Mahmoud A. A. Ibrahim* and
Ebtisam M. Z. Telb
Computational Chemistry Laboratory, Chemistry Department, Faculty of Science, Minia University, Minia 61519, Egypt. E-mail: m.ibrahim@compchem.net
First published on 19th January 2021
For the first time, unconventional ±R˙-hole interactions were unveiled in tetrel-containing complexes. The nature and characteristics of ±R˙-hole interactions were explored relative to their ±σ-hole counterparts for ˙TF3⋯ and W–T–F3⋯B/R˙/A complexes (where T = C, Si, and Ge, W = H and F, B = Lewis bases, R˙ = free radicals, and A = Lewis acids). In an effort to thoroughly investigate such interactions, a plethora of quantum mechanical calculations, including molecular electrostatic potential (MEP), maximum positive electrostatic potential (Vs,max), point-of-charge (PoC), interaction energy, symmetry adapted perturbation theory (SAPT), and reduced density gradient–noncovalent interaction (RDG–NCI) calculations, were applied. The most notable findings to emerge from this study are that (i) from the electrostatic perspective, the molecular stabilization energies of ˙TF3 and W–T–F3 monomers became more negative as the Lewis basicity increased, (ii) the most stable complexes were observed for the ones containing Lewis bases, forming −σ-hole and −R˙-hole interactions, and the interaction energies systematically increased in the order H–T–F3⋯B < ˙TF3⋯B < F–T–F3⋯B, (iii) contrariwise, the +σ-hole and +R˙-hole interactions with Lewis acids are more energetically favorable in the order F–T–F3⋯A < ˙TF3⋯A < H–T–F3⋯A, and (iv) generally, the dispersion force plays a key role in stabilizing the tetrel-containing complexes, jointly with the electrostatic and induction forces for the interactions with Lewis bases and acids, respectively. Concretely, the findings presented in this paper add to our understanding of the characteristics and nature of such intriguing interactions.
Among the σ-hole interactions, tetrel bonding has received immense interest due to its technological and fundamental magnitude in supramolecular chemistry25 and dynamical processes such as protein folding and ligand–acceptor interactions.26,27 Several studies have documented the potentiality of the tetrel-containing molecules to form −σ-hole interactions as a consequence of the interaction of Group IV elements (T) via the σ-hole with Lewis bases (B) such as lone-pair (lp),28–31 anion,32,33 π-systems.34,35 Recently, a thought-provoking study has given prominence to +σ-hole interactions of tetrel-containing complexes, which demonstrated the ability of tetrel-containing molecules to preferentially interact with Lewis acids (A).30 Nonetheless, only a handful of reports in the literature were introduced to investigate the intermolecular interactions of tetrel-containing molecules with Lewis acids.
In light of recent research, single-electron noncovalent interactions were reported, referring to an interaction between the unpaired electron of a radical (R˙) acting as a Lewis base and the σ-hole on the Group IV–VIII elements acting as a Lewis acid.36–44 Considering the pivotal role of radicals in combustion, polymerization, biochemistry, and many other chemical processes,45–48 further studies are needed to characterize the interactions of radical species with tetrel-containing molecules.
One of the most significant current investigations in noncovalent interactions is that ascribed to the lp-hole. Lp-hole interactions were accentuated for Group V–VIII elements as a result of a positive region opposite to the lp (called lp-hole).49–51 According to the term lp-hole, it is not plausible for sp3-hybridized tetrel-containing molecules to form this type of interaction. A question has been raised about whether sp3-hybridized tetrel radicals (˙TX3) can form another intriguing interaction so-called R˙-hole interaction. This interaction is attributed to the presence of a positive region opposite to the unpaired electron on the tetrel (termed, R˙-hole). As far as our knowledge extends, no previous study has assessed the occurrence of R˙-hole interaction.
