Next Article in Journal
Emulsion Stabilization with Functionalized Cellulose Nanoparticles Fabricated Using Deep Eutectic Solvents
Next Article in Special Issue
Crystallographic and Computational Characterization of Methyl Tetrel Bonding in S-Adenosylmethionine-Dependent Methyltransferases
Previous Article in Journal
The Influence of Vacuum Impregnation on Nutritional Properties of Fluidized Bed Dried Kale (Brassica oleracea L. Var. Acephala) Leaves
Previous Article in Special Issue
On the Power of Geometry over Tetrel Bonds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantitative Assessment of Tetrel Bonding Utilizing Vibrational Spectroscopy

Computational and Theoretical Chemistry Group, Department of Chemistry, Southern Methodist University, 3215 Daniel Avenue, Dallas, TX 75275-0314, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2018, 23(11), 2763; https://doi.org/10.3390/molecules23112763
Submission received: 2 October 2018 / Revised: 19 October 2018 / Accepted: 20 October 2018 / Published: 25 October 2018
(This article belongs to the Special Issue Tetrel Bonds)

Abstract

:
A set of 35 representative neutral and charged tetrel complexes was investigated with the objective of finding the factors that influence the strength of tetrel bonding involving single bonded C, Si, and Ge donors and double bonded C or Si donors. For the first time, we introduced an intrinsic bond strength measure for tetrel bonding, derived from calculated vibrational spectroscopy data obtained at the CCSD(T)/aug-cc-pVTZ level of theory and used this measure to rationalize and order the tetrel bonds. Our study revealed that the strength of tetrel bonds is affected by several factors, such as the magnitude of the σ-hole in the tetrel atom, the negative electrostatic potential at the lone pair of the tetrel-acceptor, the positive charge at the peripheral hydrogen of the tetrel-donor, the exchange-repulsion between the lone pair orbitals of the peripheral atoms of the tetrel-donor and the heteroatom of the tetrel-acceptor, and the stabilization brought about by electron delocalization. Thus, focusing on just one or two of these factors, in particular, the σ-hole description can only lead to an incomplete picture. Tetrel bonding covers a range of −1.4 to −26 kcal/mol, which can be strengthened by substituting the peripheral ligands with electron-withdrawing substituents and by positively charged tetrel-donors or negatively charged tetrel-acceptors.

Graphical Abstract

1. Introduction

Noncovalent interactions (NCIs) have received increasing attention in the last two decades [1,2,3] due to their technological and fundamental importance in physics, chemistry, and biology [4,5,6]. Despite of the fact that NCIs are weak compared to covalent bonds (about an order of magnitude smaller), the importance of NCIs absolutely cannot be neglected [7,8,9]. They are ubiquitous and play a significant role in determining the properties of matter from small molecules to supramolecular systems like DNA and proteins [7,10]. They stabilize molecular structures [11,12], construct supramolecular materials [13], lower the activation energy of chemical reactions [14], and regulate the properties of crystal materials [15]. A series of different types of NCIs has been reported, namely, hydrogen bonds [16,17,18,19,20,21,22], aerogen bonds (group 18) [23,24,25,26], halogen bonds (Group 17) [27,28,29,30,31,32], chalcogen bonds (Group 16) [33,34,35,36,37], pnicogen bonds (Group 15) [38,39,40,41,42], tetrel bonds (Group 14) [43,44,45,46,47,48,49], and triel bonds (Group 13) [50,51].
Recently, tetrel bonding has found many applications due to its unique properties, such as strength, directionality, and origin of attraction [52]. Tetrel bonds play an important role in crystal engineering and supramolecular chemistry as a new potential molecular linker [44,53] and in dynamical processes such as protein folding and ligand–acceptor interactions [54,55,56]. Tetrel bonds also play an important role in the preliminary stages of SN2 reactions [57] and hydrophobic interactions [58,59]. The formation of tetrel bonds can be understood as an interaction between an electron-deficient tetrel atom of a Lewis acid (tetrel donor, T-donor) and an electron-rich of a Lewis base (tetrel acceptor, T-acceptor) (see Figure 1) [53]. The Lewis base (T-acceptor) can be any electron-rich entity possesing a lone pair [60,61,62,63], a π -system [55,64], an anion [44,65], etc. To explain the formation of a tetrel bond via σ -hole interactions, Politzer, Murray, and Clark suggested an interaction between a region of positive electrostatic potential as a result of diminished electron density on the tetrel atom (T-donor) and a region of negative electrostatic potential on an electron-rich atom (T-acceptor) [3,58,66,67,68]. The diminished electron density on the tetrel atom occurs as a result of electrons being mostly localized in the bonding region, which leaves a deficiency of electronic density in the outer lobe of the p-type valence orbital along the extension of the covalent bonding on the tetrel atom [69].
A series of experimental studies was conducted to identify and characterize tetrel bonding. The first convincing evidence of tetrel bonding was reported by Jönson and co-workers in 1975, where they observed that the carbon atoms of the carbon dioxide dimer can attractively interact with the lone pair of the oxygen of water [70] which was confirmed nine years later via microwave spectroscopic analysis by Klemperer and co-workers [71]. Recently, Guru-Row and co-workers provided experimental evidence of tetrel bonding based on an X-ray charge density analysis [43]. They revealed the existence of a bond path connecting the oxygen atom with the −CH3 carbon atom in R3N+−CH3···OH complexes [43]. Mitzel and co-workers discussed Si···N tetrel bonding in the crystalline Si(ONMe2)4 [72,73]. Evidence of tetrel bonding has also been observed by NMR spectroscopy [74]. The chemical shifts, quadrupolar couplings, and J-coupling are sensitive to the presence of tetrel bonding. For example, the J-coupling constant for (secondary) tetrel bonds has a magnitude of about 3 Hz [75].
The strength of tetrel bonding can be enhanced by cooperative effects [76,77,78] in conjunction with hydrogen bonding [79,80], halogen bonding [80], chalcogen bonding [81,82], lithium bonding [83], or with other tetrel bonding [84]. Cooperative effects in tetrel bonding [85,86,87] play an important role in crystal materials, chemical reactions, and biological systems [78,88,89]. Thus, the understanding of the strength and the nature of tetrel bonding is the key to understanding its properties. The molecular electrostatic potential ( V s ) and its maximum value ( V s , m a x ) are commonly used measures to quantify the strength of the σ -hole interaction [58,90,91]. A limited correlation between the interaction energies and the value of the V s , m a x has been reported by several authors [27,28,61,66,92]. However, very recently, Scheiner and co-workers pointed out that the maximum magnitude of the molecular electrostatic potential is not an ideal bond strength indicator [93,94]. Therefore, there is an urgent need for a qualified intrinsic bond strength descriptor, which we address in the present work.
One of the most common measures for quantifying the chemical bond strength is the bond dissociation energy (BDE) or the bond dissociation enthalpy (BDH). It has been shown that the BDE or BDH has the limitation of describing the intrinsic strength of a bond [95,96,97,98] because it includes the geometry relaxation of the fragments as well as the reorganization of the electron density. The intrinsic bond strength based on the local mode force constants k a measures the bond strength with only infinitesimal changes in the electronic structure of the molecule, thus excluding misleading additional contributions from the relaxation of the fragments. Many examples show that a chemical bond may have a large value of k a but a low BDE, vice versa [40,96,97].
Vibrational spectroscopy is an important tool that has been used to identify and characterize small-to-medium-sized molecules [99,100,101]. However, normal vibrational modes are of limited use as bond strength measure due to mode–mode coupling. A major breakthrough was achieved by the work of Konkoli and Cremer where the use of vibrational spectroscopy as an intrinsic bond strength measure via local vibrational modes was refined [102,103]. The intrinsic strength of chemical bonds is probed using the associated local stretching force constants k a [104,105,106]. The local stretching force constants k a have been successfully used to determine the intrinsic bond strength of covalent bonds such as CC bonds [105,107,108,109], NN bonds [110], NF bonds [98], CO bonds [111], and CX (X = F, Cl, Br, I) bonds [112,113,114,115] as well as weak chemical interactions such as hydrogen bonding [18,19,20,21,22], halogen bonding [30,116,117,118], pnicogen bonding [40,41,42], chalcogen bonding [96,97], and recently, BH···π interactions [119,120]. In this study, we investigate the strength and the nature of the tetrel bonds for a representative set of 35 complexes (see Figure 2) and also compare tetrel bonding with halogen and chalcogen bonding.
The main objectives of the present work are (i) to quantify the impact of changing the tetrel atom, its substituents, and the tetrel acceptor on the tetrel bond strength; (ii) to better understand the interplay between various electronic effects such as electrostatics, covalent contributions to tetrel bonding, exchange-repulsion between the tetrel acceptor and the peripheral ligands (R, R’, and R’’) of the tetrel-donor, etc.; (iii) to compare the strength of tetrel bonds with halogen (XBs), chalcogen (ChB), and pnicogen bonds (PnB); and (iv) to develop an effective strategy to tune the strength of the tetrel bond.

2. Computational Methods

To evaluate the key factors that influence the strength of the tetrel bonds, geometry optimizations and normal vibrational modes of complexes 135 (see Figure 2), monomers 36–60 (see Table 2, Figure 3, and Supporting Information Figure S1), and reference molecules R1R2 were calculated using the coupled cluster theory with singles, doubles, and perturbative triples (CCSD(T)) [121,122] in combination with the Dunning’s aug-cc-pVTZ basis set [123,124,125] which contains diffuse basis functions for describing the charge distribution of hetero-atoms, anions, and also, the dispersion interactions in tetrel bonds. All CCSD(T) calculations were carried out using a convergence criterion of 10 7 Hartree bohr−1 for geometry and a threshold of 10 9 for self-consistent field and CC-amplitudes.
Normal vibrational modes were converted into local vibrational modes using the Konkoli–Cremer method [102,103,104,107]. The electronic and mass coupling between normal vibrational modes were eliminated using the mass-decoupled analogue of the Wilson equation [107,126]. The resulting local vibrational modes, which were free from any mode-mode coupling, were associated with a given internal coordinate q n (bond length, bond angle, dihedral angle, etc.), which could be connected to normal vibrational modes in an one-to-one relationship via the Adiabatic Connection Scheme (ACS) [104,108]. The local force constant k a , obtained from the corresponding local vibrational mode, was used to measure the intrinsic bond strength of the tetrel bonds.
For simplification, the local force constant k a was converted to the bond strength order (BSO n) using a power relationship of the generalized Badger rule [127]:
n = a ( k a ) b .
The constants a = 0.418 and b = 0.564 were obtained by two references of well-defined bond order, namely the FF bond in F2 (n = 1.0) and the three-center-four-electron bond in F3 (n = 0.5), as previously done in a study of halogen bonds [30] and other noncovalent interactions [117].
The binding energy was separated into two contributions:
Δ E = E i n t + E d e f .
E i n t is the interaction energy for the frozen geometry of the monomers, and the deformation energy ( E d e f ) is the energetic difference between the monomers’ frozen geometry and their minimum energy geometry. The counterpoise correction (CP) [128] is usually used to eliminate the basis set superposition error (BSSE) present in E i n t . However, BSSE often compensates for the error caused by an incomplete basis set; consequently, uncorrected E i n t values can be closer to the complete basis set limit (CBS) [117,129]. To test if this was the case, we compared CP-corrected and uncorrected CCSD(T)/aug-cc-pVTZ results to the ones obtained using the more saturated pentuple zeta basis set aug-cc-pV5Z [123] (see Supporting Information Table S1). The latter results were obtained by employing the domain-based local pair natural orbital (DLPNO) [130,131] approximation to CCSD(T) using a tight convergence criteria to ensure that the errors caused by the DLPNO approximation were negligible. It turned out that the uncorrected CCSD(T)/aug-cc-pVTZ values of E i n t were, on average, closer to the CP-corrected and uncorrected DLPNO-CCSD(T)/aug-cc-pV5Z results than the CP-corrected CCSD(T)/aug-cc-pVTZ values (see Supporting Information Tables S1 and S2). Therefore, in the next sections, only Δ E values without counterpoise correction are discussed (see Table 1).
CCSD(T) calculations were performed using the CFOUR program [132,133], whereas DLPNO- CCSD(T) calculations were done in ORCA 4.0 [134]. Analytical vibrational frequencies were used to verify that each equilibrium geometry obtained by CCSD(T) corresponded to a geometry minimum. The charge distribution was calculated with the natural population analysis (NPA) within the Natural Bond Orbital (NBO) scheme [135,136] using the NBO6 program [137,138]. The electron density ρ (r) and the energy density H(r) at the T···A (T: tetrel atom and A: tetrel-acceptor) electron density critical point (r) were calculated using the AIMALL program [139]. The molecular electrostatic potentials of T-donors and T-acceptors mapped onto the 0.001 e/bohr3 electron density surface were calculated using the Multiwfn3.5 [140] program. Noncovalent interaction (NCI) plots were calculated using the NCIplot program [141]. NBO charges, as well as, H(r), ρ ( r ) and V(r) were derived from CCSD(T) response densities obtained from CFOUR calculations with the help of MOLBO and Molden2AIM scripts [142]. Local mode force constants and frequencies were calculated using the COLOGNE18 program package [143].

