [CrossRef]
  20. Pople, J.A. A theoretical search for the methylenefluoronium ylide. Chem. Phys. Lett. 1986, 132, 144–146. [Google Scholar] [CrossRef]
  21. Arduengo, A.J., III; Gamper, S.F.; Tamm, M.; Calabrese, J.C.; Davidson, F.; Craig, H.A. A Bis(carbene)–Proton Complex: Structure of a C–H–C Hydrogen Bond. J. Am. Chem. Soc. 1995, 117, 572–573. [Google Scholar] [CrossRef]
  22. Alkorta, I.; Elguero, J. Carbenes and Silylenes as Hydrogen Bond Acceptors. J. Phys. Chem. 1996, 100, 19367–19370. [Google Scholar] [CrossRef]
  23. Jabłoński, M.; Palusiak, M. Divalent carbon atom as the proton acceptor in hydrogen bonding. Phys. Chem. Chem. Phys. 2009, 11, 5711–5719. [Google Scholar] [CrossRef]
  24. Giffin, N.A.; Makramalla, M.; Hendsbee, A.D.; Robertson, K.N.; Sherren, C.; Pye, C.C.; Masuda, J.D.; Clyburne, J.A.C. Anhydrous TEMPO-H: Reactions of a good hydrogen atom donor with low-valent carbon centres. Org. Biomol. Chem. 2011, 9, 3672–3680. [Google Scholar] [CrossRef] [Green Version]
  25. Gerbig, D.; Ley, D. Computational methods for contemporary carbene chemistry. WIREs Comput. Mol. Sci. 2013, 3, 242–272. [Google Scholar] [CrossRef]
  26. Samanta, R.C.; De Sarkar, S.; Fröhlich, R.; Grimme, S.; Studer, A. N-Heterocyclic carbene (NHC) catalyzed chemoselective acylation of alcohols in the presence of amines with various acylating reagents. Chem. Sci. 2013, 4, 2177–2184. [Google Scholar] [CrossRef]
  27. Jabłoński, M. Coexistence of the Carbene⋯H-D Hydrogen Bond and Other Accompanying Interactions in Forty Dimers of N-Heterocyclic-Carbenes (I, IMe2, IiPr2, ItBu2, IMes2, IDipp2, IAd2; I = imidazol-2-ylidene) and Some Fundamental Proton Donors (HF, HCN, H2O, MeOH, NH3). Molecules 2022, 27, 5712. [Google Scholar] [CrossRef]
  28. Li, Q.; Wang, H.; Liu, Z.; Li, W.; Cheng, J.; Gong, B.; Sun, J. Ab Initio Study of Lithium-Bonded Complexes with Carbene as an Electron Donor. J. Phys. Chem. A 2009, 113, 14156–14160. [Google Scholar] [CrossRef]
  29. Wang, Y.; Xie, Y.; Abraham, M.Y.; Wei, P.; Schaefer, H.F., III; Schleyer, P.v.R.; Robinson, G.H. A Viable Anionic N-Heterocyclic Dicarbene. J. Am. Chem. Soc. 2010, 132, 14370–14372. [Google Scholar] [CrossRef] [PubMed]
  30. Zhi-Feng, L.; Sheng, Y.; Hui-Xue, L. Theoretical prediction characters of unconventional weak bond with carbene as electron donors and Li–Y (Y = OH, H, F, NC and CN) as electron acceptors. J. Mol. Struct. THEOCHEM 2010, 952, 56–60. [Google Scholar] [CrossRef]
  31. Herrmann, W.A.; Runte, O.; Artus, G. Synthesis and structure of an ionic beryllium—“Carbene” complex. J. Organomet. Chem. 1995, 501, C1–C4. [Google Scholar] [CrossRef]
  32. Gilliard, R.J., Jr.; Abraham, M.Y.; Wang, Y.; Wei, P.; Xie, Y.; Quillian, B.; Schaefer, H.F., III; Schleyer, P.v.R.; Robinson, G.H. Carbene-Stabilized Beryllium Borohydride. J. Am. Chem. Soc. 2012, 134, 9953–9955. [Google Scholar] [CrossRef] [PubMed]
  33. Arrowsmith, M.; Hill, M.S.; Kociok-Köhn, G.; MacDougall, D.J.; Mahon, M.F. Beryllium-Induced C–N Bond Activation and Ring Opening of an N-Heterocyclic Carbene. Angew. Chem. Int. Ed. 2012, 51, 2098–2100. [Google Scholar] [CrossRef]
  34. Walley, J.E.; Wong, Y.-O.; Freeman, L.A.; Dickie, D.A.; Gilliard, R.J., Jr. N-Heterocyclic Carbene-Supported Aryl- and Alk- oxides of Beryllium and Magnesium. Catalysts 2019, 9, 934. [Google Scholar] [CrossRef] [Green Version]
  35. Jabłoński, M. Study of Beryllium, Magnesium, and Spodium Bonds to Carbenes and Carbodiphosphoranes. Molecules 2021, 26, 2275. [Google Scholar] [CrossRef]
  36. Arduengo, A.J., III; Dias, H.V.R.; Davidson, F.; Harlow, R.L. Carbene adducts of magnesium and zinc. J. Organomet. Chem. 1993, 462, 13–18. [Google Scholar] [CrossRef]
  37. Arrowsmith, M.; Hill, M.S.; MacDougall, D.J.; Mahon, M.F. A Hydride-Rich Magnesium Cluster. Angew. Chem. Int. Ed. 2009, 48, 4013–4016. [Google Scholar] [CrossRef]
  38. Arnold, P.L.; Casely, I.J.; Turner, Z.R.; Bellabarba, R.; Tooze, R.B. Magnesium and zinc complexes of functionalised, saturated N-heterocyclic carbene ligands: Carbene lability and functionalisation, and lactide polymerisation catalysis. Dalton Trans. 2009, 7236–7247. [Google Scholar] [CrossRef]
  39. Arduengo, A.J., III; Dias, H.V.R.; Calabrese, J.C.; Davidson, F. A Stable Carbene-Alane Adduct. J. Am. Chem. Soc. 1992, 114, 9724–9725. [Google Scholar] [CrossRef]
  40. Li, X.-W.; Su, J.; Robinson, G.H. Syntheses and molecular structure of organo-group 13 metal carbene complexes. Chem. Commun. 1996, 23, 2683–2684. [Google Scholar] [CrossRef]