Subsequently, the objectives of the current work were to (i) uncover the occurrence of the R˙-hole interactions and (ii) gain insight into the characteristics and nature of the σ-hole and R˙-hole interactions of tetrel-containing molecules with Lewis bases, free radicals, and Lewis acids. −R˙-hole and +R˙-hole interactions were assigned for the interactions with Lewis bases and acids, by analogy with −σ-hole and +σ-hole interactions, respectively. The point-of-charge (PoC) approach was utilized to precisely determine the R˙-hole location and elucidate the ±σ-hole and ±R˙-hole interactions from the electrostatic perspective.30,49,52–56 The interaction energies for ˙TF3⋯ and W–T–F3⋯B/R˙/A complexes (where T = C, Si, and Ge, and W = H and F) were calculated at the MP2/aug-cc-pVTZ(PP) and CCSD(T)/CBS levels of theory. The symmetry adapted perturbation theory (SAPT) and reduced density gradient–noncovalent interaction (RDG–NCI) index calculations were executed to reconnoiter the characteristics and nature of such interactions. This study makes a major contribution to research on tetrel radicals-bonded complexes by demonstrating the characteristics and nature of R˙-hole interactions compared to the σ-hole ones for the first time.
To fulfill the aim of this investigation, the R˙-hole location was first precisely determined. The studied ˙TF3 radicals were aligned to the x-axis, and the yz plane was then scanned by −0.01 au PoC with a step size of 0.1 Å, using the point-of-charge (PoC) approach (Fig. 1(ia)). The PoC was placed at a distance of 2.0 Å from the tetrel atom and moved along both y- and z-directions in a range from 1.6 to −1.6 Å, generating 2D-molecular stabilization energy surfaces (see Fig. 1(ia)). To inspect the effect of Lewis basicity and acidity on the ±σ-hole and ±R˙-hole interactions strength, different values of negatively and positively charged points stimulating Lewis bases and acids, respectively, were applied. In the latter calculation, the studied monomers were aligned to the x-axis, and the T⋯PoC distance effect on the strength of interaction was evaluated in the range 2.5 to 5.0 Å (see Fig. 1(ib)). The employed PoC values were set to ±0.10, ±0.25, ±0.50, and ±1.00 au. The strength of the ±σ-hole and ±R˙-hole interactions was estimated at MP2/aug-cc-pVTZ(PP) level of theory in terms of the molecular stabilization energy (Establization) according to the following equation:30,49,50,52–56
Establization = Etetrel-containing molecule⋯PoC − Etetrel-containing molecule |
Fig. 1 Schematic representation of (i) implemented PoC-based calculations and (ii) ˙TF3⋯ and W–T–F3⋯B/R˙/A complexes. |