3. Results and Discussion

Table 1 summarizes the complex binding energy Δ E, the counterpoise corrected binding energy Δ E c p , the monomers’ deformation energy E d e f , the distance r(TA) between the tetrel-donor atom (T = C, Si or Ge) and an heteroatom (A = F, O, N or Cl) of the tetrel-acceptor (T-acceptor), also called the tetrel bond (TB) distance, the distance r(XT) between the T donor atom and the donor group or atom (X = H, F, Cl, Br, OH, =O or =S), the intermonomer charge transfer (CT) obtained from the natural population analysis (NPA), the electron density ρ b and the energy density H b at the density critical point associated with TB, the local stretching force constant of TA ( k a (TA)) and XT ( k a (XT)) and the BSO n of the TA and XT. The calculated NBO atomic charges are given in Figure 2. T donor properties such as the maximum electrostatic potential at the σ -hole region of the tetrel atom ( V s m a x ), the total dipole moment, and the isotropic polarizability are listed in Table 2. The BSO n values of all TB are given as functions of their local stretching force constant k a (T···A) in Figure 4. Similar to previous studies [30,34,40,116,117,118,144], we determined the covalent character of the TB by utilizing the energy density H b at the density critical point of the TB (Figure 5); electrostatic interactions were characterized by having positive H b values, whereas, according to the Cremer–Kraka criterion, [145,146], covalent interactions have negative H b values, indicating that the accumulated electron density at the interactions region stabilizes the complex. Although the relationship between BSO n and H b is scattered, the TB strength tends to increase with the increasing covalent character of the interaction, especially among neutral complexes.

3.1. Tetrel Bonds (TB) in Neutral Complexes

TBs involving C donors: Tetrel bonds (TBs) involving a neutral sp3 hybridized carbon as a tetrel donor and neutral tetrel acceptors (complexes 1–7) have weak interactions (BSO n≤ 0.081; Δ E −2.25 kcal/mol). The energy density at the TB density critical point is destabilizing ( H b ≥ 0.007 Hartree/Å3), indicating that these TBs are electrostatic in nature. TB with C donors that have peripheral H ligands (1–6) are not only stabilized by a lone pair of the T acceptor lp(A)- σ -hole electrostatic attraction but also by the electrostatic attraction between the positive charge at the hydrogen atoms (Hs) and the negative charge at lp(A) (see the NPA atomic charges in Figure 2 and the monomers’ electrostatic potential in Table 2 and Figure 3). The presence of the later interactions were also verified by noncovalent interaction plots (NCIs plots), which showed a weak (attractive) electrostatic interaction (see Supporting Information Figure S2). Complex 7 (CF4···NH3) shows that an attractive interaction can be formed even in the absence of positively charged Hs on the T-donor. However, in 7, there is no electron density path connecting the N of the T acceptor to the C of the T donor indicating the formation of a very weak dispersive interaction, as pointed by Grabowski [147]. The calculated spectra of 7 clearly shows an intermonomer stretching vibration of A1 symmetry at 73 cm−1. Decomposition of the normal vibrational mode into the local vibrational mode shows that the CN local stretching mode contributes solely to this normal vibrational mode, confirming the existence of tetrel bonding in 7. As revealed by the NCI plot analysis, the peripheral H ligands of the T acceptor have an additional weak (attractive) electrostatic interaction with the peripheral F ligands of the T donor, which provides additional stabilization to the complex (see Supporting Information Figure S2).
The TB strength in the series of C donors and neutral T acceptors (1–7) shows only a small variation (0.021 for n and 0.87 kcal/mol for Δ E ) which is affected by several factors such as the positive charge at the Hs (5>4>36), the negative electrostatic potential at lp(A) of the T-acceptor (NH3<OH2< FH), and the intermonomer distance (e.g., 3.035 Å (2) compared to 3.218 (3)). It is noteworthy that for X−CH3···NH3 complexes, the TB strength weakens in the order of 3 (X = F) >4 (X = Cl) >5 (X = Br) >6 (X = OH) as the magnitude of the σ -hole decreases (36 (F−CH3)) >37 (Cl−CH3) >38 (Br−CH3) > 39 (OH−CH3) (see Figure 3).
TB involving Si donors: The strength of complexes 8–14 can be understood mostly on the basis of the extreme values of the electrostatic potential of the monomers. First, by varying the T-acceptor in FSiH3···A, where A = FH (8), OH2 (9), NH3 (10), the TB strength trends follows the increase in magnitude of the negative potential at the lp(A) of the Lewis base. It is noticeable that V s , m i n (A) and V s , m a x (T) in 9 are not aligned as in 2, indicating that even in this case, the stabilization brought by electron delocalization involving the highest occupied orbital HOMO of H2O (the lp(O) orbital of B1 symmetry, see Supporting Information Figure S3) and the lowest unoccupied orbital of FSiH3 ( σ (SiF) orbital) can influence the geometry of the complex (see Supporting Information Figure S4). Second, by varying the T donors in the XSiH3···NH3 complexes, the TB strength decreases in the order 10 (X = F) > 11 (X = Cl) > 12 (X = Br) ≈13 (X = OH) > 14 (X = H) as the Vs at the σ -hole decreases (41 (F−SiH3) > 42 (Cl−SiH3) > 43 (Br−SiH3) > 44 (HO−SiH3) > 45 (SiH4)). The only exception is 44, which has a more positive NBO charge at Si (1277 me compared to 852 me in 43) but a less positive V s , m a x (1.14 (44) compared to 1.46 (43), see Table 2 and Supporting Information Figure S1). This could be caused by the stronger σ (O-Si) orbital contraction, whereby the Si atom in HOSiH3 is more electron-deficient than the Si atom in BrSiH3, but due to the higher electronegativity of O, the σ (O-Si) is more compact than σ (Br-Si), resulting in a better shielded Si nucleus, reflected in the less positive potential at the σ -hole region (given by V s , m a x ). Substituting F in FSiH3 would lead to an even stronger σ (F-Si) orbital contraction, but this effect would be overcome by the higher electron deficiency of Si.
Stepwise fluorination of SiH4: The successive fluorination of SiH4 (complexes 10 and 15–17) impacts both the strength and the nature of the TB. Substituting the H collinear to the TB in complex 14 results in complex 10, which has a stronger TB interaction due to the higher V s at the σ -hole region and is due to the partial covalent character of the interaction ( H b = −0.033 Hartree/Å for complex 10), which can be understood on the basis of molecular orbital interactions, as the electron delocalization from the highest occupied molecular orbital (HOMO) of the NH3 (lone pair orbital of N, lp(N)) into the lowest unoccupied molecular orbital (LUMO) of FSiH3 (a σ (FSi) orbital see Supporting Information Figure S5).
A second fluorine substituent (complex 15) shows three different electronic effects: (i) the second fluorine withdraws the electron density from Si, decreasing its covalent radius and thus increasing the V s at the σ -hole region, resulting in a stronger σ -hole-lp(N) electrostatic attraction; (ii) the lp(N) σ (SiF) electron delocalization is not restricted to the σ orbital of the Si–F bond that is collinear to the lp(A) but can also take place to the σ orbital of the second Si–F bond (see Table 3); (iii) the orbital effect of the exchange-repulsion between the lone pair orbitals of the peripheral fluorine lp(F) and lp(N) orbital. Effects (i) and (ii) are responsible for the 0.123 Å shorter TB in 15 compared to 10. However, due to effect (iii), complexes 10 and 15 have similar TB strengths (n = 0.116 for 10, 0.103 for 15; Δ E = 6.8 kcal/mol for 10, and 7.0 kcal/mol for 15).
The addition of a third fluorine substituent (complexes 16a and 16b) leads to shorter and stronger TBs (BSO n = 0.191 (16a) and 0.280 (16b) compared to 0.103 (15)). However, this great increase in the TB strength (especially for 16b) is not reflected by the binding energies of 16a and 16b ( Δ E = −7.7 kcal/mol for 16a and −6.3 kcal/mol for 16b). The reason for the unexpectedly low Δ E values of 16a and 16b is due to the energetic cost associated with the geometric deformation of the monomers upon complexation ( E d e f = 11.8 kcal/mol for 16a, 21.2 kcal/mol for 16b). Monomer deformation is mostly caused by the lp(F)–lp(N) exchange-repulsion (effect (iii)), which pushes the peripheral ligands towards the bond collinear to the TB. For example, there is a decrease of 12.3° in the H–Si–F bond angle of HSiF3 upon the formation of 16b. Monomer deformation and the steric effect on TB complexes were also topics of a recent study carried by Scheiner [148,149]. It is noteworthy that the strongest TB between SiF3H and NH3 is collinear to the V s at the σ -hole of the Si–H bond (complex 16b), instead of the most positive potential at the σ -hole of the Si–F bond as one would expect from the σ -hole model or from steric considerations. The stronger and more covalent bond in 16b is due to the higher stabilization energies brought by electron delocalization from lp(N) into σ (FSi) and into the σ (HSi) unoccupied orbital, as shown in Table 3. Even if the TB of 16b is elongated to match the TB distance of 16a, the NBO second-order delocalization energies of 16b are still higher than those of 16a (see Table 3).
The addition of a fourth fluorine (complex 17) makes the TB even stronger (n = 0.335 for 17), compared to 0.280 for 16), which is a consequence of the more positive V s and of the higher electron delocalization that occurs from lp(N) to σ (FSi) compared to σ (H-Si) (see Table 3 and Supporting Information Figure S5). The substitution of the H collinear to the TB in 16b by a fluorine does not increase the steric repulsion between the monomer ( E d e f of 17 is almost the same of 16b), and as a result, Δ E is 5.1 kcal/mol more stable.
TB involving Ge donors: The germanium electron density is more easily polarized by an electronegative substituent than silicon. As a result, V s at the σ -hole of Ge-donors (49–52) are higher than Si donors (see Table 2), the only exception being GeH4 (53), which is a consequence of the higher electronegativity of Ge ( χ (Si): 1.74 compared to χ (Ge): 2.02). Due to the stronger lp(N)- σ -hole electrostatic attraction, mono-substituted Ge-donors (18, 19, 20, 21) form stronger TBs than mono-substituted Si donors (10, 11, 12, 13) when paired with the NH3 T acceptor. Conversely, the covalent component of this interaction is slightly reduced because of the more diffuse nature of Ge orbitals (see CT and H b values on Table 1). Similar to C donors and Si donors, for X−GeH3···NH3 complexes, the TB strength decreases in the order of 18 (X = F) >19 (X = Cl) >20 (X = Br) >21 (X = OH), as the magnitude of the σ -hole decreases (49 (F−GeH3)) >50 (Cl−GeH3) >51 (Br−GeH3) > 52 (OH−GeH3) (see Table 2 and Figure 3).
TB in double bonded C and Si donors: When a carbon atom forms a π -bond, density is moved from a p-orbital of the carbon into the π -bond, resulting in a depletion of electron density at the C and the formation of a region of positive potential called a π -hole [150,151]. As noted previously by various authors [152,153,154], the lp(A)- π -hole electrostatic attraction is an important component of the TB involving double bonded C, Si donors. Another important characteristic of these T-donors is the existence of a low lying empty π (XT) orbital which is capable of accepting electron density from the lp(A) orbital of the T-acceptor. In order to evaluate strategies to strengthen TB involving double bonded C and Si donors, complexes 23–27 were investigated. Due to the D h and C v symmetry of CO2 (54) and SCO (55), respectively, the π -bond density of these monomers has a central constriction with a negative V s , leaving a belt-shaped π -hole around the C atom (see Figure 3). In SCO (55), a chalcogen bond is also possible due to the formation of a positive V s at the σ -hole of sulfur. The TBs between CO2 (23) and SCO (24) T donors and the prototypical T acceptor NH3 are weak (n = 0.100 (23), 0.074 (24)) and electrostatic in nature ( H b > 0.002 Hartree/Å).
Substituting a CO double bond by two CF single bonds in 23 and 24 results in complexes CF2O···NH3 (25) and CF2S···NH3 (26), respectively. The T-donors of these complexes are characterized by having a higher V s at the π -hole ( V s = 1.85 eV (56), 1.29 eV (57) compared to 1.18 eV (54) and 0.64 eV (55)), resulting in stronger TBs (n = 0.127 (25), 0.105 (26a) compared to 0.100 (23) and 0.074 (24)). The atypically strong (n = 0.508), highly covalent ( H b = −1.339 Hartree/Å 3 ) and short interactions (r(TA) = 1.7 Å) found in complex 26b are formed at the expense of breaking the CS π -bond (n(C=S) decreases from 1.214 in 57 to 0.891 in 26b). The energetic cost involved in the deformation of the monomers of 26b is 1.45 kcal/mol higher than the stabilization brought by complexation ( E i n t ); hence, the 26b is less stable than the separated monomers. A small energetic barrier in the dissociative direction separates 26b from the electrostatic TB complex 26a (see Figure 6). An even stronger (n = 0.589) but less covalent interaction ( H b = −0.224 Hartree/Å 3 ) is formed between SiF2O and NH3 (complex 27). In this complex, the SiO double bond is kept almost unaltered (n = 1.465 (58); 1.425 (27)); consequently, the deformation energy is relatively low compared to the stabilization energy brought by the complexation, resulting in a binding energy of −44.14 kcal/mol.