  41. Hibbs, D.E.; Hursthouse, M.B.; Jones, C.; Smithies, N.A. Synthesis, crystal and molecular structure of the first indium trihydride complex, [InH3{CN(Pri)C2Me2N(Pri)}]. Chem. Commun. 1998, 8, 869–870. [Google Scholar] [CrossRef]
  42. Merceron, N.; Miqueu, K.; Baceiredo, A.; Bertrand, G. Stable (Amino)(phosphino)carbenes: Difunctional Molecules. J. Am. Chem. Soc. 2002, 124, 6806–6807. [Google Scholar] [CrossRef]
  43. Wang, Y.; Robinson, G.H. Unique homonuclear multiple bonding in main group compounds. Chem. Commun. 2009, 5201–5213. [Google Scholar] [CrossRef]
  44. Del Bene, J.E.; Alkorta, I.; Elguero, J. Carbon–Carbon Bonding between Nitrogen Heterocyclic Carbenes and CO2. J. Phys. Chem. A 2017, 121, 8136–8146. [Google Scholar] [CrossRef]
  45. Liu, M.; Li, Q.; Li, W.; Cheng, J. Carbene tetrel-bonded complexes. Struct. Chem. 2017, 28, 823–831. [Google Scholar] [CrossRef]
  46. Wang, Y.; Xie, Y.; Abraham, M.Y.; Gilliard, R.J., Jr.; Wei, P.; Schaefer, H.F., III; Schleyer, P.v.R.; Robinson, G.H. Carbene-Stabilized Parent Phosphinidene. Organometallics 2010, 29, 4778–4780. [Google Scholar] [CrossRef]
  47. Abraham, M.Y.; Wang, Y.; Xie, Y.; Wei, P.; Schaefer, H.F., III; Schleyer, P.v.R.; Robinson, G.H. Carbene Stabilization of Diarsenic: From Hypervalency to Allotropy. Chem. Eur. J. 2010, 16, 432–435. [Google Scholar] [CrossRef]
  48. Patel, D.S.; Bharatam, P.V. Divalent N(I) Compounds with Two Lone Pairs on Nitrogen. J. Phys. Chem. A 2011, 115, 7645–7655. [Google Scholar] [CrossRef]
  49. Zhao, Q.; Feng, D.; Sun, Y.; Hao, J.; Cai, Z. Theoretical Investigations on the Weak Nonbonded C=S⋯CH2 Interactions: Chalcogen-Bonded Complexes With Singlet Carbene as an Electron Donor. Int. J. Quant. Chem. 2011, 111, 3881–3887. [Google Scholar] [CrossRef]
  50. Arduengo, A.J., III; Kline, M.; Calabrese, J.C.; Davidson, F. Synthesis of a Reverse Ylide from a Nucleophilic Carbene. J. Am. Chem. Soc. 1991, 113, 9704–9705. [Google Scholar] [CrossRef]
  51. Kuhn, N.; Kratz, T.; Henkel, G. A Stable Carbene Iodine Adduct: Secondary Bonding in 1,3-Diethyl-2-iodo-4,5-dimethylimidazolium Iodide. J. Chem. Soc. Chem. Commun. 1993, 1778–1779. [Google Scholar] [CrossRef]
  52. Li, Q.; Wang, Y.; Liu, Z.; Li, W.; Cheng, J.; Gong, B.; Sun, J. An unconventional halogen bond with carbene as an electron donor: An ab initio study. Chem. Phys. Lett. 2009, 469, 48–51. [Google Scholar] [CrossRef]
  53. Esrafili, M.D.; Mohammadirad, N. Insights into the strength and nature of carbene⋯halogen bond interactions: A theoretical perspective. J. Mol. Model. 2013, 19, 2559–2566. [Google Scholar] [CrossRef] [PubMed]
  54. Esrafili, M.D.; Sabouri, A. Carbene–aerogen bonds: An Ab Initio Study. Mol. Phys. 2017, 115, 971–980. [Google Scholar] [CrossRef]
  55. Cases, M.; Frenking, G.; Duran, M.; Solà, M. Molecular Structure and Bond Characterization of the Fischer-Type Chromium–Carbene Complexes (CO)5Cr=C(X)R (X = H, OH, OCH3, NH2, NHCH3 and R = H, CH3, CH=CH2, Ph, C≡CH). Organometallics 2002, 21, 4182–4191. [Google Scholar] [CrossRef]
  56. Nemcsok, D.; Wichmann, K.; Frenking, G. The Significance of π Interactions in Group 11 Complexes with N-Heterocyclic Carbenes. Organometallics 2004, 23, 3640–3646. [Google Scholar] [CrossRef]
  57. Frenking, G.; Solà, M.; Vyboishchikov, S.F. Chemical bonding in transition metal carbene complexes. J. Organomet. Chem. 2005, 690, 6178–6204. [Google Scholar] [CrossRef]
  58. Tonner, R.; Heydenrych, G.; Frenking, G. Bonding Analysis of N-Heterocyclic Carbene Tautomers and Phosphine Ligands in Transition-Metal Complexes: A Theoretical Study. Chem. Asian J. 2007, 2, 1555–1567. [Google Scholar] [CrossRef]
  59. Radius, U.; Bickelhaupt, F.M. Bonding of Imidazol-2-ylidene Ligands in Nickel Complexes. Organometallics 2008, 27, 3410–3414. [Google Scholar] [CrossRef]
  60. Srebro, M.; Michalak, A. Theoretical Analysis of Bonding in N-Heterocyclic Carbene–Rhodium Complexes. Inorg. Chem. 2009, 48, 5361–5369. [Google Scholar] [CrossRef]
  61. Andrada, D.M.; Holzmann, N.; Hamadi, T.; Frenking, G.; Beilstein, J. Direct estimate of the internal π-donation to the carbene centre within N-heterocyclic carbenes and related molecules. Org. Chem. 2015, 11, 2727–2736. [Google Scholar] [CrossRef]
  62. Santoro, O.; Nahra, F.; Cordes, D.B.; Slawin, A.M.Z.; Nolan, S.P.; Cazin, C.S.J. Synthesis, characterization and catalytic activity of stable [(NHC)H][ZnXY2] (NHC = N-Heterocyclic carbene, X, Y = Cl, Br) species. J. Mol. Catal. 2016, 423, 85–91. [Google Scholar] [CrossRef] [Green Version]