To analyze the potentiality of ˙TF3 and W–T–F3 systems to favorably interact via the σ-hole and R˙-hole with B, R˙, and A (see Fig. 1(ii)), full geometrical optimization of the complexes in C3V space group was performed at the MP2/aug-cc-pVTZ level of theory (with PP functions for Ge atom). Vibrational frequency calculations were not performed for the binary complexes; thus, there was a possibility that the structures were not energetic minima. The interaction energies of the binary complexes were then estimated at the same level of optimization and performed with taking basis set superposition error (BSSE) correction into account.62 The interaction energies were computed as the difference in energy between the complex and the sum of monomers with the same geometries as in the complex. The frozen-core (FC) approximation was adopted for all MP2 calculations. Furthermore, the interaction energies were calculated for the optimized complexes at the CCSD(T)/CBS level of theory. The CCSD(T)/CBS energies were calculated based on the following equation:63
ECCSD(T)/CBS = ΔEMP2/CBS + ΔECCSD(T) |
ΔEMP2/CBS = (64EMP2/aug-cc-pVQZ − 27EMP2/aug-cc-pVTZ)/37 |
ΔECCSD(T) = ECCSD(T)/aug-cc-pVdZ − EMP2/aug-cc-pVDZ |
To have a more in-depth insight into the nature of σ-hole and R˙-hole interactions, the energy decomposition calculations were performed through the symmetry adapted perturbation theory-based energy decomposition analysis (SAPT-EDA) method using PSI4 code.64,65 The SAPT-EDA method provides a separation of interaction energies into four physically meaningful components, such as those arising from electrostatics (Eelst), exchange (Eexch), induction (Eind), and dispersion (Edisp). The interaction energies were estimated at SAPT0 truncation along with an aug-cc-pVTZ basis set, given as follows:66–68
ESAPT0int = Eelst + Eexch + Eind + Edisp |
Eelst = Eelst(10) |
Eexch = Eexch(10) |
Eind = Eind,resp(20) + Eexch-ind,resp(20) + δEHF(2) |
Edisp = Edisp(20) + Eexch-disp(20) |
To analyze the bonding characteristics of σ-hole and R˙-hole interactions, reduced density gradient–noncovalent interaction (RDG–NCI) indices were constructed, and the NCI plots were also illustrated.69 The gradient isosurfaces were colored on a blue-green-red (BGR) scale, where blue surfaces refer to strong attractions, green surfaces show weak interactions, and red surfaces imply strong repulsions. The coloring scale of electron density (ρ) was from −0.035 (blue) to 0.020 (red) au. The RDG–NCI analysis was carried out using Multiwfn3.7 software and then visualized by VMD1.9.2 software.70 The geometrical optimization, MEP analysis, interaction energy, and PoC-based calculations were performed using Gaussian09 software.71
According to the data presented in Fig. 2, the studied tetrel-containing molecules exhibited variable-in-size σ-holes along the W–T bond extensions (where T = C, Si, and Ge, and W = H and F). The σ-hole size was increased with moving from lighter to heavier tetrel atoms in order C < Si < Ge. For example, the Vs,max values at the σ-hole were found to be 31.6, 64.1, and 75.4 kcal mol−1 for F–C–F3, F–Si–F3, and F–Ge–F3 molecules, respectively. As well, a direct correlation between the electronegativity of the W atom and σ-hole size of the tetrel atom was observed. Taking W–Ge–F3 system as an example, the obtained Vs,max values at the σ-hole for H–Ge–F3, and F–Ge–F3 were 48.5 and 75.4 kcal mol−1, respectively.