3.2. Charge-Assisted Tetrel Bonds

Charge-assisted interactions: Similar to other NCIs [30,34,116,118,155], TBs can be strengthened by having a positively-charged T donor or a negatively-charged T acceptor 3035. The simplest positively charged C donor is CH3+ (isoelectronic to BH3). However, due to the availability of an empty p-orbital to coordinate with the lone pairs of the NH3, the C–N bond in 28 clearly differs from the tetrel bonds. This covalent bond in [CH3−NH3]+ (28) is much stronger (n = 0.882; Δ E = −110.26 kcal/mol) and covalent ( H b = −1.952 Hartree/Å 3 ) compared to a TB. The existence of a covalent bond in complex 28 is also confirmed by the NCI plot, showing that there is no noncovalent interaction (see Supporting Information Figure S2). On the other hand, the cationic pnicogen donor FNH3+, isoelectronic to FCH3, forms a noncovalent interaction with the NH3 (complex 29) which closely resembles the ones formed by neutral C donors (1–7) characterized by an electrostatic nature ( H b ) and an interaction collinear to the X–T bond. The only difference is the higher V s at the σ -hole ( V s = 8.58 eV), which results in a stronger electrostatic interaction (n = 0.236; Δ E = −23.14 kcal/mol).
An increase in the TB strength of as much as 105% (31) occurs for complexes involving a neutral T donor and a chloride anion as a T acceptor. This increase does not affect TB strength trends, such as FGeH3 >FSiH3>FCH3, and CO2>SCO, nor the trends in the covalent character of the TB in these series. Conversely, complex CF2S forms a weaker bond with Cl compared to NH3 (n = 0.441 (35) 0.508 (26b)), but has a highly negative binding energy ( Δ E = −16.81 kcal/mol of 35 compared to the 1.45 kcal/mol of 26b). The inverse relationships between the bond strength and interaction energies ( E i n t ) or binding energies ( Δ E) between these complexes indicate that the local C–N or C–Cl stretching is not the determining factor for the complex stabilization. Other components of the interaction which do not contribute to the strength of T-A local stretching also stabilize the complex. Figure 6 shows the dissociation curves for the CF2S···NH3 (26) and CF2S···Cl (35). Only 26 has a minimum geometry with an electrostatic TB (26a), whereas the dissociation curve for 35 has a flat region around r(TA) = 2.9Å separating the electrostatic interaction found for long TBs from the strong covalent interactions in 35. The barrier energy from 26a to 26b is about 3 kcal/mol.

3.3. Tetrel Bonds vs. Other Noncovalent Interactions

We also compared TB with other noncovalent interactions, such as halogen (XB), chalcogen (ChB), and pnicogen (PnB), in mono-fluorinated systems involving the third period series FCl, FSH, FPH2 with a medium (OH2) and a strong (NH3) Lewis base (see Table 4). It was clearly shown that the TB formed by FSiH3 tend to be weaker than the other noncovalent interactions, the only exception being FSiH3···Cl (31) (n = 0.238) which is slightly stronger than FPH2···Cl (n = 0.214) but weaker than FSH···Cl (n = 0.264) and FCl···Cl (n = 0.382). Increasing the polarizability of the T donor moving FSiH3 to FGeH3 does increase the strength of the tetrel bond enough to compete with the halogen bonds formed by FCl. A better strategy for obtaining TBs that are strong enough to compete with halogen bonds and other noncovalent interactions is the substitution of peripheral Hs in FSiH3 by fluorine atoms. SiF4···NH3 (17), for example, has a BSO of n = 0.335 compared to n = 0.216 for the FCl···NH3. A clear advantage of tetrel bonding is that even in the absence of a strong polarizing group collinear to the TB, such as in HSiF3···NH3, the TB is still stronger than other noncovalent interactions (n = 0.280 for 16b). This TB feature should be extensively explored to tune the strength of TB involving not only Si donors but also the heavier and more polarizable Ge and Sn donors.

4. Conclusions

In the present work, we investigated a set of 35 representative tetrel complexes ( Δ E = −1.4 to −26 kcal/mol) with the objective of finding the factors that influence the strengths of neutral and charged tetrel bonds involving C donors, Si donors, Ge donors, and double bonded C or Si donors. The strength of a tetrel bond is affected by the complex interplay of several factors, such as the magnitude of the σ -hole in the tetrel atom, the negative electrostatic potential at the lp(A) of the T acceptors, the positive charge at the peripheral hydrogen (Hs) of the T donors, exchange-repulsion between the lone pair orbitals of the peripheral atoms of the T donor, and the covalent character which can be rationalized on the basis of electron delocalization from the highest occupied molecular orbital (HOMO) of the T acceptor into the lowest unoccupied orbitals (LUMOs) of the T donor, which is not limited to σ (X−T) orbital but can also involve the peripheral substituents (orbital of σ (R−Si) character), allowing the formation of strong tetrel bonds, even in the absence of an electronegative X substituent collinear to the TB. This clearly shows that focusing on just one or two of these factors, in particular, the σ -hole description, can only lead to an incomplete picture [93,94,156,157]. In this work, we derived, for the first time, the intrinsic bond strength of tetrel bonds from calculated vibrational spectroscopy data, which, combined with NBO charges, charge transfer, dipole moments, electrostatic potentials, electron and energy density distributions, difference density distributions, and noncovalent interaction plots calculated at the CCSD(T)/aug-cc-pVTZ level of theory, led to a complete insight into how different electronic effects influence the intrinsic strength of the tetrel bonding.
  • Tetrel bonding becomes stronger as the atomic mass of the tetrel center increases as a consequence of increasing the polarizability.
  • For X−TH3···NH3 complexes, the tetrel bond strength weakens in the order (X = F) > (X = Cl) > (X = Br) ≥ (X = OH) as the magnitude of the σ -hole decreases in the order of F−TH3>Cl−TH3>Br−TH3≥ OH−TH3.
  • Successive fluorination of SiH4 impacts both the strength and the nature of the tetrel bond. The successive fluorinations result in stronger tetrel bonding as a consequence of (i) higher V s at the σ -hole region; (ii) the partial covalent character of the interaction; (iii) higher electron delocalization that occurs from the highest occupied molecular orbital (HOMO) of the T acceptor to the lowest unoccupied molecular orbital (LUMO) of the T donor. In this series, the binding energy trend deviates from BSO n values due to the high energetic cost associated with the geometric deformation of the monomers upon complexation ( E d e f ) which is a consequence of the exchange-repulsion between the lone pair orbitals of the peripheral atoms of the T donor.
  • Tetrel bonds in double bonded C donors, e.g., CO2 with NH3, are weak and electrostatic in nature. Substituting a C=O double bond with an electron withdrawing group (F atoms) strengthens the tetrel bond.
  • A positively-charged Tdonor or negatively-charged T-acceptor strengthens the tetrel bond. It creates higher V s at the σ -hole, resulting in a stronger electrostatic interaction.
We suggest that future materials based on strong tetrel bonding should be based on Si or heavier tetrel atoms, such as Ge and Sn, combined with peripheral fluorine ligands. Due to the larger size of Ge and Sn, the deformation energy in XGeF3···NH3 or XSnF3···NH3 should be smaller than XSiF3···NH3, making these complexes substantially more stable than XSiF3···NH3.
Although all complexes discussed in this paper represent the most stable tetrel-bonded complexes, not all of them represent the most stable structure possible (global minimum). For example, the hydrogen bonded complexes FH···CFH3 (1), OH2···CF3H (2), NH3···CF3H (3) are more stable than the tetrel-bonded complexes. However, a detailed analysis of the competition between tetrel bonds and other noncovalent interactions will be studied in the near future.

Supplementary Materials

The following are available online, Figure S1: Schematic representation of monomers (36–63) with atomic charges from the natural population analysis, Figure S2: Noncovalent interactions (NCIs) plot of complexes 1–35, Figure S3: Selected molecular orbitals of the T-acceptors, Figure S4: Electron difference density distributions Δ ρ ( r ) for complexes 1–35, Figure S5: Combination of donor and acceptor NBO orbitals involved in the electron delocalization of selected complexes; Table S1: Comparison between DLPNO-CCSD(T)/aug-cc-pV5Z and CCSD(T)/aug-cc-pVTZ energies, Table S2: Deviation from DLPNO-CCSD(T)/aug-cc-pV5Z interaction energies, Table S3: Atomic Cartesian coordinates in Å of complexes 1–35.