  63. Dagorne, S. Recent Developments on N-Heterocyclic Carbene Supported Zinc Complexes: Synthesis and Use in Catalysis. Synthesis 2018, 50, 3662–3670. [Google Scholar] [CrossRef]
  64. Procter, R.J.; Uzelac, M.; Cid, J.; Rushworth, P.J.; Ingleson, M.J. Low-Coordinate NHC–Zinc Hydride Complexes Catalyze Alkyne C–H Borylation and Hydroboration Using Pinacolborane. ACS Catal. 2019, 9, 5760–5771. [Google Scholar] [CrossRef]
  65. Specklin, D.; Fliedel, C.; Dagorne, S. Recent Representative Advances on the Synthesis and Reactivity of N-Heterocyclic-Carbene-Supported Zinc Complexes. Chem. Rec. 2021, 21, 1130–1143. [Google Scholar] [CrossRef]
  66. Arduengo, A.J., III; Davidson, F.; Krafczyk, R.; Marshall, W.J.; Tamm, M. Adducts of Carbenes with Group II and XII Metallocenes. Organometallics 1998, 17, 3375–3382. [Google Scholar] [CrossRef]
  67. Abernethy, C.D.; Baker, R.J.; Cole, M.L.; Davies, A.J.; Jones, C. Reactions of a carbene stabilised indium trihydride complex, [InH3{CN(Mes)-C2H2N(Mes)}] Mes = mesityl, with transition metal complexes. Trans. Met. Chem. 2003, 28, 296–299. [Google Scholar] [CrossRef]
  68. Wang, D.; Wurst, K.; Buchmeiser, M.R. N-heterocyclic carbene complexes of Zn(II): Synthesis, X-ray structures and reactivity. J. Organomet. Chem. 2004, 689, 2123–2130. [Google Scholar] [CrossRef]
  69. Jensen, T.R.; Breyfogle, L.E.; Hillmyer, M.A.; Tolman, W.B. Stereoelective polymerization of D,L-lactide using N-heterocyclic carbene based compounds. Chem. Commun. 2004, 2504–2505. [Google Scholar] [CrossRef] [PubMed]
  70. Jensen, T.R.; Schaller, C.P.; Hillmyer, M.A.; Tolman, W.B. Zinc N-heterocyclic carbene complexes and their polymerization of D,L-lactide. J. Organomet. Chem. 2005, 690, 5881–5891. [Google Scholar] [CrossRef]
  71. Anantharaman, G.; Elango, K. Synthesis and Characterization of NHC-Stabilized Zinc Aryloxide and Zinc Hydroxyaryloxide. Organometallics 2007, 26, 1089–1092. [Google Scholar] [CrossRef]
  72. Budagumpi, S.; Endud, S. Group XII Metal–N-Heterocyclic Carbene Complexes: Synthesis, Structural Diversity, Intramolecular Interactions, and Applications. Organometallics 2013, 32, 1537–1562. [Google Scholar] [CrossRef]
  73. Schnee, G.; Fliedel, C.; Avilés, T.; Dagorne, S. Neutral and Cationic N-Heterocyclic Carbene Zinc Adducts and the BnOH/Zn(C6F5)2 Binary Mixture–Characterization and Use in the Ring-Opening Polymerization of β-Butyrolactone, Lactide, and Trimethylene Carbonate. Eur. J. Inorg. Chem. 2013, 2013, 3699–3709. [Google Scholar] [CrossRef]
  74. Fliedel, C.; Vila-Viçosa, D.; Calhorda, M.J.; Dagorne, S.; Avilés, T. Dinuclear Zinc–N-Heterocyclic Carbene Complexes for Either the Controlled Ring-Opening Polymerization of Lactide or the Controlled Degradation of Polylactide Under Mild Conditions. ChemCatChem 2014, 6, 1357–1367. [Google Scholar] [CrossRef]
  75. Collins, L.R.; Moffat, L.A.; Mahon, M.F.; Jones, M.D.; Whittlesey, M.K. Lactide polymerisation by ring-expanded NHC complexes of zinc. Polyhedron 2016, 103, 121–125. [Google Scholar] [CrossRef] [Green Version]
  76. Tian, J.; Chen, Y.; Vayer, M.; Djurovic, A.; Guillot, R.; Guermazi, R.; Dagorne, S.; Bour, C.; Gandon, V. Exploring the Limits of π-Acid Catalysis Using Strongly Electrophilic Main Group Metal Complexes: The Case of Zinc and Aluminium. Chem. Eur. J. 2020, 26, 12831–12838. [Google Scholar] [CrossRef]
  77. Roy, M.M.D.; Baird, S.R.; Ferguson, M.J.; Rivard, E. Toward N-heterocyclic carbene stabilized zinc sulfides. Mendeleev Commun. 2021, 31, 173–175. [Google Scholar] [CrossRef]
  78. Jabłoński, M. Theoretical Study of N-Heterocyclic-Carbene–ZnX2 (X = H, Me, Et) Complexes. Materials 2021, 14, 6147. [Google Scholar] [CrossRef]
  79. Schoeller, W.W. Electrophilicity and nucleophilicity in singlet carbenes. II. Electrophilic selectivity. Tetrahedron Lett. 1980, 21, 1509–1510. [Google Scholar] [CrossRef]
  80. Goumri-Magnet, S.; Polishchuk, O.; Gornitzka, H.; Marsden, C.J.; Baceiredo, A.; Bertrand, G. The Electrophilic Behavior of Stable Phosphanylcarbenes Towards Phosphorus Lone Pairs. Angew. Chem. Int. Ed. 1999, 38, 3727–3729. [Google Scholar] [CrossRef]
  81. Moss, R.A.; Wang, L.; Cang, H.; Krogh-Jespersen, K. Extremely reactive carbenes: Electrophiles and nucleophiles. J. Phys. Org. Chem. 2017, 30, e3555. [Google Scholar] [CrossRef]
  82. Jabłoński, M. The first theoretical proof of the existence of a hydride-carbene bond. Chem. Phys. Lett. 2018, 710, 78–83. [Google Scholar] [CrossRef]
  83. Jabłoński, M. In search for a hydride-carbene bond. J. Phys. Org. Chem. 2019, 32, e3949. [Google Scholar] [CrossRef]