Interestingly, a positive region opposite to the unpaired electron (i.e., R˙-hole) was noticed in the studied radicals. The Vs,max magnitude at the R˙-hole attenuated in the order ˙GeF3 > ˙SiF3 > ˙CF3 and obtained with values of 52.8, 51.2, and 30.9 kcal mol−1, respectively. Generally, the Vs,max values at the σ-hole were higher than those at R˙-hole for the same monomer.
As shown in Fig. 3, the largest molecular stabilization energies (i.e., more negative) were observed at the centroid of the F3 atoms with values of −0.15, −0.30, and −0.31 kcal mol−1 for ˙CF3, ˙SiF3, and ˙GeF3 radicals, respectively. These findings are in line with those for lp-hole containing molecules, where the precise lp-hole positions were identified at the centroid of the xyz plane in the covalently bonded Group V-VIII elements.49
Molecule | Molecular stabilization energies (kcal mol−1) at 2.5 Å | |||
---|---|---|---|---|
PoC = −0.10 | PoC = −0.25 | PoC = −0.50 | PoC = −1.00 | |
˙CF3 | −0.65 | −1.96 | −5.06 | −14.50 |
H–C–F3 | 0.22 | 0.14 | −1.01 | −6.94 |
F–C–F3 | −0.72 | −2.18 | −5.58 | −15.82 |
˙SiF3 | −1.19 | −3.47 | −8.51 | −22.96 |
H–Si–F3 | −1.10 | −3.24 | −8.03 | −21.91 |
F–Si–F3 | −1.84 | −5.06 | −11.57 | −28.62 |
˙GeF3 | −0.20 | −3.01 | −7.73 | −21.92 |
H–Ge–F3 | −0.76 | −2.44 | −6.61 | −19.71 |
F–Ge–F3 | −2.21 | −6.00 | −13.55 | −32.91 |
Molecule | Molecular stabilization energies (kcal mol−1) at 2.5 Å | |||
---|---|---|---|---|
PoC = +0.10 | PoC = +0.25 | PoC = +0.50 | PoC = +1.00 | |
˙CF3 | 0.45 | 0.76 | 0.20 | −5.75 |
H–C–F3 | −0.44 | −1.54 | −4.64 | −16.81 |
F–C–F3 | 0.51 | 0.87 | 0.30 | −6.07 |
˙SiF3 | 0.92 | 1.77 | 1.62 | −1.19 |
H–Si–F3 | 0.83 | 1.53 | 1.12 | −7.63 |
F–Si–F3 | 1.59 | 3.46 | 5.12 | 1.23 |
˙GeF3 | 0.69 | 1.14 | 0.17 | −10.22 |
H–Ge–F3 | 0.46 | 0.55 | −1.04 | −12.88 |
F–Ge–F3 | 1.93 | 4.30 | 6.67 | 3.74 |
Concerning −σ-hole and −R˙-hole interactions, the molecular stabilization energies were boosted by increasing the negativity of PoCs (i.e., Lewis basicity) and decreasing the T⋯PoC distances. For example, the observed molecular stabilization energies for ˙CF3 monomer were −0.65, −1.96, −5.06, and −14.5 kcal mol−1 at T⋯PoC distance of 2.5 Å in the presence of −0.10, −0.25, −0.50, and −1.00 au PoCs, respectively (see Table 1). Furthermore, the effect of Lewis basicity was most distinguishable for the F–T–F3 followed in order by the ˙TF3 and H–T–F3 systems. For instance, the obtained molecular stabilization energies at T⋯PoC distance of 2.5 Å for F–Ge–F3, ˙GeF3, and H–Ge–F3 monomers were −32.91, −21.92, and −19.71 kcal mol−1 in case of −1.00 au PoC, respectively. The observation can explain this pattern that the systematic growth of the Vs,max magnitude at σ-hole, and R˙-hole led to a gradual increase in the molecular stabilization energies (i.e., became more negative). More interestingly, according to Fig. 4, molecular stabilization energies remained to be observed at long T⋯PoC distances for ˙CF3 and F–C–F3 molecules, demonstrating the dominance of the attractive electrostatic force between the negative PoC and the positive electrophilic sites over the tetrel atom (i.e., σ-hole and R˙-hole). Otherwise, in the case of H–C–F3 molecule, molecular destabilization energies were noticed and augmented by increasing the Lewis basicity, indicating the repulsive electrostatic interaction with F3 atoms is the controlling one at long T⋯PoC distance.
For +σ-hole and +R˙-hole interactions, molecular stabilization energies were observed for all investigated molecules, except SiF4 and GeF4. The incompetence of SiF4 and GeF4 to participate in +σ-hole interaction is appertaining to the significant repulsive interaction between the positive PoC and the excessive positive σ-hole. For H–C–F3 monomer as an exemplar, H–C–F3 demonstrated molecular stabilization energies of −0.44, −1.54, −4.64, and −16.81 kcal mol−1 at T⋯PoC distance of 2.5 Å in the presence of +0.10, +0.25, +0.50, and +1.00 au PoCs, respectively. As claimed by the latter observation, the molecular stabilization energies boosted with increasing the positivity of the PoC (i.e., Lewis acidity) despite the repulsive electrostatic interaction between the positive σ-hole and positive PoC, illustrating the effectual role of polarization in stabilizing such interactions. For some tetrel-containing molecules, generally, unfavorable electrostatic interactions were noticed at very short T⋯PoC distances. It is apparent from Table 1 that the H–T–F3 system gave more favorable electrostatic interaction compared to ˙TF3 and F–T–F3 systems. For instance, the molecular stabilization energies were found to increase in the order of F–C–F3 < ˙CF3 < H–C–F3 with values of −5.75, −6.07, and −16.81 kcal mol−1 at 2.5 Å in the presence of +1.00 au PoC, respectively.
Overall, with respect to the research question, it was found that the studied monomers can electrostatically interact with Lewis bases and acids along the R˙-hole extension, forming ±R˙-hole interactions. Compared with the −σ-hole and –R˙-hole interactions, the parallel +σ-hole and +R˙-hole interactions with Lewis acids are less favorable. The substantial role of polarization effect and F3 atoms in forming such interactions cannot be neglected.