Author Contributions

Conceptualization, D.S., V.O., and E.K.; Methodology, E.K.; Formal Analysis, D.S. and V.O.; Investigation, D.S. and V.O.; Data Curation, D.S. and V.O.; Writing—Original Draft Preparation, D.S. and V.O.; Writing—Review & Editing, E.K.; Visualization, D.S. and V.O.; Supervision, E.K.

Funding

This research was funded by National Science Foundation grant number CHE 1464906.

Acknowledgments

We thank Alan Humason for proof-reading and also SMU’s HPC for providing excellent computational resources.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ACSAdibatic connection scheme
BSO nBond strength order
CCSD(T)Coupled cluster theory with singles, doubles, and perturbative triples
CTIntermonomer charge transfer
EDGElectron donating group
EWGElectron withdrawing group
HOMOHighest occupied molecular orbital
LUMOLowest unoccupied molecular orbital
NBONatural bond orbital
NCINoncovalent interaction
NPANatural population analysis
TBTetrel bond

References

  1. Schneider, H.J. Binding Mechanisms in Supramolecular Complexes. Angew. Chem. Int. Ed. 2009, 48, 3924–3977. [Google Scholar] [CrossRef] [PubMed]
  2. Bene, J.E.D.; Alkorta, I.; Elguero, J. Exploring the (H2C=PH2)+:N-Base Potential Surfaces: Complexes Stabilized by Pnicogen, Hydrogen, and Tetrel Bonds. J. Phys. Chem. A 2015, 119, 11701–11710. [Google Scholar] [CrossRef] [PubMed]
  3. Politzer, P.; Murray, J.S. Analysis of Halogen and Other σ-Hole Bonds in Crystals. Crystals 2018, 8, 42. [Google Scholar] [CrossRef]
  4. Dubecký, M.; Mitas, L.; Jurečka, P. Noncovalent Interactions by Quantum Monte Carlo. Chem. Rev. 2016, 116, 5188–5215. [Google Scholar] [CrossRef] [PubMed]
  5. Scheiner, S. Hydrogen Bonding: A Theoretical Perspective; Oxford University Press: New York, NY, USA, 1997. [Google Scholar]
  6. 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] [PubMed] [Green Version]
  7. Riley, K.E.; Hobza, P. Noncovalent interactions in biochemistry. WIREs: Comput. Mol. Sci. 2011, 1, 3–17. [Google Scholar] [CrossRef] [Green Version]
  8. Alkorta, I.; Legon, A. Nucleophilicities of Lewis Bases B and Electrophilicities of Lewis Acids A Determined from the Dissociation Energies of Complexes B···A Involving Hydrogen Bonds, Tetrel Bonds, Pnictogen Bonds, Chalcogen Bonds and Halogen Bonds. Molecules 2017, 22, 1786. [Google Scholar] [CrossRef] [PubMed]
  9. Gholipour, A. Mutual interplay between pnicogen-π and tetrel bond in PF3⊥X-Pyr···SiH3CN complexes: NMR, SAPT, AIM, NBO, and MEP analysis. Struct. Chem. 2018. [Google Scholar] [CrossRef]
  10. Christensen, A.S.; Kubař, T.; Cui, Q.; Elstner, M. Semiempirical Quantum Mechanical Methods for Noncovalent Interactions for Chemical and Biochemical Applications. Chem. Rev. 2016, 116, 5301–5337. [Google Scholar] [CrossRef] [PubMed]
  11. Sessions, R.B.; Gibbs, N.; Dempsey, C.E. Hydrogen Bonding in Helical Polypeptides from Molecular Dynamics Simulations and Amide Hydrogen Exchange Analysis: Alamethicin and Melittin in Methanol. Biophys. J. 1998, 74, 138–152. [Google Scholar] [CrossRef] [Green Version]
  12. Bene, J.E.D.; Alkorta, I.; Elguero, J. Anionic complexes of F and Cl with substituted methanes: Hydrogen, halogen, and tetrel bonds. Chem. Phys. Lett. 2016, 655–656, 115–119. [Google Scholar] [CrossRef]
  13. Priimagi, A.; Cavallo, G.; Metrangolo, P.; Resnati, G. The Halogen Bond in the Design of Functional Supramolecular Materials: Recent Advances. Acc. Chem. Res. 2013, 46, 2686–2695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Shi, F.Q.; Li, X.; Xia, Y.; Zhang, L.; Yu, Z.X. DFT Study of the Mechanisms of In Water Au(I)-Catalyzed Tandem [3,3]-Rearrangement/Nazarov Reaction/[1,2]-Hydrogen Shift of Enynyl Acetates: A Proton-Transport Catalysis Strategy in the Water-Catalyzed [1,2]-Hydrogen Shift. J. Am. Chem. Soc. 2007, 129, 15503–15512. [Google Scholar] [CrossRef] [PubMed]
  15. Shen, Q.J.; Jin, W.J. Strong halogen bonding of 1,2-diiodoperfluoroethane and 1,6-diiodoperfluorohexane with halide anions revealed by UV-Vis, FT-IR, NMR spectroscopes and crystallography. Phys. Chem. Chem. Phys. 2011, 13, 13721–13729. [Google Scholar] [CrossRef] [PubMed]
  16. Arunan, E.; Desiraju, G.R.; Klein, R.A.; Sadlej, J.; Scheiner, S.; Alkorta, I.; Clary, D.C.; Crabtree, R.H.; Dannenberg, J.J.; Hobza, P.; et al. Definition of the hydrogen bond (IUPAC Recommendations 2011). Pure Appl. Chem. 2011, 83, 1637–1641. [Google Scholar] [CrossRef] [Green Version]
  17. Arunan, E.; Desiraju, G.R.; Klein, R.A.; Sadlej, J.; Scheiner, S.; Alkorta, I.; Clary, D.C.; Crabtree, R.H.; Dannenberg, J.J.; Hobza, P.; et al. Defining the hydrogen bond: An account (IUPAC Technical Report). Pure Appl. Chem. 2011, 83, 1619–1636. [Google Scholar] [CrossRef] [Green Version]
  18. Freindorf, M.; Kraka, E.; Cremer, D. A comprehensive analysis of hydrogen bond interactions based on local vibrational modes. Int. J. Quant. Chem. 2012, 112, 3174–3187. [Google Scholar] [CrossRef]
  19. Kalescky, R.; Zou, W.; Kraka, E.; Cremer, D. Local vibrational modes of the water dimer—Comparison of theory and experiment. Chem. Phys. Lett. 2012, 554, 243–247. [Google Scholar] [CrossRef]
  20. Kalescky, R.; Kraka, E.; Cremer, D. Local vibrational modes of the formic acid dimer—The strength of the double hydrogen bond. Mol. Phys. 2013, 111, 1497–1510. [Google Scholar] [CrossRef]
  21. Kraka, E.; Freindorf, M.; Cremer, D. Chiral Discrimination by Vibrational Spectroscopy Utilizing Local Modes. Chirality 2013, 25, 185–196. [Google Scholar] [CrossRef] [PubMed]
  22. Tao, Y.; Zou, W.; Jia, J.; Li, W.; Cremer, D. Different Ways of Hydrogen Bonding in Water—Why Does Warm Water Freeze Faster than Cold Water? J. Theory Comp. Chem. 2016, 13, 55–76. [Google Scholar] [CrossRef] [PubMed]
  23. Bauzá, A.; Frontera, A. Aerogen Bonding Interaction: A New Supramolecular Force? Angew. Chem. Int. Ed. 2015, 54, 7340–7343. [Google Scholar] [CrossRef] [PubMed]
  24. Bauzá, A.; Frontera, A. Theoretical Study on the Dual Behavior of XeO3 and XeF4 toward Aromatic Rings: Lone Pair-π versus Aerogen-π Interactions. ChemPhysChem 2015, 16, 3625–3630. [Google Scholar] [CrossRef] [PubMed]
  25. Bauzá, A.; Frontera, A. π-Hole aerogen bonding interactions. Phys. Chem. Chem. Phys. 2015, 17, 24748–24753. [Google Scholar] [CrossRef] [PubMed]
  26. Frontera, A.; Bauzá, A. Concurrent aerogen bonding and lone pair/anion-π interactions in the stability of organoxenon derivatives: A combined CSD and ab initio study. Phys. Chem. Chem. Phys. 2017, 19, 30063–30068. [Google Scholar] [CrossRef] [PubMed]
  27. Desiraju, G.R.; Ho, P.S.; Kloo, L.; Legon, A.C.; Marquardt, R.; Metrangolo, P.; Politzer, P.; Resnati, G.; Rissanen, K. Definition of the halogen bond (IUPAC Recommendations 2013). Pure Appl. Chem. 2013, 85, 1711–1713. [Google Scholar] [CrossRef] [Green Version]
  28. Politzer, P.; Murray, J.S. Halogen Bonding: An Interim Discussion. ChemPhysChem 2013, 14, 278–294. [Google Scholar] [CrossRef] [PubMed]
  29. Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. The Halogen Bond. Chem. Rev. 2016, 116, 2478–2601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Oliveira, V.; Kraka, E.; Cremer, D. The intrinsic strength of the halogen bond: Electrostatic and covalent contributions described by coupled cluster theory. Phys. Chem. Chem. Phys. 2016, 18, 33031–33046. [Google Scholar] [CrossRef] [PubMed]
  31. Gilday, L.C.; Robinson, S.W.; Barendt, T.A.; Langton, M.J.; Mullaney, B.R.; Beer, P.D. Halogen Bonding in Supramolecular Chemistry. Chem. Rev. 2015, 115, 7118–7195. [Google Scholar] [CrossRef] [PubMed]
  32. Wolters, L.P.; Schyman, P.; Pavan, M.J.; Jorgensen, W.L.; Bickelhaupt, F.M.; Kozuch, S. The many faces of halogen bonding: A review of theoretical models and methods. WIREs Comput. Mol. Sci. 2014, 4, 523–540. [Google Scholar] [CrossRef]
  33. Alikhani, E.; Fuster, F.; Madebene, B.; Grabowski, S.J. Topological reaction sites—Very strong chalcogen bonds. Phys. Chem. Chem. Phys. 2014, 16, 2430–2442. [Google Scholar] [CrossRef] [PubMed]
  34. Oliveira, V.; Cremer, D.; Kraka, E. The Many Facets of Chalcogen Bonding: Described by Vibrational Spectroscopy. J. Phys. Chem. A 2017, 121, 6845–6862. [Google Scholar] [CrossRef] [PubMed]
  35. Gleiter, R.; Haberhauer, G.; Werz, D.B.; Rominger, F.; Bleiholder, C. From Noncovalent Chalcogen-Chalcogen Interactions to Supramolecular Aggregates: Experiments and Calculations. Chem. Rev. 2018, 118, 2010–2041. [Google Scholar] [CrossRef] [PubMed]
  36. Mahmudov, K.T.; Kopylovich, M.N.; da Silva, M.F.C.G.; Pombeiro, A.J.L. Chalcogen bonding in synthesis, catalysis and design of materials. Dalton Trans. 2017, 46, 10121–10138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Alkorta, I.; Elguero, J.; Bene, J.E.D. Complexes of O=C=S with Nitrogen Bases: Chalcogen Bonds, Tetrel Bonds, and Other Secondary Interactions. ChemPhysChem 2018, 19, 1886–1894. [Google Scholar] [CrossRef] [PubMed]
  38. Scheiner, S. The Pnicogen Bond: Its Relation to Hydrogen, Halogen, and Other Noncovalent Bonds. Acc. Chem. Res. 2012, 46, 280–288. [Google Scholar] [CrossRef] [PubMed]