  84. Yourdkhani, S.; Jabłoński, M. Physical nature of silane⋯carbene dimers revealed by state-of-the-art ab initio calculations. J. Comput. Chem. 2019, 40, 2643–2652. [Google Scholar] [CrossRef]
  85. Yáñez, M.; Sanz, P.; Mó, O.; Alkorta, I.; Elguero, J. Beryllium Bonds, Do They Exists? J. Chem. Theory Comput. 2009, 5, 2763–2771. [Google Scholar] [CrossRef]
  86. Martín-Sómer, A.; Lamsabhi, A.M.; Mó, O.; Yáñez, M. The importance of deformation on the strength of beryllium bonds. Comput. Theor. Chem. 2012, 998, 74–79. [Google Scholar] [CrossRef]
  87. Eskandari, K. Characteristics of beryllium bonds: A QTAIM study. J. Mol. Model. 2012, 18, 3481–3487. [Google Scholar] [CrossRef]
  88. Villanueva, E.F.; Mó, O.; Yáñez, M. On the existence and characteristics of π-beryllium bonds. Phys. Chem. Chem. Phys. 2014, 16, 17531–17536. [Google Scholar] [CrossRef]
  89. Zhong, A.; Chen, D.; Li, R. Revisiting the beryllium bonding interactions from energetic and wavefunction perspectives. Chem. Phys. Lett. 2015, 633, 265–272. [Google Scholar] [CrossRef]
  90. Eskandari, K. Nature of beryllium bonds in view of interacting quantum atoms and natural energy decomposition analysis. Comput. Theor. Chem. 2016, 1090, 74–79. [Google Scholar] [CrossRef]
  91. Montero-Campillo, M.M.; Mó, O.; Yáñez, M.; Alkorta, I.; Elguero, J. The beryllium bond. Adv. Inorg. Chem. 2019, 73, 73–121. [Google Scholar]
  92. 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]
  93. Jabłoński, M. On the Uselessness of Bond Paths Linking Distant Atoms and on the Violation of the Concept of Privileged Exchange Channels. Chem. Open 2019, 8, 497–507. [Google Scholar] [CrossRef] [Green Version]
  94. Yang, X.; Li, Q.; Cheng, J.; Li, W. A new interaction mechanism of LiNH2 with MgH2: Magnesium bond. J. Mol. Model. 2013, 19, 247–253. [Google Scholar] [CrossRef]
  95. Xu, H.-L.; Li, Q.-Z.; Scheiner, S. Effect of magnesium bond on the competition between hydrogen and halogen bonds and the induction of proton and halogen transfer. ChemPhysChem 2018, 19, 1456–1464. [Google Scholar] [CrossRef] [Green Version]
  96. Sanz, P.; Montero-Campillo, M.M.; Mó, O.; Yáñez, M.; Alkorta, I.; Elguero, J. Intramolecular magnesium bonds in malonaldehyde-like systems: A critical view of the resonance-assisted phenomena. Theor. Chem. Acc. 2018, 137, 97. [Google Scholar] [CrossRef] [Green Version]
  97. Lupinetti, A.J.; Jonas, V.; Thiel, W.; Strauss, S.H.; Frenking, G. Trends in Molecular Geometries and Bond Strengths of the Homoleptic d10 Metal Carbonyl Cations [M(CO)n]x+ (Mx+ = Cu+, Ag+, Au+, Zn2+, Cd2+, Hg2+; n = 1–6): A Theoretical Study. Chem. Eur. J. 1999, 5, 2573–2583. [Google Scholar] [CrossRef]
  98. Joy, J.; Jemmis, E.D. Contrasting Behavior of the Z Bonds in X–Z⋯Y Weak Interactions: Z = Main Group Elements Versus the Transition Metals. Inorg. Chem. 2017, 56, 1132–1143. [Google Scholar] [CrossRef]
  99. Wang, S.R.; Arrowsmith, M.; Braunschweig, H.; Dewhurst, R.D.; Dömling, M.; Mattock, J.D.; Pranckevicius, C.; Vargas, A. Monomeric 16-Electron π-Diborene Complexes of Zn(II) and Cd(II). J. Am. Chem. Soc. 2017, 139, 10661–10664. [Google Scholar] [CrossRef]
  100. Kalhor, P.; Wang, Y.; Yu, Z. The Structures of ZnCl2-Ethanol Mixtures, a Spectroscopic and Quantum Chemical Calculation Study. Molecules 2021, 26, 2498. [Google Scholar] [CrossRef]
  101. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Oxford University Press: New York, NY, USA, 1990. [Google Scholar]
  102. Cremer, D.; Kraka, E. Chemical Bonds without Bonding Electron Density–Does the Difference Electron-Density Analysis Suffice for a Description of the Chemical Bond? Angew. Chem. Int. Ed. Engl. 1984, 23, 627–628. [Google Scholar] [CrossRef]
  103. Bader, R.F.W.; Stephens, M.E. Spatial Localization of the Electronic Pair and Number Distributions in Molecules. J. Am. Chem. Soc. 1975, 97, 7391–7399. [Google Scholar] [CrossRef]
  104. Fradera, X.; Austen, M.A.; Bader, R.F.W. The Lewis Model and Beyond. J. Phys. Chem. A 1999, 103, 304–314. [Google Scholar] [CrossRef]
  105. Fradera, X.; Poater, J.; Simon, S.; Duran, M.; Solà, M. Electron-pairing analysis from localization and delocalization indices in the framework of the atoms-in-molecules theory. Theor. Chem. Acc. 2002, 108, 214–224. [Google Scholar] [CrossRef]
  106. Firme, C.L.; Antunes, O.A.C.; Esteves, P.M. Relation between bond order and delocalization index of QTAIM. Chem. Phys. Lett. 2009, 468, 129–133. [Google Scholar] [CrossRef]
  107. Rafat, M.; Popelier, P.L.A. The Quantum Theory of Atoms in Molecules: From Solid State to DNA and Drug Design; Matta, C.F., Boyed, R.J., Eds.; Wiley-VCH: Weinheim, Germany, 2007; pp. 121–140. [Google Scholar]