Complex | ˙TF3/W–T–F3⋯B systems | |||||
---|---|---|---|---|---|---|
NCH | FH | |||||
EMP2 | ECCSD(T) | dT⋯NCH | EMP2 | ECCSD(T) | dT⋯FH | |
a Optimized structure could not be obtained. | ||||||
˙CF3⋯B | −1.04 | −1.14 | 3.29 | −0.58 | −0.75 | 3.18 |
H–C–F3⋯B | 0.01 | −0.04 | 3.55 | 0.08 | −0.03 | 3.39 |
F–C–F3⋯B | −1.09 | −1.15 | 3.36 | −0.62 | −0.79 | 3.23 |
˙SiF3⋯B | −2.08 | −2.28 | 3.17 | −0.77 | −1.03 | 3.23 |
H–Si–F3⋯B | −1.95 | −2.10 | 3.18 | −0.70 | −0.95 | 3.22 |
F–Si–F3⋯B | −3.44 | −3.64 | 3.01 | −1.37 | −1.67 | 3.10 |
˙GeF3⋯B | −2.70 | −2.95 | 2.96 | −0.56 | −0.89 | 3.14 |
H–Ge–F3⋯B | −2.13 | −2.32 | 3.01 | −0.39 | −0.67 | 3.21 |
F–Ge–F3⋯B | −10.21 | −17.83 | 2.39 | −1.84 | −2.29 | 2.95 |
Complex | ˙TF3/W–T–F3⋯R˙ systems | |||||
---|---|---|---|---|---|---|
˙CH3 | ˙CF3 | |||||
EMP2 | ECCSD(T) | dT···˙CH3 | EMP2 | ECCSD(T) | dT···˙CF3 | |
˙CF3⋯R˙ | −0.41 | −0.57 | 3.74 | −0.27 | −0.47 | 3.55 |
H–C–F3⋯R˙ | −0.26 | −0.43 | 3.86 | −0.56 | −0.78 | 3.52 |
F–C–F3⋯R˙ | −0.46 | −0.63 | 3.75 | −0.34 | −0.50 | 3.53 |
˙SiF3⋯R˙ | −0.69 | −0.98 | 3.65 | −0.37 | −0.65 | 3.58 |
H–Si–F3···R˙ | −0.72 | −1.02 | 3.61 | −0.42 | −0.73 | 3.52 |
F–Si–F3···R˙ | −1.03 | −1.40 | 3.47 | −0.25 | −0.54 | 3.47 |
˙GeF3⋯R˙ | −0.74 | −1.19 | 3.47 | −0.35 | −0.85 | 3.36 |
H–Ge–F3⋯R˙ | −0.70 | −1.11 | 3.52 | −0.51 | −0.97 | 3.39 |
F–Ge–F3⋯R˙ | −1.79 | −2.42 | 3.17 | −0.19 | −0.64 | 3.25 |
Complex | ˙TF3/W–T–F3⋯A systems | |||||
---|---|---|---|---|---|---|
HCN | HF | |||||
EMP2 | ECCSD(T) | dT⋯HCN | EMP2 | ECCSD(T) | dT⋯HF | |
˙CF3⋯A | 0.07 | 0.02 | 2.96 | 0.11 | 0.09 | 3.49 |
H–C–F3⋯A | −1.07 | −1.14 | 2.88 | −0.79 | −0.88 | 2.91 |
F–C–F3⋯A | 0.00 | −0.03 | 2.99 | 0.15 | 0.14 | 3.26 |
˙SiF3⋯A | −0.14 | −0.16 | 3.41 | 0.07 | 0.04 | 3.97 |
H–Si–F3⋯A | −0.18 | −0.22 | 3.53 | −0.08 | −0.09 | 4.62 |
F–Si–F3⋯A | 0.49 | 0.48 | 3.55 | —a | —a | —a |
˙GeF3⋯A | −0.34 | −0.46 | 3.14 | −0.01 | −0.07 | 3.46 |
H–Ge–F3⋯A | −0.58 | −0.72 | 3.10 | −0.28 | −0.34 | 3.57 |
F–Ge–F3⋯A | 0.44 | 0.43 | 3.86 | —a | —a | —a |
As can be readily appreciated from Table 2, the investigated monomers exhibited a strong tendency to interact with Lewis bases, Lewis acids, and free radicals with CCSD(T)/CBS interaction energies in the range −0.03 to −17.83, 0.48 to −0.72, and −0.43 to −2.42 kcal mol−1, respectively. For the analyzed complexes, the interaction energies were almost greater (i.e., more negative) for the Ge-containing complexes compared to the Si and C counterparts, which are in agreement with the MEP results (see Fig. 2). Noteworthy, reasonable differences between the interaction energies estimated at MP2/aug-cc-pVTZ(PP) and those calculated at CCSD(T)/CBS were observed, demonstrating the accuracy and effectiveness of the MP2/aug-cc-pVTZ(PP) level of theory to investigate such noncovlanet interactions. This is in agreement with previous studies on estimating the strength of noncovlent interactions in related systems28,32,73–76 (Table 2).