  39. Sarkar, S.; Pavan, M.S.; Row, T.N.G. Experimental validation of ‘pnicogen bonding’ in nitrogen by charge density analysis. Phys. Chem. Chem. Phys. 2015, 17, 2330–2334. [Google Scholar] [CrossRef] [PubMed]
  40. Setiawan, D.; Kraka, E.; Cremer, D. Strength of the Pnicogen Bond in Complexes Involving Group 5A Elements N, P, and As. J. Phys. Chem. A 2014, 119, 1642–1656. [Google Scholar] [CrossRef] [PubMed]
  41. Setiawan, D.; Kraka, E.; Cremer, D. Description of pnicogen bonding with the help of vibrational spectroscopy—The missing link between theory and experiment. Chem. Phys. Lett. 2014, 614, 136–142. [Google Scholar] [CrossRef]
  42. Setiawan, D.; Cremer, D. Super-pnicogen bonding in the radical anion of the fluorophosphine dimer. Chem. Phys. Lett. 2016, 662, 182–187. [Google Scholar] [CrossRef] [Green Version]
  43. Thomas, S.P.; Pavan, M.S.; Row, T.N.G. Experimental evidence for ‘carbon bonding’ in the solid state from charge density analysis. Chem. Commun. 2014, 50, 49–51. [Google Scholar] [CrossRef] [PubMed]
  44. Bauzá, A.; Mooibroek, T.J.; Frontera, A. Tetrel-Bonding Interaction: Rediscovered Supramolecular Force? Angew. Chem. Int. Ed. 2013, 52, 12317–12321. [Google Scholar] [CrossRef] [PubMed]
  45. Bauzá, A.; Mooibroek, T.J.; Frontera, A. Tetrel Bonding Interactions. Chem. Rec. 2016, 16, 473–487. [Google Scholar] [CrossRef] [PubMed]
  46. Bene, J.E.D.; Alkorta, I.; Elguero, J. Carbenes as Electron-Pair Donors To CO2 for C···C Tetrel Bonds and C−C Covalent Bonds. J. Phys. Chem. A 2017, 121, 4039–4047. [Google Scholar] [CrossRef] [PubMed]
  47. Alkorta, I.; Elguero, J.; Bene, J.E.D. Azines as Electron-Pair Donors to CO2 for N···C Tetrel Bonds. J. Phys. Chem. A 2017, 121, 8017–8025. [Google Scholar] [CrossRef] [PubMed]
  48. Bene, J.E.D.; Alkorta, I.; Elguero, J. Carbon-Carbon Bonding between Nitrogen Heterocyclic Carbenes and CO2. J. Phys. Chem. A 2017, 121, 8136–8146. [Google Scholar] [CrossRef] [PubMed]
  49. Bene, J.D.; Elguero, J.; Alkorta, I. Complexes of CO2 with the Azoles: Tetrel Bonds, Hydrogen Bonds and Other Secondary Interactions. Molecules 2018, 23, 906. [Google Scholar] [CrossRef] [PubMed]
  50. Grabowski, S.J. Triel Bonds, π-Hole-π-Electrons Interactions in Complexes of Boron and Aluminium Trihalides and Trihydrides with Acetylene and Ethylene. Molecules 2015, 20, 11297–11316. [Google Scholar] [CrossRef] [PubMed]
  51. Bauzá, A.; Frontera, A. On the Versatility of BH2X (X=F, Cl, Br, and I) Compounds as Halogen-, Hydrogen-, and Triel-Bond Donors: An Ab Initio Study. ChemPhysChem 2016, 17, 3181–3186. [Google Scholar] [CrossRef] [PubMed]
  52. Esrafili, M.D.; Asadollahi, S.; Mousavian, P. Anionic tetrel bonds: An ab initio study. Chem. Phys. Lett. 2018, 691, 394–400. [Google Scholar] [CrossRef]
  53. Li, Q.Z.; Zhuo, H.Y.; Li, H.B.; Liu, Z.B.; Li, W.Z.; Cheng, J.B. Tetrel-Hydride Interaction between XH3F (X = C, Si, Ge, Sn) and HM (M = Li, Na, BeH, MgH). J. Phys. Chem. A 2014, 119, 2217–2224. [Google Scholar] [CrossRef] [PubMed]
  54. Lu, Y.; Wang, Y.; Zhu, W. Nonbonding interactions of organic halogens in biological systems: Implications for drug discovery and biomolecular design. Phys. Chem. Chem. Phys. 2010, 12, 4543–4551. [Google Scholar] [CrossRef] [PubMed]
  55. Mani, D.; Arunan, E. The X−C··· (X = F, Cl, Br, CN) Carbon Bond. J. Phys. Chem. A 2014, 118, 10081–10089. [Google Scholar] [CrossRef] [PubMed]
  56. Xu, H.; Cheng, J.; Yang, X.; Liu, Z.; Li, W.; Li, Q. Comparison of σ-Hole and π-Hole Tetrel Bonds Formed by Pyrazine and 1,4-Dicyanobenzene: The Interplay between Anion-π and Tetrel Bonds. ChemPhysChem 2017, 18, 2442–2450. [Google Scholar] [CrossRef] [PubMed]
  57. Grabowski, S.J. Tetrel bond- σ-hole bond as a preliminary stage of the SN2 reaction. Phys. Chem. Chem. Phys. 2014, 16, 1824–1834. [Google Scholar] [CrossRef] [PubMed]
  58. Murray, J.S.; Lane, P.; Politzer, P. Expansion of the σ-hole concept. J. Mol. Model. 2009, 15, 723–729. [Google Scholar] [CrossRef] [PubMed]
  59. Scilabra, P.; Kumar, V.; Ursini, M.; Resnati, G. Close contacts involving germanium and tin in crystal structures: Experimental evidence of tetrel bonds. J. Mol. Model. 2018, 24, 37. [Google Scholar] [CrossRef] [PubMed]
  60. Donald, K.J.; Tawfik, M. The Weak Helps the Strong: Sigma-Holes and the Stability of MF4·Base Complexes. J. Phys. Chem. A 2013, 117, 14176–14183. [Google Scholar] [CrossRef] [PubMed]
  61. Bundhun, A.; Ramasami, P.; Murray, J.S.; Politzer, P. Trends in σ-hole strengths and interactions of F3MX molecules (M = C, Si, Ge and X = F, Cl, Br, I). J. Mol. Model. 2013, 19, 2739–2746. [Google Scholar] [CrossRef] [PubMed]
  62. Scheiner, S. Comparison of CH···O, SH···O, Chalcogen, and Tetrel Bonds Formed by Neutral and Cationic Sulfur-Containing Compounds. J. Phys. Chem. A 2015, 119, 9189–9199. [Google Scholar] [CrossRef] [PubMed]
  63. Azofra, L.M.; Scheiner, S. Tetrel, chalcogen, and CH···O hydrogen bonds in complexes pairing carbonyl-containing molecules with 1, 2, and 3 molecules of CO2. J. Chem. Phys. 2015, 142, 034307. [Google Scholar] [CrossRef] [PubMed]
  64. Wei, Y.; Li, Q.; Scheiner, S. The π-Tetrel Bond and its Influence on Hydrogen Bonding and Proton Transfer. ChemPhysChem 2018, 19, 736–743. [Google Scholar] [CrossRef] [PubMed]
  65. McDowell, S.A.C.; Joseph, J.A. The effect of atomic ions on model σ-hole bonded complexes of AH3Y (A = C, Si, Ge; Y = F, Cl, Br). Phys. Chem. Chem. Phys. 2014, 16, 10854. [Google Scholar] [CrossRef] [PubMed]
  66. Politzer, P.; Murray, J.S.; Clark, T. Halogen bonding and other σ-hole interactions: A perspective. Phys. Chem. Chem. Phys. 2013, 15, 11178–11189. [Google Scholar] [CrossRef] [PubMed]
  67. Politzer, P.; Murray, J.S. σ-Hole Interactions: Perspectives and Misconceptions. Crystals 2017, 7, 212. [Google Scholar] [CrossRef]
  68. Politzer, P.; Murray, J.S.; Clark, T.; Resnati, G. The σ-hole revisited. Phys. Chem. Chem. Phys. 2017, 19, 32166–32178. [Google Scholar] [CrossRef] [PubMed]
  69. Clark, T.; Hennemann, M.; Murray, J.S.; Politzer, P. Halogen bonding: The σ-hole. J. Mol. Model. 2007, 13, 291–296. [Google Scholar] [CrossRef] [PubMed]
  70. Jönsson, B.; Karlström, G.; Wennerström, H. Ab initio molecular orbital calculations on the water-carbon dioxide system: Molecular complexes. Chem. Phys. Lett. 1975, 30, 58–59. [Google Scholar] [CrossRef]
  71. Peterson, K.I.; Klemperer, W. Structure and internal rotation of H2O−CO2, HDO−CO2, and D2O−CO2 van der Waals complexes. J. Chem. Phys. 1984, 80, 2439–2445. [Google Scholar] [CrossRef]
  72. Mitzel, N.W.; Blake, A.J.; Rankin, D.W.H. β-Donor Bonds in SiON Units: An Inherent Structure- Determining Property Leading to (4+4)-Coordination in Tetrakis-(N,N-dimethylhydroxylamido)silane. J. Am. Chem. Soc. 1997, 119, 4143–4148. [Google Scholar] [CrossRef]
  73. Mitzel, N.W.; Losehand, U. β-Donorbindungen in Molekülen mit SiON-Einheiten. Angew. Chem. 1997, 109, 2897–2899. [Google Scholar] [CrossRef]
  74. Southern, S.A.; Bryce, D.L. NMR Investigations of Noncovalent Carbon Tetrel Bonds. Computational Assessment and Initial Experimental Observation. J. Phys. Chem. A 2015, 119, 11891–11899. [Google Scholar] [CrossRef] [PubMed]
  75. Brammer, L. Halogen bonding, chalcogen bonding, pnictogen bonding, tetrel bonding: Origins, current status and discussion. Faraday Discuss. 2017, 203, 485–507. [Google Scholar] [CrossRef] [PubMed]
  76. Li, R.; Li, Q.; Cheng, J.; Liu, Z.; Li, W. The Prominent Enhancing Effect of the Cation-π Interaction on the Halogen-Hydride Halogen Bond in M1···C6H5X···HM2. ChemPhysChem 2011, 12, 2289–2295. [Google Scholar] [CrossRef] [PubMed]
  77. Li, Q.Z.; Sun, L.; Liu, X.F.; Li, W.Z.; Cheng, J.B.; Zeng, Y.L. Enhancement of Iodine-Hydride Interaction by Substitution and Cooperative Effects in NCX-NCI-HMY Complexes. ChemPhysChem 2012, 13, 3997–4002. [Google Scholar] [CrossRef] [PubMed]
  78. George, J.; Dronskowski, R. Tetrel Bonds in Infinite Molecular Chains by Electronic Structure Theory and Their Role for Crystal Stabilization. J. Phys. Chem. A 2017, 121, 1381–1387. [Google Scholar] [CrossRef] [PubMed]
  79. Tang, Q.; Li, Q. Interplay between tetrel bonding and hydrogen bonding interactions in complexes involving F2XO (X=C and Si) and HCN. Comput. Theor. Chem. 2014, 1050, 51–57. [Google Scholar] [CrossRef]
  80. Liu, M.; Li, Q.; Li, W.; Cheng, J.; McDowell, S.A.C. Comparison of hydrogen, halogen, and tetrel bonds in the complexes of HArF with YH3X (X = halogen, Y = C and Si). RSC Adv. 2016, 6, 19136–19143. [Google Scholar] [CrossRef]
  81. Guo, X.; Liu, Y.W.; Li, Q.Z.; Li, W.Z.; Cheng, J.B. Competition and cooperativity between tetrel bond and chalcogen bond in complexes involving F2CX (X = Se and Te). Chem. Phys. Lett. 2015, 620, 7–12. [Google Scholar] [CrossRef]
  82. Marín-Luna, M.; Alkorta, I.; Elguero, J. Cooperativity in Tetrel Bonds. J. Phys. Chem. A 2016, 120, 648–656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Solimannejad, M.; Orojloo, M.; Amani, S. Effect of cooperativity in lithium bonding on the strength of halogen bonding and tetrel bonding: (LiCN)···ClYF3 and (LiCN)···YF3Cl (Y= C, Si and n=1-5) complexes as a working model. J. Mol. Model. 2015, 21. [Google Scholar] [CrossRef] [PubMed]