  108. García-Revilla, M.; Francisco, E.; Popelier, P.L.A.; Pendás, A.M. Domain-Averaged Exchange-Correlation Energies as a Physical Underpinning for Chemical Graphs. ChemPhysChem 2013, 14, 1211–1218. [Google Scholar] [CrossRef]
  109. Johnson, E.R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A.J.; Yang, W. Revealing Noncovalent Interactions. J. Am. Chem. Soc. 2010, 132, 6498–6506. [Google Scholar] [CrossRef] [Green Version]
  110. Contreras-García, J.; Johnson, E.R.; Keinan, S.; Chaudret, R.; Piquemal, J.-P.; Beratan, D.N.; Yang, W. NCIPLOT: A Program for Plotting Noncovalent Interaction Regions. J. Chem. Theory Comput. 2011, 7, 625–632. [Google Scholar] [CrossRef]
  111. Blanco, M.A.; Pendás, A.M.; Francisco, E. Interacting Quantum Atoms: A Correlated Energy Decomposition Scheme Based on the Quantum Theory of Atoms in Molecules. J. Chem. Theory Comput. 2005, 1, 1096–1109. [Google Scholar] [CrossRef] [PubMed]
  112. Guevara-Vela, J.M.; Francisco, E.; Rocha-Rinza, T.; Pendás, A.M. Interacting Quantum Atoms–A Review. Molecules 2020, 25, 4028. [Google Scholar] [CrossRef]
  113. Ziegler, T.; Rauk, A. On the calculation of bonding energies by the Hartree Fock Slater method. Theoret. Chim. Acta (Berl.) 1977, 46, 1–10. [Google Scholar]
  114. Mitoraj, M.; Michalak, A. Donor–Acceptor Properties of Ligands from the Natural Orbitals for Chemical Valence. Organometallics 2007, 26, 6576–6580. [Google Scholar] [CrossRef]
  115. Mitoraj, M.; Michalak, A. Natural orbitals for chemical valence as descriptors of chemical bonding in transition metal complexes. J. Mol. Model. 2007, 13, 347–355. [Google Scholar] [CrossRef] [PubMed]
  116. Michalak, A.; Mitoraj, M.; Ziegler, T. Bond Orbitals from Chemical Valence Theory. J. Phys. Chem. A 2008, 112, 1933–1939. [Google Scholar] [CrossRef]
  117. Mitoraj, M.P.; Michalak, A.; Ziegler, T. A Combined Charge and Energy Decomposition Scheme for Bond Analysis. J. Chem. Theory Comput. 2009, 5, 962–975. [Google Scholar] [CrossRef]
  118. Mitoraj, M.P.; Michalak, A.; Ziegler, T. On the Nature of the Agostic Bond between Metal Centers and β-Hydrogen Atoms in Alkyl Complexes. An Analysis Based on the Extended Transition State Method and the Natural Orbitals for Chemical Valence Scheme (ETS-NOCV). Organometallics 2009, 28, 3727–3733. [Google Scholar] [CrossRef]
  119. Saebø, S.; Pulay, P. Local Treatment of Electron Correlation. Annu. Rev. Phys. Chem 1993, 44, 213–236. [Google Scholar] [CrossRef]
  120. Neese, F.; Wennmohs, F.; Hansen, A. Efficient and accurate local approximations to coupled-electron pair approaches: An attempt to revive the pair natural orbital method. J. Chem. Phys. 2009, 130, 114108. [Google Scholar] [CrossRef]
  121. Neese, F.; Hansen, A.; Liakos, D.G. Efficient and accurate approximations to the local coupled cluster singles doubles method using a truncated pair natural orbital basis. J. Chem. Phys. 2009, 131, 064103. [Google Scholar] [CrossRef]
  122. Riplinger, C.; Sandhoefer, B.; Hansen, A.; Neese, F. Natural triple excitations in local coupled cluster calculations. J. Chem. Phys. 2013, 139, 134101. [Google Scholar] [CrossRef]
  123. Guo, Y.; Riplinger, C.; Becker, U.; Liakos, D.G.; Minenkov, Y.; Cavallo, L.; Neese, F. Communication: An improved linear scaling perturbative triples correction for the domain based local pair-natural orbital based singles and doubles coupled cluster method [DLPNO-CCSD(T)]. J. Chem. Phys. 2018, 148, 011101. [Google Scholar] [CrossRef]
  124. Altun, A.; Saitow, M.; Neese, F.; Bistoni, G. Local Energy Decomposition of Open-Shell Molecular Systems in the Domain-Based Local Pair Natural Orbital Coupled Cluster Framework. J. Chem. Theory Comput. 2019, 15, 1616–1632. [Google Scholar] [CrossRef] [Green Version]
  125. Altun, A.; Izsák, R.; Bistoni, G. Local energy decomposition of coupled-cluster interaction energies: Interpretation, benchmarks, and comparison with symmetry-adapted perturbation theory. Int. J. Quantum Chem. 2021, 121, e26339. [Google Scholar] [CrossRef]
  126. Hirshfeld, F.L. Bonded-Atom Fragments for Describing Molecular Charge Densities. Theor. Chim. Acta 1977, 44, 129–138. [Google Scholar] [CrossRef]
  127. Ritchie, J.P.; Bachrach, S.M. Some Methods and Applications of Electron Density Distribution Analysis. J. Comput. Chem. 1987, 8, 499–509. [Google Scholar] [CrossRef]
  128. Ritchie, J.P. Electron Density Distribution Analysis for Nitromethane, Nitromethide, and Nitramide. J. Am. Chem. Soc. 1985, 107, 1829–1837. [Google Scholar] [CrossRef]