In regards to ˙TF3⋯ and W–T–F3⋯B complexes, −σ-hole, and −R˙-hole interaction energies were large negative values, indicating strong interactions between the interacting subunits. A linear correlation was found between the interaction energies and the Vs,max at the σ-hole and R˙-hole over the tetrel atom, where the interaction energies enhanced in order H–T–F3⋯B < ˙TF3⋯B < F–T–F3⋯B. For example, the CCSD(T)/CBS interaction energies for H–Ge–F3⋯, ˙GeF3⋯, and F–Ge–F3⋯NCH complexes were noticed with values of −2.32, −2.95, and −17.83 kcal mol−1, respectively. It was also found that the greater the atomic size of tetrel atom, the shorter T⋯B contact, which in turn resulted in strengthening interaction energies. The example given was for F–C–F3⋯, F–Si–F3⋯, and F–Ge–F3⋯NCH complexes, where the CCSD(T)/CBS interaction energies were noticed with values of −1.15, −3.64, and −17.83 kcal mol−1 at T⋯N intermolecular distances values of 3.36, 3.01, and 2.39 Å, respectively (see Table 2). Moreover, the NCH Lewis base was found to interact more strongly than the FH candidate with the studied monomers.
For ˙TF3⋯ and W–T–F3⋯A complexes, favorable +σ-hole and +R˙-hole interactions were observed for most of the studied complexes. The inability of some monomers to engage in favorable interactions with Lewis acids can be attributed to the dominance of the repulsive electrostatic interaction between the electrophilic site over the tetrel atom and the Lewis acid. Generally, the interaction energies reinforced by decreasing the Vs,max value over the tetrel atom in the order F–T–F3⋯A < ˙TF3⋯A < H–T–F3⋯A. As an exemplar, the CCSD(T)/CBS interaction energies were found with values of 0.43, −0.46, and −0.72 kcal mol−1 for F–Ge–F3⋯, ˙GeF3⋯, and H–Ge–F3⋯HCN complexes, respectively. However, the CCSD(T)/CBS interaction energies for ˙CF3⋯, ˙SiF3⋯ and ˙GeF3⋯HCN complexes were with values of 0.02, −0.16, and −0.46 kcal mol−1, respectively, despite increasing the Vs,max at R˙-hole in order of C < Si < Ge. These findings explicitly referred to the interaction of F3 atoms with the H atom of the Lewis acid and were further investigated by NCI analysis. Compared to HF-containing complexes, the interaction energies for HCN-containing complexes were systematically larger (i.e., more negative). For example, the CCSD(T)/CBS interaction energies for H–Ge–F3⋯HCN and ⋯HF complexes were −0.72 and −0.34 kcal mol−1, respectively.
In the case of ˙TF3⋯ and W–T–F3⋯R˙ complexes, which exhibited a staggered conformation with C3v symmetry, the ˙CH3, and ˙CF3 can act as a nucleophile and an electrophile, respectively, according to the generated MEP maps (see Fig. S3†). For ˙TF3⋯ and W–T–F3⋯˙CH3 complexes, the interaction energies increased in the same order as in Lewis base-containing complexes, confirming the potentiality of ˙CH3 to act as a Lewis base. For instance, the CCSD(T)/CBS interaction energies amounted to −1.11, −1.19, and −2.42 kcal mol−1 for H–Ge–F3⋯, ˙˙GeF3⋯, and F–Ge–F3⋯˙CH3 complexes, respectively. On the other hand, for ˙CF3-containing complexes, the interaction became stronger in the order F–T–F3⋯˙CF3 < ˙TF3⋯˙CF3 < H–T–F3⋯˙CF3, demonstrating that ˙CF3 could act as a Lewis acid. As an example, the CCSD(T)/CBS interaction energies were observed with values of −0.64, −0.85, and −0.97 kcal mol−1 for F–Ge–F3⋯, ˙GeF3⋯, and H–Ge–F3⋯˙CF3 complexes, respectively (see Table 2).
Overall, these results divulge that the strength of ±σ-hole interactions is higher compared to the ±R˙-hole ones, and the most stable complexes were observed for the ones containing Lewis bases. Besides, the strength of ±σ-hole and ±R˙-hole interactions in ˙TF3⋯ and W–T–F3⋯B/R˙/A complexes is not exclusive to the repulsive/attractive electrostatic interaction between the electrophilic site on tetrel atom and B/R˙/A, but also the interactions with the F3 atoms. These appealing findings seem to be consistent with other previous reports concerning the interactions of tetrel-containing molecules with Lewis bases, acids, and free radicals.28,30,36,74,77,78
Complex | ˙TF3/W–T–F3⋯B systems | |||||||||
---|---|---|---|---|---|---|---|---|---|---|
NCH | FH | |||||||||
Eelst | Eexch | Eind | Edisp | ESAPT0int | Eelst | Eexch | Eind | Edisp | ESAPT0int | |
a Optimized structure could not be obtained, so no SAPT calculations were performed. | ||||||||||
˙CF3⋯B | −1.26 | 1.77 | −0.22 | −1.63 | −1.34 | −0.54 | 0.81 | −0.11 | −0.98 | −0.82 |
H–C–F3⋯B | 0.22 | 0.92 | −0.14 | −1.19 | −0.19 | 0.30 | 0.48 | −0.08 | −0.76 | −0.04 |
F–C–F3⋯B | −1.37 | 1.58 | −0.21 | −1.57 | −1.55 | −0.67 | 0.77 | −0.11 | −0.97 | −0.98 |
˙SiF3⋯B | −3.50 | 3.76 | −0.67 | −2.42 | −2.83 | −1.04 | 1.10 | −0.21 | −1.11 | −1.26 |
H–Si–F3⋯B | −3.35 | 3.65 | −0.64 | −2.39 | −2.72 | −1.01 | 1.12 | −0.21 | −1.13 | −1.23 |
F–Si–F3⋯B | −6.13 | 5.66 | −1.24 | −2.97 | −4.68 | −2.04 | 1.51 | −0.35 | −1.29 | −2.17 |
˙GeF3⋯B | −6.12 | 7.20 | −1.56 | −3.63 | −4.11 | −1.22 | 1.71 | −0.32 | −1.44 | −1.28 |
H–Ge–F3⋯B | −5.16 | 6.42 | −1.26 | −3.42 | −3.41 | −0.84 | 1.37 | −0.25 | −1.30 | −1.02 |
F–Ge–F3⋯B | −24.0 | 28.45 | −12.1 | −7.75 | −15.4 | −3.28 | 2.57 | −0.72 | −1.76 | −3.19 |
Complex | ˙TF3/W–T–F3⋯R˙ systems | |||||||||
---|---|---|---|---|---|---|---|---|---|---|
˙CH3 | ˙CF3 | |||||||||
Eelst | Eexch | Eind | Edisp | ESAPT0int | Eelst | Eexch | Eind | Edisp | ESAPT0int | |
˙CF3⋯R˙ | −0.66 | 1.12 | −0.06 | −1.01 | −0.61 | −0.27 | 1.07 | −0.08 | −1.17 | −0.45 |
H–C–F3⋯R˙ | −0.36 | 0.93 | −0.06 | −0.95 | −0.44 | −0.86 | 1.46 | −0.12 | −1.41 | −0.94 |
F–C–F3⋯R˙ | −0.72 | 1.18 | −0.06 | −1.05 | −0.66 | −0.25 | 1.25 | −0.10 | −1.30 | −0.41 |
˙SiF3⋯R˙ | −1.43 | 2.07 | −0.18 | −1.46 | −1.01 | −0.33 | 1.48 | −0.17 | −1.46 | −0.46 |
H–Si–F3⋯R˙ | −1.53 | 2.25 | −0.20 | −1.55 | −1.03 | −0.42 | 1.72 | −0.20 | −1.58 | −0.48 |
F–Si–F3⋯R˙ | −2.33 | 3.11 | −0.39 | −1.82 | −1.43 | 0.03 | 1.91 | −0.29 | −1.65 | −0.06 |
˙GeF3⋯R˙ | −2.31 | 3.65 | −0.40 | −2.20 | −1.26 | −1.01 | 3.21 | −0.42 | −2.38 | −0.60 |
H–Ge–F3⋯R˙ | −2.02 | 3.26 | −0.32 | −2.07 | −1.15 | −1.05 | 2.96 | −0.36 | −2.28 | −0.73 |
F–Ge–F3⋯R˙ | −4.68 | 6.60 | −1.47 | −3.01 | −2.57 | −0.20 | 3.56 | −0.75 | −2.46 | 0.14 |
Complex | ˙TF3/W–T–F3⋯A systems | |||||||||
---|---|---|---|---|---|---|---|---|---|---|
HCN | HF | |||||||||
Eelst | Eexch | Eind | Edisp | ESAPT0int | Eelst | Eexch | Eind | Edisp | ESAPT0int | |
˙CF3⋯A | 0.65 | 0.45 | −0.24 | −0.81 | 0.05 | 0.40 | 0.03 | −0.09 | −0.21 | 0.13 |
H–C–F3⋯A | −0.60 | 0.81 | −0.35 | −1.06 | −1.21 | −0.12 | 0.39 | −0.33 | −0.66 | −0.72 |
F–C–F3⋯A | 0.82 | 0.44 | −0.25 | −0.83 | 0.19 | 0.75 | 0.08 | −0.15 | −0.34 | 0.35 |
˙SiF3⋯A | 0.36 | 0.10 | −0.14 | −0.44 | −0.11 | 0.29 | 0.01 | −0.06 | −0.14 | 0.10 |
H–Si–F3⋯A | 0.43 | 0.14 | −0.16 | −0.51 | −0.10 | 0.02 | 0.01 | −0.03 | −0.06 | −0.06 |
F–Si–F3⋯A | 1.18 | 0.07 | −0.13 | −0.39 | 0.74 | —a | —a | —a | —a | —a |
˙GeF3⋯A | 0.41 | 0.49 | −0.30 | −0.91 | −0.31 | 0.53 | 0.08 | −0.16 | −0.34 | 0.10 |
H–Ge–F3⋯A | 0.23 | 0.56 | −0.33 | −0.97 | −0.51 | 0.21 | 0.05 | −0.13 | −0.29 | −0.16 |
F–Ge–F3⋯A | 1.03 | 0.03 | −0.08 | −0.27 | 0.67 | —a | —a | —a | —a | —a |
A perusal of the data in Table 3 shows that for −σ-hole and −R˙-hole interactions, the Edisp, Eelst, and Eind terms are negative, unlikely; the Eexch is positive. As well, all these terms increased in the order C < Si < Ge. Generally, it is apparent that the Edisp governs the −σ-hole and −R˙-hole interactions, followed by the Eelst, and Eind. For instance, Edisp contributions were 89, 90, and 69% out of the total attractive terms for the H–C–F3⋯NCH, ⋯FH, and ⋯˙CH3 complexes, respectively. On the other hand, for Ge-containing molecules interacting with NCH as Lewis base, the interactions were dominated by Eelst; where the Eelst contribution to the total interaction energy ranged from −5.16 to −24.0 kcal mol−1 and the Edisp contribution ranged from −3.42 to −7.75 kcal mol−1 (see Table 3). These findings demonstrated the importance of electrostatic and dispersion energies in stabilizing ˙TF3⋯ and W–T–F3⋯NCH/FH/˙CH3 complexes.
Turning now to +σ-hole and +R˙-hole interactions, the Edisp term was the most important attractive contributor. Besides, generally, the Eelst and Eind terms help in promoting the strength of interactions for ˙TF3/W–T–F3⋯˙CF3 and ⋯HCN/HF complexes, respectively. For example, the Edisp, Eelst, and Eind were −2.38, −1.01, and −0.42 kcal mol−1 for ˙GeF3⋯˙CF3 complex, respectively (see Table 3).
Taken together, these results disclose that the stability of the studied complexes-containing Lewis bases, Lewis acids, and radicals are attributed mainly to the dispersion forces, with enhanced contributions from electrostatic and induction forces.
The results obtained from the NCI isosurfaces demonstrated the favorable −σ-hole and −R˙-hole interactions, showing disc-shaped green isosurfaces between the interacting fragments (Fig. S4†). Besides, the green isosurfaces became larger with increasing interaction strength, converting to blue ones for the strong interactions as observed in the F–Ge–F3⋯NCH complex. Moreover, the corresponding spikes of sign (λ2)ρ at low densities confirmed the existence of such attractive interactions (sign (λ2)ρ < 0). The location of spikes had a greater deviation from zero, and its shape became broader with enhancing the interaction energy (see Fig. S5†).
For +σ-hole and +R˙-hole interactions, disc-shaped green isosurfaces were also noticed for the studied complexes, asserting the existence of such interactions (see Fig. S4†). Furthermore, green isosurfaces were observed between the F3 atoms and Lewis acids, revealing their role in stabilizing the Lewis acids-containing complexes. The favorable interactions were also characterized by spikes at the negative sign (λ2)ρ. For instance, the spikes corresponding to each of H–C–F3⋯HCN and ⋯˙CF3 complexes were located at negative sign (λ2)ρ, indicating the probability of such interactions to dwell (see Fig. S5†).
In summary, these findings underlined that the −σ-hole and −R˙-hole interactions are the most favorable ones, where the large green isosurfaces and broad spikes were observed in their complexes.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d0ra09564h |
This journal is © The Royal Society of Chemistry 2021 |