  84. Esrafili, M.D.; Mohammadirad, N.; Solimannejad, M. Tetrel bond cooperativity in open-chain (CH3CN)n and (CH3NC)n clusters (n = 2–7): An ab initio study. Chem. Phys. Lett. 2015, 628, 16–20. [Google Scholar] [CrossRef]
  85. Liu, M.; Li, Q.; Li, W.; Cheng, J. Tetrel bonds between PySiX3 and some nitrogenated bases: Hybridization, substitution, and cooperativity. J. Mol. Graphics Modell. 2016, 65, 35–42. [Google Scholar] [CrossRef] [PubMed]
  86. Wei, Y.; Li, Q.; Yang, X.; McDowell, S.A.C. Intramolecular Si···O Tetrel Bonding: Tuning of Substituents and Cooperativity. ChemistrySelect 2017, 2, 11104–11112. [Google Scholar] [CrossRef]
  87. Wei, Y.; Cheng, J.; Li, W.; Li, Q. Regulation of coin metal substituents and cooperativity on the strength and nature of tetrel bonds. RSC Adv. 2017, 7, 46321–46328. [Google Scholar] [CrossRef] [Green Version]
  88. Mani, D.; Arunan, E. The X−C···Y Carbon Bond. In Noncovalent Forces; Springer International Publishing: Cham, Switzerland, 2015; pp. 323–356. [Google Scholar]
  89. Mahadevi, A.S.; Sastry, G.N. Cooperativity in Noncovalent Interactions. Chem. Rev. 2016, 116, 2775–2825. [Google Scholar] [CrossRef] [PubMed]
  90. Politzer, P.; Murray, J.S. The fundamental nature and role of the electrostatic potential in atoms and molecules. Theor. Chem. Acc. 2002, 108, 134–142. [Google Scholar] [CrossRef]
  91. Murray, J.S.; Politzer, P. The electrostatic potential: An overview. WIREs Comput. Mol. Sci. 2011, 1, 153–163. [Google Scholar] [CrossRef]
  92. Mani, D.; Arunan, E. The X−C···Y (X = O/F, Y = O/S/F/Cl/Br/N/P) ‘carbon bond’ and hydrophobic interactions. Phys. Chem. Chem. Phys. 2013, 15, 14377–14383. [Google Scholar] [CrossRef] [PubMed]
  93. Zierkiewicz, W.; Michalczyk, M.; Scheiner, S. Comparison between Tetrel Bonded Complexes Stabilized by σ and π Hole Interactions. Molecules 2018, 23, 1416. [Google Scholar] [CrossRef] [PubMed]
  94. Zierkiewicz, W.; Michalczyk, M.; Scheiner, S. Implications of monomer deformation for tetrel and pnicogen bonds. Phys. Chem. Chem. Phys. 2018, 20, 8832–8841. [Google Scholar] [CrossRef] [PubMed]
  95. Cremer, D.; Kraka, E. From Molecular Vibrations to Bonding, Chemical Reactions, and Reaction Mechanism. Curr. Org. Chem. 2010, 14, 1524–1560. [Google Scholar] [CrossRef]
  96. Kraka, E.; Setiawan, D.; Cremer, D. Re-evaluation of the Bond Length-Bond Strength Rule: The Stronger Bond Is not Always the Shorter Bond. J. Comp. Chem. 2016, 37, 130–142. [Google Scholar] [CrossRef] [PubMed]
  97. Setiawan, D.; Kraka, E.; Cremer, D. Hidden Bond Anomalies: The Peculiar Case of the Fluorinated Amine Chalcogenides. J. Phys. Chem. A 2015, 119, 9541–9556. [Google Scholar] [CrossRef] [PubMed]
  98. Setiawan, D.; Sethio, D.; Cremer, D.; Kraka, E. From strong to weak NF bonds: On the design of a new class of fluorinating agents. Phys. Chem. Chem. Phys. 2018, 20, 23913–23927. [Google Scholar] [CrossRef] [PubMed]
  99. Lin, C.Y.; Gilbert, A.T.B.; Gill, P.M.W. Calculating molecular vibrational spectra beyond the harmonic approximation. Theor. Chem. Acc. 2007, 120, 23–35. [Google Scholar] [CrossRef]
  100. Roy, T.K.; Gerber, R.B. Vibrational self-consistent field calculations for spectroscopy of biological molecules: New algorithmic developments and applications. Phys. Chem. Chem. Phys. 2013, 15, 9468–9492. [Google Scholar] [CrossRef] [PubMed]
  101. Panek, P.T.; Jacob, C.R. Anharmonic Theoretical Vibrational Spectroscopy of Polypeptides. J. Phys. Chem. Lett. 2016, 7, 3084–3090. [Google Scholar] [CrossRef] [PubMed]
  102. Konkoli, Z.; Cremer, D. A New Way of Analyzing Vibrational Spectra I. Derivation of Adiabatic Internal Modes. Int. J. Quant. Chem. 1998, 67, 1–9. [Google Scholar] [CrossRef]
  103. Konkoli, Z.; Larsson, J.A.; Cremer, D. A new way of analyzing vibrational spectra. IV. Application and testing of adiabatic modes within the concept of the characterization of normal modes. Int. J. Quant. Chem. 1998, 67, 41–55. [Google Scholar] [CrossRef]
  104. Zou, W.; Kalescky, R.; Kraka, E.; Cremer, D. Relating Normal Vibrational Modes to Local Vibrational Modes with the Help of an Adiabatic Connection Scheme. J. Chem. Phys. 2012, 137, 084114. [Google Scholar] [CrossRef] [PubMed]
  105. Kalescky, R.; Kraka, E.; Cremer, D. Description of Aromaticity with the Help of Vibrational Spectroscopy: Anthracene and Phenanthrene. J. Phys. Chem. A 2014, 118, 223–237. [Google Scholar] [CrossRef] [PubMed]
  106. Freindorf, M.; Tao, Y.; Sethio, D.; Cremer, D.; Kraka, E. New Mechanistic Insights into the Claisen Rearrangement of Chorismate—A Unified Reaction Valley Approach Study. Mol. Phys. 2018, in press. [Google Scholar] [CrossRef]
  107. Cremer, D.; Larsson, J.A.; Kraka, E. New developments in the analysis of vibrational spectra On the use of adiabatic internal vibrational modes. In Theoretical and Computational Chemistry; Parkanyi, C., Ed.; Elsevier: Amsterdam, The Netherlands, 1998; pp. 259–327. [Google Scholar]
  108. Zou, W.; Kalescky, R.; Kraka, E.; Cremer, D. Relating normal vibrational modes to local vibrational modes: Benzene and naphthalene. J. Mol. Model. 2012, 19, 2865–2877. [Google Scholar] [CrossRef] [PubMed]
  109. Humason, A.; Zou, W.; Cremer, D. 11,11-Dimethyl-1,6-methano[10]annulene—An Annulene with an Ultralong CC Bond or a Fluxional Molecule? J. Phys. Chem. A 2015, 119, 1666–1682. [Google Scholar] [CrossRef] [PubMed]
  110. Kalescky, R.; Kraka, E.; Cremer, D. Identification of the Strongest Bonds in Chemistry. J. Phys. Chem. A 2013, 117, 8981–8995. [Google Scholar] [CrossRef] [PubMed]
  111. Kalescky, R.; Kraka, E.; Cremer, D. New Approach to Tolman’s Electronic Parameter Based on Local Vibrational Modes. Inorg. Chem. 2014, 53, 478–495. [Google Scholar] [CrossRef] [PubMed]
  112. Kraka, E.; Cremer, D. Characterization of CF Bonds with Multiple-Bond Character: Bond Lengths, Stretching Force Constants, and Bond Dissociation Energies. ChemPhysChem 2009, 10, 686–698. [Google Scholar] [CrossRef] [PubMed]
  113. Kalescky, R.; Kraka, E.; Cremer, D. Are carbon-halogen double and triple bonds possible? Int. J. Quantum Chem. 2014, 114, 1060–1072. [Google Scholar] [CrossRef]
  114. Kalescky, R.; Zou, W.; Kraka, E.; Cremer, D. Quantitative Assessment of the Multiplicity of Carbon-Halogen Bonds: Carbenium and Halonium Ions with F, Cl, Br, and I. J. Phys. Chem. A 2014, 118, 1948–1963. [Google Scholar] [CrossRef] [PubMed]
  115. Oomens, J.; Kraka, E.; Nguyen, M.K.; Morton, T.H. Structure, Vibrational Spectra, and Unimolecular Dissociation of Gaseous 1-Fluoro-1-phenethyl Cations. J. Phys. Chem. A 2008, 112, 10774–10783. [Google Scholar] [CrossRef] [PubMed]
  116. Oliveira, V.; Kraka, E.; Cremer, D. Quantitative Assessment of Halogen Bonding Utilizing Vibrational Spectroscopy. Inorg. Chem. 2016, 56, 488–502. [Google Scholar] [CrossRef] [PubMed]
  117. Oliveira, V.; Kraka, E. Systematic Coupled Cluster Study of Noncovalent Interactions Involving Halogens, Chalcogens, and Pnicogens. J. Phys. Chem. A 2017, 121, 9544–9556. [Google Scholar] [CrossRef] [PubMed]
  118. Oliveira, V.; Cremer, D. Transition from metal-ligand bonding to halogen bonding involving a metal as halogen acceptor a study of Cu, Ag, Au, Pt, and Hg complexes. Chem. Phys. Lett. 2017, 681, 56–63. [Google Scholar] [CrossRef]
  119. Zhang, X.; Dai, H.; Yan, H.; Zou, W.; Cremer, D. B−H···π Interaction: A New Type of Nonclassical Hydrogen Bonding. J. Am. Chem. Soc. 2016, 138, 4334–4337. [Google Scholar] [CrossRef] [PubMed]
  120. Zou, W.; Zhang, X.; Dai, H.; Yan, H.; Cremer, D.; Kraka, E. Description of an unusual hydrogen bond between carborane and a phenyl group. J. Organ. Chem. 2018, 865, 114–127. [Google Scholar] [CrossRef]
  121. Purvis, G.D.; Bartlett, R.J. A full coupled-cluster singles and doubles model: The inclusion of disconnected triples. J. Chem. Phys. 1982, 76, 1910–1918. [Google Scholar] [CrossRef]
  122. Pople, J.A.; Head-Gordon, M.; Raghavachari, K. Quadratic configuration interaction. A general technique for determining electron correlation energies. J. Chem. Phys. 1987, 87, 5968–5975. [Google Scholar] [CrossRef]
  123. Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  124. Woon, D.E.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98, 1358–1371. [Google Scholar] [CrossRef]
  125. Woon, D.E.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. IV. Calculation of static electrical response properties. J. Chem. Phys. 1994, 100, 2975–2988. [Google Scholar] [CrossRef]
  126. Wilson, E.B.; Decius, J.C.; Cross, P.C. Molecular Vibrations. The Theory of Infrared and Raman Vibrational Spectra; McGraw-Hill: New York, NY, USA, 1955. [Google Scholar]
  127. Kraka, E.; Larsson, J.A.; Cremer, D. Generalization of the Badger Rule Based on the Use of Adiabatic Vibrational Modes. In Computational Spectroscopy; Grunenberg, J., Ed.; Wiley: New York, NY, USA, 2010; pp. 105–149. [Google Scholar]
  128. Boys, S.F.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  129. Mentel, ŁM.; Baerends, E.J. Can the Counterpoise Correction for Basis Set Superposition Effect Be Justified? J. Chem. Theory Comput. 2013, 10, 252–267. [Google Scholar] [CrossRef] [PubMed]
  130. Riplinger, C.; Neese, F. An efficient and near linear scaling pair natural orbital based local coupled cluster method. J. Chem. Phys. 2013, 138, 034106. [Google Scholar] [CrossRef] [PubMed]
  131. Riplinger, C.; Sandhoefer, B.; Hansen, A.; Neese, F. Natural triple excitations in local coupled cluster calculations with pair natural orbitals. J. Chem. Phys. 2013, 139, 134101. [Google Scholar] [CrossRef] [PubMed]
  132. Stanton, J.F.; Gauss, J.; Cheng, L.; Harding, M.E.; Matthews, D.A.; Szalay, P.G. CFOUR, Coupled-Cluster techniques for Computational Chemistry, a Quantum-Chemical Program Package. Available online: http://www.cfour.de (accessed on 1 October 2018).
  133. Harding, M.; Mezroth, T.; Gauss, J.; Auer, A. Parallel Calculation of CCSD and CCSD(T) Analytic First and Second Derivatives. J. Chem. Theory Comput. 2008, 4, 64–74. [Google Scholar] [CrossRef] [PubMed]
  134. Neese, F. The ORCA program system. WIREs Comput. Mol. Sci. 2011, 2, 73–78. [Google Scholar] [CrossRef]
  135. Weinhold, F.; Landis, C.R. Valency and Bonding: A Natural Bond Orbital Donor-Acceptor Perspective; Cambridge University Press: Cambridge, UK, 2003. [Google Scholar]
  136. Reed, A.; Curtiss, L.; Weinhold, F. Intermolecular Interactions from A Natural Bond Orbital, Donor-Acceptor Viewpoint. Chem. Rev. 1988, 88, 899–926. [Google Scholar] [CrossRef]
  137. Glendening, E.D.; Badenhoop, J.K.; Reed, A.E.; Carpenter, J.E.; Bohmann, J.A.; Morales, C.M.; Landis, C.R.; Weinhold, F. NBO6. In Theoretical Chemistry Institute; University of Wisconsin: Madison, WI, USA, 2013. [Google Scholar]
  138. Glendening, E.D.; Landis, C.R.; Weinhold, F. NBO 6.0: Natural bond orbital analysis program. J. Comput. Chem. 2013, 34, 1429–1437. [Google Scholar] [CrossRef] [PubMed]
  139. Keith, T. TK Gristmill Software. Overland Park, KS, USA. Available online: http//aim.tkgristmill.com (accessed on 1 October 2018).
  140. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2011, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  141. 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] [PubMed] [Green Version]
  142. Zou, W.; Nori-Shargh, D.; Boggs, J.E. On the Covalent Character of Rare Gas Bonding Interactions: A New Kind of Weak Interaction. J. Phys. Chem. A 2012, 117, 207–212. [Google Scholar] [CrossRef] [PubMed]
  143. Kraka, E.; Zou, W.; Filatov, M.; Gräfenstein, J.; Izotov, D.; Gauss, J.; He, Y.; Wu, A.; Konkoli, Z.; Polo, V.; et al. COLOGNE. 2018. Available online: http://www.smu.edu/catco (accessed on 1 October 2018).
  144. Li, Y.; Oliveira, V.; Tang, C.; Cremer, D.; Liu, C.; Ma, J. The Peculiar Role of the Au3 Unit in Aum Clusters: σ-Aromaticity of the Au5Zn+ Ion. Inorg. Chem. 2017, 56, 5793–5803. [Google Scholar] [CrossRef] [PubMed]
  145. 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. 1984, 23, 627–628. [Google Scholar] [CrossRef]
  146. Cremer, D.; Kraka, E. A Description of the Chemical Bond in Terms of Local Properties of Electron Density and Energy. Croatica Chem. Acta 1984, 57, 1259–1281. [Google Scholar]
  147. Grabowski, S. Lewis Acid Properties of Tetrel Tetrafluorides—The Coincidence of the σ-Hole Concept with the QTAIM Approach. Crystals 2017, 7, 43. [Google Scholar] [CrossRef]
  148. Scheiner, S. Systematic Elucidation of Factors That Influence the Strength of Tetrel Bonds. J. Phys. Chem. A 2017, 121, 5561–5568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Scheiner, S. Steric Crowding in Tetrel Bonds. J. Phys. Chem. A 2018, 122, 2550–2562. [Google Scholar] [CrossRef] [PubMed]
  150. Angarov, V.; Kozuch, S. On the σ, π and δ hole interactions: A molecular orbital overview. New J. Chem. 2018, 42, 1413–1422. [Google Scholar] [CrossRef]
  151. Wang, H.; Wang, W.; Jin, W.J. σ-Hole Bond vs π-Hole Bond: A Comparison Based on Halogen Bond. Chem. Rev. 2016, 116, 5072–5104. [Google Scholar] [CrossRef] [PubMed]
  152. Dong, W.; Wang, Y.; Cheng, J.; Yang, X.; Li, Q. Competition between σ-hole pnicogen bond and π-hole tetrel bond in complexes of CF2=CFZH2 (Z= P, As, and Sb). Mol. Phys. 2018, 1–9. [Google Scholar] [CrossRef]
  153. Grabowski, S.J. Hydrogen bonds, and σ-hole and π-hole bonds—Mechanisms protecting doublet and octet electron structures. Phys. Chem. Chem. Phys. 2017, 19, 29742–29759. [Google Scholar] [CrossRef] [PubMed]
  154. Shen, S.; Zeng, Y.; Li, X.; Meng, L.; Zhang, X. Insight into the π-holebond···π-electrons tetrel bonds between F2ZO (Z = C, Si, Ge) and unsaturated hydrocarbons. Int. J. Quantum Chem. 2017, 118, e25521. [Google Scholar] [CrossRef]
  155. Liu, M.; Li, Q.; Li, W.; Cheng, J. Carbene tetrel-bonded complexes. Struct. Chem. 2017, 28, 823–831. [Google Scholar] [CrossRef]
  156. Xu, H.; Cheng, J.; Yu, X.; Li, Q. Abnormal Tetrel Bonds between Formamidine and TH3F: Substituent Effects. Chem. Sel. 2018, 3, 2842–2849. [Google Scholar]
  157. Zierkiewicz, W.; Michalczyk, M. On the opposite trends of correlations between interaction energies and electrostatic potentials of chlorinated and methylated amine complexes stabilized by halogen bond. Theor. Chem. Acc. 2017, 136. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic representation of tetrel complexes between the electron-deficient tetrel atom of a Lewis acid (tetrel donor, T-donor, T = C, Si, Ge) and the electron-rich tetrel atom of a Lewis base (tetrel acceptor, T-acceptor, A = FH, OH2, NH3, Cl).
Figure 1. Schematic representation of tetrel complexes between the electron-deficient tetrel atom of a Lewis acid (tetrel donor, T-donor, T = C, Si, Ge) and the electron-rich tetrel atom of a Lewis base (tetrel acceptor, T-acceptor, A = FH, OH2, NH3, Cl).
Molecules 23 02763 g001
Figure 2. Schematic representation of complexes 1-35 with atomic charges (in me) from the natural population analysis calculated at the CCSD(T)/aug-cc-pVTZ level of theory. Colors are used to correlate charges to specific atoms.
Figure 2. Schematic representation of complexes 1-35 with atomic charges (in me) from the natural population analysis calculated at the CCSD(T)/aug-cc-pVTZ level of theory. Colors are used to correlate charges to specific atoms.
Molecules 23 02763 g002
Figure 3. Molecular electrostatic potential of neutral tetrel-donors mapped onto the 0.001 a.u electron density surface. Blue and red correspond, respectively, to the positive and negative potential. The extreme values are ±1.9 eV. The V s m a x at the tetrel σ or π -hole are given in bold blue, while the V s m a x at the H (36, 37, 38) and at the chalcogen atoms (55, 57) are shown in black. Calculated at the CCSD(T)/aug-cc-pVTZ level of theory.
Figure 3. Molecular electrostatic potential of neutral tetrel-donors mapped onto the 0.001 a.u electron density surface. Blue and red correspond, respectively, to the positive and negative potential. The extreme values are ±1.9 eV. The V s m a x at the tetrel σ or π -hole are given in bold blue, while the V s m a x at the H (36, 37, 38) and at the chalcogen atoms (55, 57) are shown in black. Calculated at the CCSD(T)/aug-cc-pVTZ level of theory.
Molecules 23 02763 g003
Figure 4. Power relationship between the relative bond strength order (BSO) n and the local stretching force constants k a of the TA interaction of complexes 1–35. C donors are gray, Si donors are blue, Ge donors are purple, double bonded donors are green, and charge-assisted TBs are orange. Complex 28 is not shown. Calculated at the CCSD(T)/aug-cc-pVTZ level of theory.
Figure 4. Power relationship between the relative bond strength order (BSO) n and the local stretching force constants k a of the TA interaction of complexes 1–35. C donors are gray, Si donors are blue, Ge donors are purple, double bonded donors are green, and charge-assisted TBs are orange. Complex 28 is not shown. Calculated at the CCSD(T)/aug-cc-pVTZ level of theory.
Molecules 23 02763 g004
Figure 5. Comparison of the relative bond strength order (BSO) n and the energy density at the bond critical point H b of the tetrel bond of complexes. C tetrel bonds are gray, Si tetrel bonds are blue, Ge tetrel bonds are purple, double bonded tetrel bonds are green, and anionic tetrel bonds are orange. Complexes 26b–28 and 35 are not shown. Calculated at the CCSD(T)/aug-cc-pVTZ level of theory.
Figure 5. Comparison of the relative bond strength order (BSO) n and the energy density at the bond critical point H b of the tetrel bond of complexes. C tetrel bonds are gray, Si tetrel bonds are blue, Ge tetrel bonds are purple, double bonded tetrel bonds are green, and anionic tetrel bonds are orange. Complexes 26b–28 and 35 are not shown. Calculated at the CCSD(T)/aug-cc-pVTZ level of theory.
Molecules 23 02763 g005
Figure 6. Relationship between the binding energy and interactomic distance computed at the CCSD(T)/aug-cc-pVTZ level of theory. All geometric parameters were optimized at each point of the curves for fixed r(TA) values. The blue dots represent the binding energy at the minima of complex 26 and the minimum of 35; the black lines connecting points were used to improve interpretation.
Figure 6. Relationship between the binding energy and interactomic distance computed at the CCSD(T)/aug-cc-pVTZ level of theory. All geometric parameters were optimized at each point of the curves for fixed r(TA) values. The blue dots represent the binding energy at the minima of complex 26 and the minimum of 35; the black lines connecting points were used to improve interpretation.
Molecules 23 02763 g006
Table 1. Summary of energetics, electron density, energy density, geometric, bond strength order, and vibrational spectroscopy data for complexes 1–35 *.
Table 1. Summary of energetics, electron density, energy density, geometric, bond strength order, and vibrational spectroscopy data for complexes 1–35 *.
#Complex (symm.) Δ E Δ E cp E def rrCT ρ b H b k a n k a n
TAXT TATATATAXTXT
Neutral tetrel bonds involving C donors
1FCH3···FH (C3v)−1.50−1.290.012.9721.39220.0340.0120.0450.0735.0181.038
2FCH3···OH2 (Cs)−2.10−1.870.023.0351.39430.0410.0120.0550.0814.9561.030
3FCH3···NH3 (C3v)−2.25−2.050.023.2181.39550.0400.0090.0490.0764.9121.025
4ClCH3···NH3 (C3v)−2.08−1.880.023.2891.79860.0370.0080.0430.0712.9430.768
5BrCH3···NH3 (C3v)−2.00−1.800.023.3041.95360.0370.0080.0410.0692.5150.703
6(HO)CH3···NH3 (Cs)−1.38−1.210.013.3621.42930.0310.0080.0320.0604.6520.994
7CF4···NH3 (C3v)−1.62−1.240.063.4261.32810.030 a 0.007 a 0.0440.0725.9261.140
Neutral tetrel bonds involving Si donors
8FSiH3···FH (Cs)−2.28−1.850.062.9641.61790.0550.0050.0620.0874.9701.032
9FSiH3···OH2 (Cs)−4.20−3.610.352.7741.623250.0920.0020.0880.1064.7621.007
10FSiH3···NH3 (C3v)−6.80−5.942.112.5231.637810.179−0.0330.1030.1164.2090.940
11ClSiH3···NH3 (C3v)−6.13−5.412.022.5802.117840.165−0.0240.0730.0951.9410.607
12BrSiH3···NH3 (C3v)−6.11−5.352.232.5662.290900.170−0.0270.0660.0901.5050.526
13(HO)SiH3···NH3 (Cs)−4.13−3.610.682.8251.680420.108−0.0030.0700.0934.0650.921
14SiH4···NH3 (C3v)−2.27−1.970.153.2021.490180.0600.0040.0490.0762.7930.746
15SiF2H2···NH3 (Cs)−6.99−5.734.742.4001.613950.225−0.0660.0830.1034.5730.985
16aSiF3H···NH3 (Cs)−7.66−5.7711.772.2051.6171390.320−0.1260.2490.1914.6981.000
16bHSiF3···NH3 (C3v)−6.30−4.1421.222.1041.4741720.390−0.1490.4930.2802.9740.772
17SiF4···NH3 (C3v)−11.40−8.8621.152.0721.6091760.419−0.1640.6780.3355.0461.041
Neutral tetrel bonds involving Ge donors
18FGeH3···NH3 (C3v)−7.77−7.181.402.6241.816440.169−0.0080.1490.1434.1250.929
19ClGeH3···NH3 (C3v)−6.22−5.751.072.7552.216640.134−0.0010.1030.1161.9210.604
20BrGeH3···NH3 (C3v)−6.01−5.531.072.7762.375660.1320.0000.0970.1121.5910.543
21(HO)GeH3···NH3 (Cs)−4.58−4.180.502.9101.818390.1010.0040.0890.1073.490.845
22GeH4···NH3 (C3v)−1.99−1.790.093.3231.550150.0520.0050.0470.0742.5800.713
Neutral tetrel bonds involving double bonded C or Si donors
23CO2···NH3 (Cs)−3.09−2.840.112.9221.16750.1070.0020.0790.10015.1831.938
24SCO···NH3 (Cs)−1.97−1.690.023.2091.57330.0460.0090.0470.0747.0811.260
25CF2O···NH3 (Cs)−5.55−4.820.272.6871.178120.1130.0050.1220.12714.3931.880
26aCF2S···NH3 (Cs)−3.91−3.230.112.8971.60790.0780.0080.0860.1056.3971.190
26bCF2S···NH3 (Cs)1.454.2824.131.5871.7015451.388−1.3391.4140.5083.8280.891
27SiF2O···NH3 (Cs)−44.14−42.167.961.9171.5292290.569−0.2241.8380.5898.8031.425
Charge-assisted interactions
28CH3+···NH3 (C3v)−110.25 b −109.0124.951.5111.0873291.517−1.9523.7660.8825.4581.088
29FNH3+···NH3 (C3v)−23.14−22.770.432.6191.374350.1420.0120.3640.2365.2261.062
30FCH3···Cl (C3v)−9.77−9.340.393.1791.419230.0640.0100.1280.1314.1550.933
31FSiH3···Cl (C3v)−20.73−19.4912.032.5041.7032630.277−0.1150.3700.2382.7930.746
32FGeH3···Cl (C3v)−26.10−25.0910.712.5661.8922380.290−0.0690.4550.2682.4510.693
33CO2···Cl (Cs)−7.45−6.991.442.9201.170310.1070.0020.1090.12014.8791.916
34SCO···Cl (Cs)−5.36−4.960.523.1431.581240.0730.0060.0790.1006.5681.208
35CF2S···Cl (Cs)−16.81−13.8332.631.8981.7257981.031−0.5931.1000.4413.4140.835
* Binding energies (ΔE), counterpoise corrected binding energies ΔEcp and monomers’ deformation energies upon complexation (Edef) in kcal/mol. XT bond distance r(XT) and tetrel bond distance r(TA) in Å. Density at the TA critical point ρb in e/Å3, energy density at the TA critical point Hb in Hartree/Å3. Natural population analysis (NPA) charge transfer in mili-electrons (me). TA and XT local stretching force constant (ka) in mdyn/Å and bond strength order (BSO) n values. Computed at the CCSD(T)/aug-cc-pVTZ level of theory. a Calculated at a cage critical point (see Ref. [147]). b Covalent bond, see text.
Table 2. Geometry, vibrational spectroscopy data, and values of the electrostatic potential for the monomers *.
Table 2. Geometry, vibrational spectroscopy data, and values of the electrostatic potential for the monomers *.
#Monomers V s max (X)r(XT) k a (XT)n(XT)Dipole α iso
36F−CH30.901.3895.1071.0481.882.5
37Cl−CH30.761.7923.0680.7861.924.3
38Br−CH30.691.9482.6160.7181.865.4
39HO−CH30.471.4264.7491.006 3.1
40F−CF30.931.3216.2041.1700.002.8
41F−SiH31.641.6135.1201.0491.384.1
42Cl−SiH31.512.0722.7990.7461.416.2
43Br−SiH31.462.2382.3210.6721.387.4
44HO−SiH31.141.6644.5170.978 4.9
45H−SiH30.811.4832.9030.7620.004.6
46F−SiH2F1.721.5975.4971.092 3.5
47aF−SiF2H1.781.5835.8841.135 3.8
47bH−SiF31.441.4583.2730.8151.433.8
48F−SiF31.981.5716.2811.1780.003.3
49F−GeH31.931.7934.9511.0302.254.7
50Cl−GeH31.662.1752.4910.6992.046.9
51Br−GeH31.572.3302.0910.6331.938.1
52HO−GeH31.331.8023.8720.896 5.5
53H−GeH30.731.5422.6930.7300.005.2
54O=CO1.181.16715.6131.9690.002.6
55S=CO0.641.5757.2271.2750.685.2
56O=CF21.851.17714.6801.9021.002.8
57S=CF21.291.6036.6261.2140.165.2
58O=SiF23.461.5179.2431.4652.314.0
59CH3+10.01 0.001.3
60F−NH3+8.581.3685.6421.1094.781.7
* Maximum electrostatic potential at the σ -hole of X ( V s m a x (X)) in eV. XT bond distance r(XT) in Å, XT local stretching force k a ( X T ) in mdyn/Å , XT bond strength order n(XT). Dipole moment in Debye and static isotropic polarizability in Å 3 . All values were calculated with CCSD(T)/aug-cc-pVTZ.
Table 3. Naural Bond Orbital (NBO) electron delocalization energies involving the lone pair of NH3 *.
Table 3. Naural Bond Orbital (NBO) electron delocalization energies involving the lone pair of NH3 *.
#Complex σ (X-Si) σ (Si-R) σ (Si-R’) σ (Si-R”)
10FSiH3···NH315.72.42.42.4
15SiF2H2···NH312.76.23.13.1
16aSiF3H···NH316.311.911.97.5
16bSiF3H···NH311.423.523.523.5
16b a SiHF3···NH37.916.816.816.8
17SiF4···NH320.719.519.519.5
* NBO electron delocalization energies from the second-order perturbation analysis referent to the interaction involving the lp(N) orbital of NH3 and the σ (X-Si) (collinear to the TB), the σ (Si-R), σ (Si-R’) and the σ (Si-R”) (peripheral to the TB, see Figure 1) of selected tetrel donors (see Supporting Information Figure S5). Values are in kcal/mol. Calculated with ω B97XD/aug-cc-pVTZ. a Complex 16b with an elongated tetrel bond (TB) to match the TB distance of 16a.
Table 4. Summary of energetics, geometric and vibrational spectroscopy data for other types of interactions *.
Table 4. Summary of energetics, geometric and vibrational spectroscopy data for other types of interactions *.
Complex Δ E Δ E cp rCT ρ H b k a n k a n
TA TATATATAXTXT
F2···OH2 (Cs)−1.42−1.152.6620.0050.0660.0220.0570.0834.4880.974
Cl2···OH2 (Cs)−2.98−2.622.8080.0150.0980.0180.0970.1122.8960.761
FCl···OH2 (Cs)−5.22−4.752.5660.0320.1630.0160.1700.1543.9670.909
FSH···OH2 (Cs)−5.69−5.152.6590.0280.1380.0100.1520.1444.0110.914
FPH2···OH2 (Cs)−4.63−4.022.7800.0210.1070.0060.1180.1254.1980.938
F2···NH3 (C3v)−2.00−1.692.6150.0170.0970.0270.0620.0873.8210.890
Cl2···NH3 (C3v)−4.92−4.432.6640.0550.1720.0060.1320.1332.3700.680
FCl···NH3 (C3v)−10.13−9.392.3200.1450.358−0.0580.3110.2162.6870.729
FSH···NH3 (Cs)−8.23−7.582.5120.0810.235−0.0200.1940.1663.3090.820
FPH2···NH3 (Cs)−6.81−6.102.6630.0570.171−0.0120.1440.1403.7940.886
FCl···Cl (Cv)−30.07−28.982.3160.4960.547−0.1610.8550.3821.2120.465
FSH···Cl (Cs)−23.46−22.482.4930.3050.377−0.0920.4430.2641.4660.518
FPH2···Cl (Cs)−19.62−18.622.6490.2080.266−0.0580.3070.2142.1360.641
* Binding energies ( Δ E) and conterpoise corrected binding energies Δ E c p in kcal/mol. Intermonomer bond distance r(TA) in Å. Density at the TA critical point ρ b in e/Å3, energy density at the TA critical point H b in Hartree/Å3. NPA charge transfer (CT) in e. TA and XT local stretching force constant ( k a ) in mdyn/Å and BSO n values. Computed at the CCSD(T)/aug-cc-pVTZ level of theory.

Share and Cite

MDPI and ACS Style

Sethio, D.; Oliveira, V.; Kraka, E. Quantitative Assessment of Tetrel Bonding Utilizing Vibrational Spectroscopy. Molecules 2018, 23, 2763. https://doi.org/10.3390/molecules23112763

AMA Style

Sethio D, Oliveira V, Kraka E. Quantitative Assessment of Tetrel Bonding Utilizing Vibrational Spectroscopy. Molecules. 2018; 23(11):2763. https://doi.org/10.3390/molecules23112763

Chicago/Turabian Style

Sethio, Daniel, Vytor Oliveira, and Elfi Kraka. 2018. "Quantitative Assessment of Tetrel Bonding Utilizing Vibrational Spectroscopy" Molecules 23, no. 11: 2763. https://doi.org/10.3390/molecules23112763

Article Metrics

Back to TopTop