  129. Wiberg, K.B.; Rablen, P.R. Atomic Charges. J. Org. Chem. 2018, 83, 15463–15469. [Google Scholar] [CrossRef]
  130. Jabłoński, M.; Krygowski, T.M. Study of the influence of intermolecular interaction on classical and reverse substituent effects in para-substituted phenylboranes. New J. Chem. 2020, 44, 9656–9670. [Google Scholar] [CrossRef]
  131. Walsh, A.D. The Electronic Orbitals, Shapes, and Spectra of Polyatomic Molecules. Part I. AH2 Molecules. J. Chem. Soc. 1953, 2260–2266. [Google Scholar] [CrossRef]
  132. Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom–atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef] [PubMed]
  133. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef] [Green Version]
  134. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef] [Green Version]
  135. Parr, R.G.; Yang, W. Density-Functional Theory of Atoms and Molecules; Oxford University Press: New York, NY, USA, 1989. [Google Scholar]
  136. Mardirossian, N.; Head-Gordon, M. Thirty years of density functional theory in computational chemistry: An overview and extensive assessment of 200 density functionals. Mol. Phys. 2017, 115, 2315–2372. [Google Scholar] [CrossRef] [Green Version]
  137. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  138. McLean, A.D.; Chandler, G.S. Contracted Gaussian basis sets for molecular calculations. I. second row atoms, Z = 11–18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  139. Clark, T.; Chandrasekhar, J.; Spitznagel, G.W.; Schleyer, P.v.R. Efficient Diffuse Function-Augmented Basis Sets for Anion Calculations. III. The 3-21+G Basis Set for First-Row Elements, Li–F. J. Comput. Chem. 1983, 4, 294–301. [Google Scholar] [CrossRef]
  140. Frisch, M.J.; Pople, J.A.; Binkley, J.S. Self-consistent molecular orbital methods 25. Supplementary functions for Gaussian basis sets. J. Chem. Phys. 1984, 80, 3265–3269. [Google Scholar] [CrossRef]
  141. Curtiss, L.A.; McGrath, M.P.; Blandeau, J.-P.; Davis, N.E.; Binning, R.C., Jr.; Radom, L. Extension of Gaussian-2 theory to molecules containing third-row atoms Ga–Kr. J. Chem. Phys. 1995, 103, 6104–6113. [Google Scholar] [CrossRef]
  142. Jensen, F. Introduction to Computational Chemistry; John Wiley & Sons Ltd.: Chichester, UK, 2007. [Google Scholar]
  143. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus, T.L. New Basis Set Exchange: An Open, Up-to-Date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59, 4814–4820. [Google Scholar] [CrossRef]
  144. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  145. Keith, T.A. AIMAll (Version 15.05.18); TK Gristmill Software: Overland Park, KS, USA, 2015; Available online: Aim.tkgristmill.com (accessed on 17 September 2022).
  146. Møller, C.; Plesset, M.S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46, 618–622. [Google Scholar] [CrossRef]
  147. Head-Gordon, M.; Pople, J.A.; Frisch, M.J. MP2 Energy Evaluation by Direct Methods. Chem. Phys. Lett. 1988, 153, 503–506. [Google Scholar] [CrossRef]
  148. Becke, A.D. Density-fnnctional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  149. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  150. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  151. van Lenthe, E.; Baerends, E.J.; Snijder, J.G. Relativistic regular two-component Hamiltonians. J. Chem. Phys. 1993, 99, 4597–4610. [Google Scholar] [CrossRef]
  152. Snijder, J.G.; Sadlej, A.J. Perturbation versus variation treatment of regular relativistic Hamiltonians. Chem. Phys. Lett. 1996, 252, 51–61. [Google Scholar] [CrossRef]
  153. Baerends, E.J.; Ziegler, T.; Atkins, A.J.; Autschbach, J.; Baseggio, O.; Bashford, D.; Bééces, A.; Bickelhaupt, F.M.; Bo, C.; Boerrigter, P.M.; et al. ADF 2022.1; SCM-Theoretical Chemistry; Vrije Universiteit: Amsterdam, The Netherlands, 2022. [Google Scholar]
  154. Sagan, F.; Mitoraj, M.P. Non-covalent Interactions in Selected Transition Metal Complexes. In Transition Metals in Coordination Environments: Computational Chemistry and Catalysis Viewpoints; Broclawik, E., Borowski, T., Radoń, M., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 65–89. [Google Scholar]
  155. Stasyuk, O.A.; Sedlak, R.; Fonseca-Guerra, C.; Hobza, P. Comparison of the DFT-SAPT and Canonical EDA Schemes for the Energy Decomposition of Various Types of Noncovalent Interactions. J. Chem. Theory Comput. 2018, 14, 3440–3450. [Google Scholar] [CrossRef]
Figure 1. Fully optimized cyclopropenylidene⋯ZnBr 2 (left) and imidazol-2-ylidene⋯ZnBr 2 (right) dimer structures.
Figure 1. Fully optimized cyclopropenylidene⋯ZnBr 2 (left) and imidazol-2-ylidene⋯ZnBr 2 (right) dimer structures.
Ijms 23 14668 g001
Figure 2. Dominant NOCV-based deformation density contributions determined for imidazol-2-ylidene⋯BeBr 2 dimer.
Figure 2. Dominant NOCV-based deformation density contributions determined for imidazol-2-ylidene⋯BeBr 2 dimer.
Ijms 23 14668 g002
Figure 3. Relation between the relative energy (i.e., determined in relation to the minimum for a given dimer) (left) or the distance d C Zn (right) and the twist angle of the ZnBr 2 unit plane with respect to the carbene plane for dimers cyclopropenylidene⋯ZnBr 2 (cpy-ZnBr 2 ) and imidazol-2-ylidene⋯ZnBr 2 (imi-ZnBr 2 ).
Figure 3. Relation between the relative energy (i.e., determined in relation to the minimum for a given dimer) (left) or the distance d C Zn (right) and the twist angle of the ZnBr 2 unit plane with respect to the carbene plane for dimers cyclopropenylidene⋯ZnBr 2 (cpy-ZnBr 2 ) and imidazol-2-ylidene⋯ZnBr 2 (imi-ZnBr 2 ).
Ijms 23 14668 g003
Figure 4. Dependence of the orbital energy of the three highest occupied orbitals (i.e., HOMO, HOMO-1 and HOMO-2) on the twist angle of the ZnBr 2 unit plane with respect to the carbene plane for dimers cyclopropenylidene⋯ZnBr 2 (cpy-ZnBr 2 ) and imidazol-2-ylidene⋯ZnBr 2 (imi-ZnBr 2 ).
Figure 4. Dependence of the orbital energy of the three highest occupied orbitals (i.e., HOMO, HOMO-1 and HOMO-2) on the twist angle of the ZnBr 2 unit plane with respect to the carbene plane for dimers cyclopropenylidene⋯ZnBr 2 (cpy-ZnBr 2 ) and imidazol-2-ylidene⋯ZnBr 2 (imi-ZnBr 2 ).
Ijms 23 14668 g004
Figure 5. Changes in the values of the ETS-NOCV-based interaction energy components (left) or the orbital energy components (right) with the change of the rotation angle of the ZnBr 2 unit plane with respect to the carbene plane for dimers cyclopropenylidene⋯ZnBr 2 (changes relative to θ = 0 ; ZnBr 2 -cpy) and imidazol-2-ylidene⋯ZnBr 2 (changes relative to θ = 90 ; ZnBr 2 -imi).
Figure 5. Changes in the values of the ETS-NOCV-based interaction energy components (left) or the orbital energy components (right) with the change of the rotation angle of the ZnBr 2 unit plane with respect to the carbene plane for dimers cyclopropenylidene⋯ZnBr 2 (changes relative to θ = 0 ; ZnBr 2 -cpy) and imidazol-2-ylidene⋯ZnBr 2 (changes relative to θ = 90 ; ZnBr 2 -imi).
Ijms 23 14668 g005
Table 1. Some fundamental data characterizing cyclopropenylidene⋯MX 2 (X = H, Br) and imidazol-2-ylidene⋯MBr 2 dimers (M = Be, Mg, Zn): C⋯M and M-X distance, change of the MX bond length (in Å), XMX, CMX, LCL angles (in degrees), dissociation energy (in kcal/mol), charge transfer (in a.u.).
Table 1. Some fundamental data characterizing cyclopropenylidene⋯MX 2 (X = H, Br) and imidazol-2-ylidene⋯MBr 2 dimers (M = Be, Mg, Zn): C⋯M and M-X distance, change of the MX bond length (in Å), XMX, CMX, LCL angles (in degrees), dissociation energy (in kcal/mol), charge transfer (in a.u.).
MX 2 d C M d MX Δ d MX α XMX α CMX α LCL Δ α LCL D 0 CT a
cyclopropenylidene
BeH 2 1.7431.3740.039135.1112.456.60.829.1−0.317
MgH 2 2.2681.7400.035148.3105.956.70.920.9−0.273
ZnH 2 2.1211.5750.036144.6107.757.30.915.2−0.279
BeBr 2 1.7642.0430.093134.5112.857.31.535.4−0.414
MgBr 2 2.2062.3760.052145.4107.357.01.332.1−0.331
ZnBr 2 2.0612.2980.083137.1111.557.31.528.2−0.380
imidazol-2-ylidene
BeBr 2 1.7652.0590.109132.9113.6104.23.548.6−0.464
MgBr 2 2.1732.393 b0.069 c147.9106.1 d103.52.843.7−0.363
ZnBr 2 2.0372.3110.097139.2110.4104.23.441.2−0.432
a The charge transfer was calculated based on Hirshfeld’s atomic charges [126,127]. b The average value determined from the two values 2.386 Å and 2.400 Å. c The average value determined from the two values 0.062 Å and 0.076 Å. d The average value determined from the two values 102.9 and 109.2 .
Table 2. The IQA-based energy (in a.u.) terms (see Methods) computed for the C⋯M interaction in the carbene⋯MX 2 dimers.
Table 2. The IQA-based energy (in a.u.) terms (see Methods) computed for the C⋯M interaction in the carbene⋯MX 2 dimers.
MX 2 E neen E nn E ee E ee , C E ee , xc % E ee , xc E elst E int
cyclopropenylidene
BeH 2 −12.28397.28464.65054.6963−0.045813.1−0.3030−0.3488
MgH 2 −32.499916.801715.509815.5405−0.030716.3−0.1577−0.1884
ZnH 2 −90.127744.898745.063445.1726−0.109165.9−0.0564−0.1655
BeBr 2 −12.10177.19904.53834.5810−0.042811.7−0.3217−0.3644
MgBr 2 −33.393417.271315.904015.9383−0.034315.7−0.1838−0.2181
ZnBr 2 −92.581646.221746.168246.2894−0.121263.2−0.0705−0.1917
imidazol-2-ylidene
BeBr 2 −11.36097.19764.10934.1554−0.046185.5−0.0079−0.0539
MgBr 2 −32.034117.535114.537014.5767−0.0397−104.70.07770.0379
ZnBr 2 −88.308646.754141.541841.6735−0.13171037.00.1190−0.0127
Table 3. Sums of the IQA-based energy terms (in kcal/mol) computed (MP2/6-311+G(d,p)) for groups of atoms reflecting the ETS-NOCV fragmentation, that is {carbene}⋯{MX 2 } .
Table 3. Sums of the IQA-based energy terms (in kcal/mol) computed (MP2/6-311+G(d,p)) for groups of atoms reflecting the ETS-NOCV fragmentation, that is {carbene}⋯{MX 2 } .
MX 2 E int E elst % E elst E ee , xc % E ee , xc
cyclopropenylidene
BeH 2 −143.28−73.6851.4−69.6048.6
MgH 2 −71.81−38.5553.7−33.2546.3
ZnH 2 −91.45−14.8616.2−76.5983.8
BeBr 2 −162.74−90.9855.9−71.7644.1
MgBr 2 −88.25−49.4156.0−38.8544.0
ZnBr 2 −105.00−18.6117.7−86.3882.3
imidazol-2-ylidene
BeBr 2 −190.24−110.0057.8−80.2742.2
MgBr 2 −113.04−63.8756.5−49.1743.5
ZnBr 2 −127.25−26.1920.6−101.0579.4
Table 4. The ETS-NOCV-based energy (in kcal/mol) contributions to carbene⋯MX 2 bonding.
Table 4. The ETS-NOCV-based energy (in kcal/mol) contributions to carbene⋯MX 2 bonding.
MX 2 Δ E int Δ E elst Δ E Pauli Δ E disp Δ E orb Δ E orb σ Δ E orb π Δ E orb NH Br
cyclopropenylidene
BeH 2 −45.69−69.29 (56%)77.03−2.56 (2%)−50.87 (41%)−40.33−10.11n/a
MgH 2 −26.20−45.85 (70%)39.55−2.58 (4%)−17.32 (26%)−14.16−2.88n/a
ZnH 2 −23.63−75.38 (68%)87.86−2.87 (3%)−33.24 (30%)−27.31−5.19n/a
BeBr 2 −52.23−82.67 (58%)90.98−4.90 (3%)−55.64 (39%)−47.43−6.34n/a
MgBr 2 −38.57−57.23 (69%)44.25−4.14 (5%)−21.44 (26%)−17.67−2.56n/a
ZnBr 2 −37.74−95.45 (67%)104.58−4.52 (3%)−42.34 (30%)−36.96−3.53n/a
imidazol-2-ylidene
BeBr 2 −66.86−103.25 (61%)102.10−6.61 (4%)−59.10 (35%)−47.92−5.24−1.71
MgBr 2 −50.49−77.00 (71%)58.24−6.01 (5%)−25.72 (24%)−18.57−3.13−1.10
ZnBr 2 −50.92−122.82 (68%)128.22−6.45 (4%)−49.87 (28%)−42.15−3.70−1.11
Table 5. DPLNO-CCSD(T)/ccpv-tz Local Energy Decomposition analysis of the {carbene}⋯{MX 2 } bonding (carbene names shortened for clarity: cpy = cyclopropenylidene and imi = imidazol-2-ylidene). Energies in kcal/mol.
Table 5. DPLNO-CCSD(T)/ccpv-tz Local Energy Decomposition analysis of the {carbene}⋯{MX 2 } bonding (carbene names shortened for clarity: cpy = cyclopropenylidene and imi = imidazol-2-ylidene). Energies in kcal/mol.
cpy–BeBr 2 cpy–MgBr 2 cpy–ZnBr 2 imi–BeBr 2 imi–MgBr 2 imi–ZnBr 2
E int −39.60−34.28−29.08−54.28−46.37−42.63
E elst ref −218.44−152.51−327.12−251.84−190.31−393.52
E exch ref −20.12−11.84−33.73−20.8414.59−39.32
E geo prep ref 14.886.2310.7617.586.7311.24
E el prep ref 190.33126.75325.62210.63158.28387.67
E non disp C -0.690.121.91−2.74−1.83−0.17
E disp C −5.35−2.93−6.14−6.43−4.10−7.67
E T C −0.21−0.10−0.38−0.63−0.55−0.85
CT 1→2−16.96−9.52−13.44−19.17−11.11−16.32
CT 2→1−3.33−2.43−7.58−3.56−2.53−8.14
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sagan, F.; Mitoraj, M.; Jabłoński, M. Nature of Beryllium, Magnesium, and Zinc Bonds in Carbene⋯MX2 (M = Be, Mg, Zn; X = H, Br) Dimers Revealed by the IQA, ETS-NOCV and LED Methods. Int. J. Mol. Sci. 2022, 23, 14668. https://doi.org/10.3390/ijms232314668

AMA Style

Sagan F, Mitoraj M, Jabłoński M. Nature of Beryllium, Magnesium, and Zinc Bonds in Carbene⋯MX2 (M = Be, Mg, Zn; X = H, Br) Dimers Revealed by the IQA, ETS-NOCV and LED Methods. International Journal of Molecular Sciences. 2022; 23(23):14668. https://doi.org/10.3390/ijms232314668

Chicago/Turabian Style

Sagan, Filip, Mariusz Mitoraj, and Mirosław Jabłoński. 2022. "Nature of Beryllium, Magnesium, and Zinc Bonds in Carbene⋯MX2 (M = Be, Mg, Zn; X = H, Br) Dimers Revealed by the IQA, ETS-NOCV and LED Methods" International Journal of Molecular Sciences 23, no. 23: 14668. https://doi.org/10.3390/ijms232314668

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop