Next Article in Journal
The Phase Transition and Dehydration in Epsomite under High Temperature and High Pressure
Next Article in Special Issue
3D Numerical Analysis of the Asymmetric Three-Phase Line of Floating Zone for Silicon Crystal Growth
Previous Article in Journal
Improvement on Thermal Stability of Nano-Domains in Lithium Niobate Thin Films
Previous Article in Special Issue
Recent Advancements in Crystalline Pb-Free Halide Double Perovskites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interaction and Polarization Energy Relationships in σ-Hole and π-Hole Bonding

Department of Chemistry, University of New Orleans, New Orleans, LA 70148, USA
*
Author to whom correspondence should be addressed.
Crystals 2020, 10(2), 76; https://doi.org/10.3390/cryst10020076
Submission received: 4 January 2020 / Revised: 19 January 2020 / Accepted: 25 January 2020 / Published: 30 January 2020

Abstract

:
We demonstrate that a wide range of σ- and π-hole interaction energies can be related to (a) the electrostatic potentials and electric fields of the σ- and π-hole molecules at the approximate positions of the negative sites and (b) the electrostatic potentials and polarizabilities of the latter. This is consistent with the Coulombic nature of these interactions, which should be understood to include both electrostatics and polarization. The energies associated with polarization were estimated and were shown to overall be greater for the stronger interactions; no new factors need be introduced to account for these. All of the interactions can be treated in the same manner.

1. σ-Holes and π-Holes

The electronic density of a free atom is, on average, spherically symmetrical [1]. However, in forming a covalent bond and becoming part of a molecule, the atom normally loses this symmetry and becomes anisotropic, with regions of higher and lower electronic density [2,3,4,5,6,7,8,9].
There tends to be a lower electronic density on the side of the atom opposite to the bond, on its extension; the atom’s radius in that direction is less than to a lateral side. This region of lower electronic density on the extension of a bond is known as a σ-hole [10]. There is also frequently a region of lower electronic density above and below a planar portion of a molecule, which has been labeled a π-hole [11].
Our present interest is in the effects of σ-holes and π-holes upon the energetics of intermolecular interactions. Since these effects involve molecular electrostatic potentials and electric fields, we will briefly discuss these two properties.

2. Molecular Electrostatic Potentials and Electric Fields

The nuclei and electrons of any molecule create an electrostatic potential V(r) at each point r in the surrounding space. V(r) is given rigorously by Equation (1),
V ( r ) = A Z A | R A r | ρ ( r ) d r | r r |
in which ZA is the charge on nucleus A, located at RA, and ρ(r) is the molecule’s electronic density.
The sign of V(r) in any region is the result of the positive contributions of the nuclei and the negative ones of the electrons. It is sometimes incorrectly assumed that the electrostatic potential follows the electronic density and that “electron-rich” and “electron-poor” regions will have negative and positive V(r), respectively. This is sometimes the case but not always; the nuclear contributions must be taken into account as well as the electronic ones. For instance, this is shown by there usually being buildups of electronic density in the internuclear regions of covalent bonds [12,13,14], but the electrostatic potentials in those regions are generally positive [15,16,17], which is due to the proximities of the positive nuclei.
A key feature of the electrostatic potential is that it is a real physical property, an observable. It can be determined experimentally, by diffraction methods [14,15,18], as well as computationally.
The significance of V(r) for the present discussion is in relation to interaction energies. If a molecule A has an electrostatic potential VA(r), then its interaction energy ΔE with a point charge Q placed at the position R is ΔE = QVA(R). If Q and VA(R) have opposite signs, then ΔE < 0 and the interaction is attractive; if they have the same sign, ΔE > 0 and the interaction is repulsive. This can readily be extended to the interaction of A with a second molecule B, since B is simply a collection of point charges (nuclei and electrons). Regions of positive (negative) electrostatic potential on A will interact attractively with negative (positive) regions on B.
It might be argued that atomic charges can achieve the same purpose as electrostatic potentials. However, the charge on an atom in a molecule is not a physical observable; it cannot be determined experimentally. Numerous definitions of atomic charge have been proposed, but they are all arbitrary and have no physical basis [19,20,21,22,23]. The perils of atomic charges are illustrated by the case of H3C-NO2; seven different definitions of atomic charge assigned values to the carbon that ranged from −0.478 to +0.564 [20]. They do not even agree as to whether it is positive or negative!
In the context of molecular interactions, electrostatic potentials are now commonly presented on molecular surfaces defined by outer contours of the electronic densities [24] and labeled VS(r). The 0.001 au contour is most often used for this purpose. The local maxima and minima of VS(r) (its most positive and most negative values) are designated by VS,max and VS,min, respectively. There may be several of each.
The electrostatic potential of a molecule creates an electric field ε(r). From classical physics, this is equal to the negative gradient of the electrostatic potential, ε ( r ) = V ( r ) . If a charge Q is placed at the point R, then the electric field exerts a force F(R) upon Q, which is given by F ( R ) = Q ε ( R ) = Q V ( R ) . If V ( R ) and Q have the same sign, then the electric field attracts Q toward the molecule.

3. σ- and π-Hole Interactions

There are frequently (although not always) positive electrostatic potentials associated with the lower electronic densities of σ-holes and π-holes. Through these positive potentials, the molecule can interact attractively with negative sites such as lone pairs, π electrons, and anions [11,25,26,27]. These are known as σ- and π-hole interactions, even though they do not actually involve the σ- and π-holes but rather the positive potentials associated with them [28,29]. Such interactions have been widely observed, both experimentally and computationally, for covalently bonded atoms from different groups of the periodic table [2,11,25,26,30,31,32,33,34,35,36,37]. Hydrogen bonding is also a σ-hole interaction [38,39,40,41,42].
As examples of positive potentials associated with σ- and π-holes, Figure 1 and Figure 2 display the computed electrostatic potentials on the 0.001 au molecular surfaces of SeClBr and SeO3. These were computed at the density functional B3PW91/6-31G(d,p) level using Gaussian 09 [43] and the WFA-SAS code [44].
In SeClBr, Figure 1, there are four positive maxima, arising from two σ-holes on the selenium, and one on each halogen atom, which are on the extensions of the bonds to these atoms. Note that all three atoms have regions of both positive and negative electrostatic potential, which means that they can interact attractively with both negative and positive sites. Such dual reactivity is well-established experimentally [4,6,34,35] and computationally [45,46]. It could not have been predicted on the basis of atomic charges, since each atom in a molecule is assigned a single charge.
Figure 2 shows the positive potential that is above and below the selenium in SeO3, due to its π-holes. The three oxygens are entirely negative. They do have σ-holes (regions of lower electronic density) on the extensions of the Se–O bonds, but the potentials associated with them are negative, although less so than the surrounding oxygen surfaces.

4. Polarization

The electrostatic potentials plotted on molecular surfaces, as in Figure 1 and Figure 2, can reveal positive regions that are likely to interact favorably with negative sites. However, these potentials are typically computed for the free molecules prior to interaction. They do not reflect that as two molecules approach each other and interact, the charge distribution of each becomes increasingly polarized by the electric field of the other. This is clearly visible in density difference plots [47,48,49,50]; the electronic densities of the negative sites are polarized toward the positive potentials of the σ- or π-hole atoms, and the electronic densities of those atoms are polarized away from the negative sites. It follows that the electrostatic potentials computed for the free molecules, e.g., their VS,max and VS,min, become somewhat less relevant as the molecules interact.
Polarization is an intrinsic part of a Coulombic interaction [48,49,51,52,53], and it is always stabilizing, strengthening the interaction. Hennemann et al. have shown how putting a point charge in the vicinity of a σ-hole can affect the magnitude of the electrostatic potential associated with that σ-hole [39]. Polarization can actually result in an interaction becoming favorable even though the electrostatic potentials of the free molecules would predict it to be unfavorable. For instance, the σ-hole potentials of the chlorine in H3C–Cl and the phosphorus in H3P are essentially neutral (near-zero or even slightly negative, depending upon the computational procedure). Yet, these molecules were found to computationally to interact favorably with O=CH2 and NSH to form H3C-Cl---O=CH2 [54] and H3P---NSH [55]. The electric fields of the negative sites induce positive potentials on the chlorine and phosphorus.
Substituted aromatic systems provide some dramatic examples of the effects of polarization. 1,4-Difluorobenzene has weakly negative electrostatic potentials above and below the ring; nevertheless, it interacts there with the nitrogen lone pair of HCN to form a T-shaped complex [56]. The negative lone pair induces a positive ring potential. 1,3,5-Trifluorobenzene has weakly positive potentials above and below the ring, but it forms a T-shaped complex with the positive hydrogen of HCN, which induces a negative potential. See also Geronimo et al. [57].
The important role that polarization can play has been recognized since the earliest days of using electrostatic potentials to analyze molecular interactions [58,59,60]. Several approaches have been used to account for it, including perturbation theory [58,60,61] and modeling with point charges [59,62,63,64]. For overviews of this work, see Politzer et al. [17,65].

5. Strengths of σ- and π-Hole Interactions

The strength of noncovalent bonding is typically measured by the interaction energy, ΔE. For the formation of a complex A---B, this is given in terms of the respective energies by
ΔE = E(A---B) − [E(A) + E(B)].
The more negative ΔE is, the stronger the interaction.
ΔE reflects the contributions of both electrostatics and polarization. Sometimes, the polarization is relatively minor, and ΔE can be related to just electrostatics. Thus, for series of σ-hole interactions with a given negative site, it has on occasion been possible to correlate ΔE quite well with just the VS,max of the σ-hole molecules [25,30,66,67,68]. If the negative site is not held constant, there has been some success in relating ΔE to combinations of the VS,max of the σ-hole molecules and the VS,min of the negative sites [25,42,52]. For groups of interactions with a given σ- or π-hole molecule, correlations have sometimes been found between ΔE and the VS,min of the negative sites [59,68].
However, polarization does often have a significant effect and cannot be neglected. It has two components in σ- and π-hole interactions: namely, the effect of the electric field of the σ- or π-hole molecule upon the negative site and the effect of the electric field of the negative site upon the σ- or π-hole molecule.
We have recently been investigating possible procedures for expressing σ- and π-hole interaction energies in terms of both electrostatics and polarization. Initially, we used data bases of 20 σ-hole and 21 π-hole interactions, involving a variety of σ- and π-hole molecules and negative sites [69]. We verified that electrostatics—i.e., VS,max or a combination of VS,max and VS,min—is not sufficient for an acceptable correlation with either the σ-hole or the π-hole interaction energies. Polarization needs to be included.
Accordingly, we tested the triple regression relationship given by Equation (3). It represents the electrostatics of the interactions by the most positive electrostatic potentials, the VS,max, of the σ- or π-holes on the 0.001 au molecular surfaces. Polarization is taken into account by explicitly including the electric fields ε(R) that the σ- or π-hole molecules exert upon the negative sites and the average polarizabilities α of these sites [65,69].
ΔE = c1[VS,max] + c2[ε(R)]2 + c3[α] + c4
In Equation (3), R is the intermolecular separation. For a σ-hole molecule, ε(R) is evaluated at the distance R along the extension of the bond that produced the σ-hole, and for a π-hole molecule, it is evaluated at R in the direction perpendicular to the planar portion of the molecule. This should give the magnitudes of the fields at the approximate positions of the negative sites. ε(R) is raised to the second power, because the energy of an induced dipole is −0.5αε2 [70]. The coefficients c1–c4 were obtained through the regression procedure for each database separately [71].
The σ-hole and π-hole databases were separately fit to Equation (3) to obtain expressions for predicting the respective interaction energies [69]. The results were encouraging. Therefore, in the present work, we have expanded the analyses, in modified form, to considerably larger databases of σ-hole and π-hole interactions. For each interaction, we have also directly calculated the energy associated with the polarization of the negative site by the electric field of the σ- or π-hole molecule.

6. Present Results

Table 1 lists the computed properties for 54 σ-hole complexes involving σ-hole molecules from Groups IV–VII, with several nitrogen and oxygen Lewis bases as negative sites. About one-fourth of the complexes are hydrogen-bonded. Structures and energies were obtained with the MP2/aug-cc-pVDZ procedure [43], electrostatic potentials, and electric fields with B3PW91/6-31G(d,p) using the MP2 geometries.
Table 2 presents the same properties for 33 π-hole complexes. Table 3 shows the computed VS,min and the experimental polarizabilities [72] of the negative sites. The VS,min are on the 0.001 au surfaces of the negative sites prior to interaction.
In the present study, in addition to using larger databases, we have modified our approach. In Equation (3), the electrostatics are modeled entirely by the σ- or π-hole VS,max, which is on the 0.001 au surface of that molecule in the free, unperturbed state. However as recently suggested [65], it should be more realistic to use the potential V(R) that the σ- or π-hole molecule creates at the point at which the electric field is evaluated [65]. This should better reflect what the negative site actually experiences, and it would also be consistent with earlier work [38,64].
In order to further improve the representation of the electrostatic contribution, we also introduce the most negative potential VS,min on the 0.001 au surface of the negative site. This should be at least somewhat indicative of what is interacting with V(R).
Accordingly, our regression relationship becomes,
ΔE = c1[V(R)] + c2[ε(R)]2 + c3[α] + c4[VS,min] + c5
However, we have found that V(R) and ε(R)2 are rather closely related for the π-hole interactions in Table 2; the coefficient of determination R2 for [ε(R)]2 versus V(R) is 0.918. Therefore, using both of them in the regression relationship would be redundant and statistically undesirable. For the σ-hole interactions, on the other hand, R2 is just 0.649 for [ε(R)]2 versus V(R), indicating that both of them are needed. Therefore, we have used Equation (4) for the σ-hole interactions in Table 1 and Equation (5) for the π-hole interactions in Table 2.
ΔE = c1[ε(R)]2 + c2[α] + c3[VS,min] + c4
The results are gratifying. In Figure 3 and Figure 4, the ΔE predicted by Equations (4) and (5), respectively, are plotted against the computed ΔE in Table 1 and Table 2. For the σ-hole interactions in Figure 3, R2 = 0.938 and the root-mean-square error is 1.04 kcal/mol. For the π-hole interactions in Figure 4, R2 = 0.952 and the root-mean-square error is 2.51 kcal/mol. There is one major outlier, which is in Figure 4. It is for the π-hole interaction between F2Si=O and NCH, for which the predicted ΔE is −30.73 kcal/mol, which is much more negative than the computed −22.58 kcal/mol. If this interaction were omitted from the database, R2 would increase to 0.974, and the root-mean-square error would decrease to 1.89 kcal/mol.
The discrepancy between the predicted and computed ΔE for this complex can be attributed largely to the secondary interactions involving portions of the molecules other than the σ- or π-hole and the primary negative site. In the case of F2Si=O and NCH, the negative fluorines and oxygen of F2Si=O interact repulsively with the negative nitrogen of NCH; the computed F---N and O---N separations, 2.66 and 2.87 Å, are considerably less than the sums of the respective van der Waals radii, 3.12 and 3.16 Å [73]. The computed ΔE takes these destabilizing repulsions into account, but the ΔE predicted by Equation (5) does not. As a result, the predicted ΔE exaggerates the strength of the interaction. Secondary interactions are certainly also implicated in the other discrepancies between predicted and computed ΔE, although their effects are not nearly as great. They are likely to be more significant in π-hole interactions than in σ-hole interactions, because the planar nature of the molecule around the π-hole puts more atoms in closer proximity to the atoms of the negative site.
Could all of the interaction energies in Table 1 and Table 2, both σ-hole and π-hole, be represented by a single regression relationship? We investigated this using Equation (4), which was parametrized to fit all 54 σ-hole and 33 π-hole interaction energies. The resulting correlation between predicted and computed ΔE values is shown in Figure 5. It has an R2 of 0.927 and root-mean-square error of 2.12 kcal/mol. These results are very encouraging, considering that 87 interactions of a variety of types are involved.
The same outlier that appeared in Figure 4 is again shown in Figure 5. It is for the π-hole interaction between F2SiO and NCH, and it is due to significant secondary interactions, as discussed above.

7. Polarization Energies

For each interaction in Table 1 and Table 2, we have calculated the energy of the polarization of the negative site by the electric field of the σ- or π-hole molecule, using Equation (6) [70]:
ΔE(pol) = −0.5[α][ε(R)]2
Note that we are not presently taking into account the polarization of the σ- or π-hole molecule by the negative site.
The calculated ΔE(pol) values are listed in Table 1 and Table 2. The relationship between the computed total interaction energies and the polarization energies is shown in Figure 6 for the σ-hole interactions and in Figure 7 for the π-hole interactions. The overall trends are that the stronger interactions are associated with greater degrees of polarization.
However, most of the interactions are weak to moderate, and polarization makes only a minor contribution. In the σ-hole category (Table 1), 69% have ΔE between zero and −8 kcal/mol, and their average polarization energy ΔE(pol), by Equation (6), is just -0.37 kcal/mol. Among the π-hole (Table 2), 58% have ΔE between zero and −8 kcal/mol, and their average ΔE(pol) is -0.59 kcal/mol.
This explains why polarization can often be neglected and the interaction energies expressed satisfactorily in terms of just the most positive and/or most negative potentials, VS,max and/or VS,min, of the free molecules prior to interaction [25,30,42,52,59,66,67,68]. On the other hand, there are three π-hole complexes in Table 2 (F3B and NCH, FPO2 and CO, FPO2 and OCO) that have interaction energies ΔE between −4 and −7 kcal/mol, but yet have an ΔE(pol) value of about −2 kcal/mol. Thus, a weak or moderate interaction does not necessarily have insignificant polarization. Nevertheless, it seems generally valid to say that the stronger interactions are accompanied by greater degrees of polarization.
Figure 6 and Figure 7 show that strong interactions and high degrees of polarization are more likely for the π-hole interactions in Table 2 than for the σ-hole in Table 1. More than one-third (39%) of the π-holes have ΔE values that are more negative than −10 kcal/mol, reaching −42.38 kcal/mol, and their average ΔE(pol) is -13.8 kcal/mol. In contrast, just 20% of the σ-hole interactions have ΔE values that are more negative than -10 kcal/mol; the most negative is −21.96 kcal/mol, and their average ΔE(pol) is −4.0 kcal/mol.
Why are there such differences between the energetics of the σ-hole and π-hole interactions? One possible factor is that the planar surfaces around π-holes results in more atoms of the interacting molecules being in closer contact. Furthermore, σ-hole interactions are usually accompanied by increases or decreases in the lengths of the bonds that produce the σ-holes [69,74,75], while the π-hole interactions in Table 2 involve bond bending as the π-hole atoms change from trigonal to partial tetrahedral character [11,69]. Since bond stretching or compression requires a greater input of energy than bond bending [76], this would contribute to making σ-hole interactions on the average weaker (i.e., less negative ΔE).

8. Discussion and Summary

We have demonstrated that the energies of σ- and π-hole interactions can be expressed quite well in terms of four properties: the electrostatic potentials V(R) and electric fields ε(R) of the σ- and π-hole molecules at the approximate positions of the negative sites, the most negative potentials VS,min on the 0.001 au surfaces of the negative sites prior to interaction, and the average polarizabilities α of these sites. It would of course be more accurate to use the components of the polarizabilities in the directions of the electric fields, but it is difficult to find reliable values for these [77]. The relative importance of the four properties cited varies somewhat depending upon the particular database; for instance, the interaction energies of the π-hole interactions in Table 2 can be represented without V(R).
The success achieved with Equations (3)–(5) supports the interpretation of these σ- and π-hole interactions as Coulombic in nature, which is consistent with the rigorous Hellmann–Feynman theorem [78,79]. Coulombic should be understood to mean both electrostatics and polarization. Dispersion is included in the polarization contribution, as demonstrated by Feynman [79].
It is noteworthy that interactions with ΔE ranging from −1.7 to −42.4 kcal/mol can be handled in the same manner. No new factors need to be invoked to account for the stronger ones. Overall, they simply involve greater degrees of polarization of the electronic densities of the negative sites toward the σ- or π-hole molecules. It might be suggested that this be described using terms such as “dative” or “coordinate covalent” character [11,25,30,80], but the gradation in interaction energies is in fact simply a manifestation of the continuum nature of chemical bonding, from weak to strong [81,82,83,84].
A major cause of deviations from Equations. (3)–(5) is the secondary interactions between portions of the molecules other than the σ- or π-hole potentials and the negative sites. Examples have been discussed earlier in this paper and in other work [11,65,69,85,86]. The problem arises because the actual computed ΔE does reflect secondary interactions, but Equations (3)–(5) do not.
The importance of secondary interactions increases as more atoms of the two molecules are in closer proximity. Thus, they can be expected to be less significant for σ-hole interactions than for π-hole interactions, and least so when the σ-hole atom is hydrogen or a halogen, since these have only one immediate neighbor, which is the atom to which the hydrogen or halogen is bonded. For example, there are 34 interactions in Table 1 in which the σ-hole is on hydrogen or a halogen. If Equation (4) is fit to the data for these 34 interactions, the R2 is an excellent 0.980, and the root-mean-square error is just 0.60 kcal/mol. This correlation is clearly superior to the one for all 54 σ-hole interactions, as shown in Figure 3, which has an R2 = 0.938 and root-mean-square error = 1.04 kcal/mol.
We conclude by noting again that we have considered only the polarization of the negative site by the σ- or π-hole molecule, not the polarization of the latter by the negative site. It seems reasonable to expect the polarization of the negative site to be the more important, since the negative site is likely to have the higher electronic density and be more polarizable. However, in principle, the present procedure could also be applied to the polarization of the σ- or π-hole molecule.

Author Contributions

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

Funding

This research received no external funding.

Acknowledgments

We greatly appreciate the very useful suggestion by Tore Brinck that we investigate the use of the negative site VS,min in our regression relationships.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Delgado-Barrio, G.; Prat, R.F. Deformed Hartree-Fock Solutions for Atoms. III. Convergent Iterative Process and Results for O. Phys Rev. A 1975, 12, 2288–2297. [Google Scholar] [CrossRef]
  2. Bent, H.A. Structural Chemistry of Donor-Acceptor Interactions. Chem. Rev. 1968, 68, 587–648. [Google Scholar] [CrossRef]
  3. Stevens, E.D. Experimental Electron Density Distribution of Molecular Chlorine. Mol. Phys. 1979, 37, 27–45. [Google Scholar] [CrossRef]
  4. Guru Row, T.N.; Parthasarathy, R. Directional Preferences of Nonbonded Atomic Contacts with Divalent Sulfur in Terms of its Orbital Orientations. 2. S---S Interactions and Nonspherical Shape of Sulfur in Crystals. J. Am. Chem. Soc. 1981, 103, 477–479. [Google Scholar]
  5. Nyburg, S.C.; Faerman, C.H. A Revision of van der Waals Atomic Radii for Molecular Crystals: N, O, F, S, Cl, Se, Br and I Bonded to Carbon. Acta Cryst. B 1985, 41, 274–279. [Google Scholar] [CrossRef]
  6. Ramasubbu, N.; Parthasarathy, R. Stereochemistry of Incipient Electrophilic and Nucleophilic Reactions at Divalent Selenium Center: Electrophilic-Nucleophlic Pairing and Anisotropic Shape of Se in Se-Se Interactions. Phosphorus Sulfur. 1987, 31, 221–229. [Google Scholar] [CrossRef]
  7. Tsirelson, V.G.; Zou, P.F.; Tang, T.H.; Bader, R.F.W. Topological Definition of Crystal Structure: Determination of the Bonded Interactions in Solid Molecular Chlorine. Acta Cryst. A 1995, 51, 143–153. [Google Scholar] [CrossRef]
  8. Batsanov, S.S. Anisotropy of Atomic van der Waals Radii in the Gas Phase and Condensed Molecules. Struct. Chem. 2000, 11, 177–183. [Google Scholar] [CrossRef]
  9. Eramian, H.; Tian, Y.-H.; Fox, Z.; Beneberu, H.Z.; Kertesz, M. On the Anisotropy of van der Waals Atomic Radii of O, S, Se, F, Cl, Br and I. J. Phys. Chem. A 2013, 117, 14184–14190. [Google Scholar] [CrossRef]
  10. Clark, T.; Hennemann, M.; Murray, J.S.; Politzer, P. Halogen Bonding: The σ-Hole. J. Mol. Model. 2007, 13, 291–296. [Google Scholar] [CrossRef]
  11. Murray, J.S.; Lane, P.; Clark, T.; Riley, K.E.; Politzer, P. σ-Holes, π-Holes and Electrostatically-Driven Interactions. J. Mol. Model. 2012, 18, 541–548. [Google Scholar] [CrossRef] [PubMed]
  12. Coppens, P. Experimental electron densities and chemical bonding. Angew. Chem. Int. Ed. 1977, 16, 32–40. [Google Scholar] [CrossRef]
  13. Feil, D. The electron density distribution in molecular crystals: Exploring the boundary between X-ray diffraction and quantum chemistry. Chem. Scr. 1986, 26, 395–408. [Google Scholar]
  14. Klein, C.L.; Stevens, E.D. Charge density studies of drug molecules. In Structure and Reactivity; Liebman, J.F., Greenberg, A., Eds.; VCH Publishers: New York, NY, USA, 1988; pp. 26–64. [Google Scholar]
  15. Stewart, R.F. On the mapping of electrostatic properties from Bragg diffraction data. Chem. Phys. Lett. 1979, 65, 335–342. [Google Scholar] [CrossRef]
  16. Politzer, P.; Murray, J.S. Molecular electrostatic potentials and chemical reactivity. In Reviews in Computational Chemistry; Lipkowitz, K.B., Boyd, D.B., Eds.; VCH Publishers: New York, NY, USA, 1991; Volume 2, pp. 273–312. [Google Scholar]
  17. Politzer, P.; Daiker, K.C. Models for chemical reactivity. In The Force Concept in Chemistry; Deb, B.M., Ed.; Van Nostrand Reinhold: New York, NY, USA, 1981; pp. 294–387. [Google Scholar]
  18. Politzer, P.; Truhlar, D.G. (Eds.) Chemical Applications of Atomic and Molecular Electrostatic Potentials; Plenum Press: New York, NY, USA, 1981. [Google Scholar]
  19. Stone, A.J.; Price, S.L. Some new ideas in the theory of intermolecular forces: Anisotropic atom-atom potentials. J. Phys. Chem. 1988, 92, 3325–3335. [Google Scholar] [CrossRef]
  20. Wiberg, K.B.; Rablen, P.R. Comparison of atomic charges derived via different procedures. J. Comput. Chem. 1993, 14, 1504–1518. [Google Scholar] [CrossRef]
  21. Bachrach, S.M. Population analysis and electron densities from quantum mechanics. In Reviews in Computational Chemistry; Lipkowitz, K.B., Boyd, D.B., Eds.; VCH Publishers: New York, NY, USA, 1994; Volume 5, pp. 171–227. [Google Scholar]
  22. Price, S.L. Applications of realistic electrostatic modelling to molecules in complexes, solids and proteins. J. Chem. Soc. Faraday Trans. 1996, 92, 2997–3008. [Google Scholar] [CrossRef]
  23. Murray, J.S.; Politzer, P. The electrostatic potential: An overview. Wires Comput. Mol. Sci. 2011, 1, 153–163. [Google Scholar] [CrossRef]
  24. Bader, R.F.W.; Carroll, M.T.; Cheeseman, J.R.; Chang, C. Properties of atoms in molecules: Atomic volumes. J. Am. Chem. Soc. 1987, 109, 7968–7979. [Google Scholar] [CrossRef]
  25. Politzer, P.; Murray, J.S.; Clark, T. Halogen bonding and other σ-hole interactions: A perspective. Phys. Chem. 2013, 15, 11178–11189. [Google Scholar] [CrossRef]
  26. Bauzá, A.; Mooibroek, T.J.; Frontera, A. The bright future of unconventional σ/π-hole interactions. Chem. Phys. 2015, 16, 2496–2517. [Google Scholar] [CrossRef]
  27. Politzer, P.; Murray, J.S. An overview of strengths and directionalities of noncovalent interactions: σ-Holes and π-holes. Crystals 2019, 9, 165. [Google Scholar] [CrossRef] [Green Version]
  28. Politzer, P.; Murray, J.S.; Clark, T.; Resnati, G. The σ-hole revisited. Chem. Phys. 2017, 19, 32166–32178. [Google Scholar] [CrossRef]
  29. Politzer, P.; Murray, J.S. σ-Hole interactions: Perspectives and misconceptions. Crystals 2017, 7, 212. [Google Scholar] [CrossRef] [Green Version]
  30. Politzer, P.; Murray, J.S. Halogen bonding: An interim discussion. Chem. Phys. 2013, 14, 278–294. [Google Scholar] [CrossRef]
  31. Benesi, H.A.; Hildebrand, J.H. A spectrophotometric investigation of the interaction of iodine with aromatic hydrocarbons. J. Am. Chem. Soc. 1949, 71, 2703–2707. [Google Scholar] [CrossRef]
  32. Hassel, O.; Romming, C. Direct structural evidence for weak charge transfer bonds in solids containing chemically saturated molecules. Quart. Rev. Chem. Soc. 1962, 16, 1–18. [Google Scholar] [CrossRef]
  33. Blackstock, S.C.; Lorand, J.P.; Kochi, J.K. Charge-transfer interactions of amines with tetrahalomethanes. J. Org. Chem. 1987, 52, 1451–1460. [Google Scholar] [CrossRef]
  34. Ramasubbu, N.; Parthasarathy, R.; Murray-Rust, P. Angular preferences of intermolecular forces around halogen centers: Preferred directions of approach of electrophiles and nucleophiles around carbon-halogen bond. J. Am. Chem. So. 1986, 108, 4308–4314. [Google Scholar] [CrossRef]
  35. Rosenfeld, R.E., Jr.; Parthasarathy, R.; Dunitz, J.D. Directional preferences of nonbonded atomic contacts with divalent sulfur. 1. Electrophiles and nucleophiles. J. Am. Chem. Soc. 1977, 99, 4860–4862. [Google Scholar] [CrossRef]
  36. Politzer, P.; Murray, J.S.; Janjić, G.V.; Zarić, S.D. σ-Hole interactions of covalently- bonded nitrogen, phosphorus and arsenic: A survey of crystal structures. Crystal 2014, 4, 12–31. [Google Scholar] [CrossRef] [Green Version]
  37. Andree, S.N.L.; Sinha, A.S.; Aakeröy, C.B. Structural examination of halogen-bonded co-crystals of tritopic acceptors. Molecules 2018, 23, 163. [Google Scholar] [CrossRef]
  38. Shields, Z.P.-I.; Murray, J.S.; Politzer, P. Directional tendencies of halogen and hydrogen bonding. Intern. J. Quantum Chem. 2010, 110, 2823–2832. [Google Scholar] [CrossRef]
  39. Hennemann, M.; Murray, J.S.; Politzer, P.; Riley, K.E.; Clark, T. Polarization-induced σ-holes and hydrogen bonding. J. Mol. Model. 2012, 18, 2461–2469. [Google Scholar] [CrossRef]
  40. Politzer, P.; Murray, J.S.; Clark, T. σ-Hole bonding: A physical interpretation. Top. Curr. Chem. 2014, 358, 19–42. [Google Scholar]
  41. Politzer, P.; Murray, J.S.; Clark, T. Intermolecular interactions in crystals. In Intermolecular Interactions in Crystals: Fundamentals of Crystal Engineering; Novoa, J.J., Ed.; RSC Publishing: London, UK, 2017. [Google Scholar]
  42. Murray, J.S.; Politzer, P. Hydrogen bonding: A Coulombic σ-hole interaction. J. Indian Inst. Sci. 2019. [Google Scholar] [CrossRef]
  43. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J. Gaussian 09, Revision A.1; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  44. Bulat, F.A.; Toro-Labbé, A.; Brinck, T.; Murray, J.S.; Politzer, P. Quantitative analysis of molecular surfaces: Volumes, electrostatic potentials and average local ionization energies. J. Mol. Model. 2010, 16, 1679–1691. [Google Scholar] [CrossRef]
  45. Politzer, P.; Murray, J.S.; Clark, T. Halogen bonding: An electrostatically-driven highly directional noncovalent interaction. Phys. Chem. Chem. Phys. 2010, 12, 7748–7757. [Google Scholar] [CrossRef]
  46. Politzer, P.; Murray, J.S.; Concha, M.C. σ-Hole bonding between like atoms. A fallacy of atomic charges. J. Mol. Model. 2008, 14, 659–665. [Google Scholar] [CrossRef]
  47. Solimannejad, M.; Malekani, M.; Alkorta, I. Cooperative and diminutive unusual weak bonding in F3CX---HMgH---Y and F3CX---Y---HMgH trimers (X = Cl, Br; Y = HCN and HNC). J. Phys. Chem. A 2010, 114, 12106–12111. [Google Scholar] [CrossRef]
  48. Politzer, P.; Riley, K.E.; Bulat, F.A.; Murray, J.S. Perspectives on halogen bonding and other σ-hole interactions: lex parsimoniae (Occam’s Razor). Comput. Chem. 2012, 998, 2–8. [Google Scholar] [CrossRef]
  49. Clark, T. Polarization, donor-acceptor interactions and covalent contributions in weak interactions: A clarification. J. Mol. Model. 2017, 23, 297. [Google Scholar] [CrossRef]
  50. Edwards, A.J.; Mackenzie, C.F.; Spackman, P.R.; Jayatilaka, D.; Spackman, M.A. Intermolecular interactions in molecular crystals: What’s in a name? Faraday Discuss. 2017, 203, 93–112. [Google Scholar] [CrossRef] [Green Version]
  51. Politzer, P.; Murray, J.S.; Clark, T. Mathematical modeling and physical reality in noncovalent interactions. J. Mol. Model. 2015, 21, 52. [Google Scholar] [CrossRef]
  52. Politzer, P.; Murray, J.S. A unified view of halogen bonding, hydrogen bonding and other σ-hole interactions. In Noncovalent Forces; Scheiner, S., Ed.; Springer: Heidelberg, Germany, 2015; pp. 291–321. [Google Scholar]
  53. Clark, T.; Murray, J.S.; Politzer, P. A perspective on quantum mechanics and chemical concepts in describing noncovalent interactions. Phys. Chem. Chem. Phys. 2018, 20, 30076–30082. [Google Scholar] [CrossRef]
  54. Riley, K.E.; Hobza, P. Investigations into the nature of halogen bonding including symmetry adapted perturbation theory analyses. J. Chem. Theory Comput. 2008, 4, 232–242. [Google Scholar] [CrossRef]
  55. Solimannejad, M.; Gharabaghi, M.; Scheiner, S. SH---N and SH---P blue shifting H-bonds and and N---P interactions in complexes pairing HSN with amines and phosphines. J. Chem. Phys. 2011, 134, 24312. [Google Scholar] [CrossRef]
  56. Murray, J.S.; Shields, Z.P.-I.; Seybold, P.G.; Politzer, P. Intuitive and counterintuitive noncovalent interactions of aromatic π regions with the hydrogen and nitrogen of HCN. J. Comput. Sci. 2015, 10, 209–216. [Google Scholar] [CrossRef]
  57. Geronimo, I.; Singh, N.J.; Kim, K.S. Can electron-rich π systems bind anions? J. Chem. Theory Comput. 2011, 7, 825–829. [Google Scholar] [CrossRef]
  58. Bartlett, R.J.; Weinstein, H. Theoretical treatment of multiple site reactivity in large molecules. Chem. Phys. Lett. 1975, 30, 441–447. [Google Scholar] [CrossRef]
  59. Bonaccorsi, R.; Scrocco, E.; Tomasi, J. A representation of the polarization term in the interaction energy between a molecule and a point-like charge. Theoret. Chim. Acta 1976, 43, 63–73. [Google Scholar] [CrossRef]
  60. Francl, M.M. Polarization correction to electrostatic potentials. J. Phys. Chem. 1985, 89, 428–433. [Google Scholar] [CrossRef]
  61. Orozco, M.; Luque, F.J. Generalization of the molecular electrostatic potential for the study of noncovalent interactions. In Molecular Electrostatic Potentials: Concepts and Applications; Murray, J.S., Sen, K., Eds.; Elsevier: Amsterdam, Holland, 1996; pp. 181–218. [Google Scholar]
  62. Alkorta, I.; Perez, J.J.; Villar, H.O. Molecular polarization maps as a tool for studies of intermolecular interactions and chemical reactivity. J. Mol. Graph. 1994, 12, 3–13. [Google Scholar] [CrossRef]
  63. Clark, T.; Murray, J.S.; Politzer, P. The σ-hole Coulombic interpretation of trihalide anion formation. Chem. Phys. Chem. 2018, 19, 3044–3049. [Google Scholar] [CrossRef]
  64. Brinck, T.; Borrfors, A.N. Electrostatics and polarization determine the strength of the halogen bond: A red card for charge transfer. J. Mol. Model. 2019, 25, 125. [Google Scholar] [CrossRef] [Green Version]
  65. Politzer, P.; Murray, J.S. Electrostatics and polarization in σ- and π-hole noncovalent interactions: An overview. Chem. Phys. Chem. 2019. [Google Scholar] [CrossRef]
  66. Riley, K.E.; Murray, J.S.; Politzer, P.; Concha, M.C.; Hobza, P. Br---O complexes as probes of factors affecting halogen bonding: Interactions of bromobenzenes and bromopyrimidines with acetone. J. Chem. Theory Comput. 2009, 5, 155–163. [Google Scholar] [CrossRef]
  67. 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]
  68. Wang, H.; Wang, W.; Jin, W.H. σ-Hole bond vs π-hole bond: A comparison based on halogen bond. Chem. Rev. 2016, 116, 5072–5104. [Google Scholar] [CrossRef]
  69. Politzer, P.; Murray, J.S.; Clark, T. Explicit inclusion of polarizing electric fields in σ- and π-hole interactions. J. Phys. Chem. A 2019, 123, 10123–10130. [Google Scholar] [CrossRef]
  70. Hirschfelder, J.O.; Curtiss, C.F.; Bird, R.B. Molecular Theory of Gases and Liquids; Wiley: New York, NY, USA, 1954; p. 852. [Google Scholar]
  71. NCSS 11 Statistical Software, version 2016; NCSS, LLC: Kaysville, UT, USA, 2016; Available online: https://www.ncss.com/software/ncss/ (accessed on 27 January 2020).
  72. Johnson, R.D., III (Ed.) NIST Computational Chemistry Comparison and Benchmark DataBase; NIST Standard Reference DataBase No. 101; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2018. Available online: https://cccbdb.nist.gov/ (accessed on 27 January 2020).
  73. Alvarez, S. A cartography of the van der Waals territories. Dalton Trans. 2013, 42, 8617–8636. [Google Scholar] [CrossRef] [Green Version]
  74. Wang, W.; Wong, N.B.; Zheng, W.; Tian, A. Theoretical study on the blueshifting halogen bond. J. Phys. Chem. A 2004, 108, 1799–1805. [Google Scholar] [CrossRef]
  75. Murray, J.S.; Concha, M.S.; Lane, P.; Hobza, P.; Politzer, P. Blue shifts vs. red shifts in σ-hole bonding. J. Mol. Model. 2008, 14, 699–704. [Google Scholar] [CrossRef]
  76. Crews, P.; Jaspars, M.; Rodriguez, J. Organic Structure Analysis; Oxford University Press: New York, NY, USA, 2009. [Google Scholar]
  77. Couling, V.W.; Halliburton, B.W.; Keir, R.L.; Ritchie, G.L.D. Anisotropic molecular polarizabilities of HCHO, CH3CHO and CH3COCH3, Rayleigh depolarization ratios of HCHO and CH3CHO and first and second Kerr virial coefficients of CH3COCH. J. Phys. Chem. A 2001, 105, 4365–4370. [Google Scholar] [CrossRef]
  78. Hellmann, H. Einführung in die Quantenchemie; Deuticke: Leipzig, Germany, 1937. [Google Scholar]
  79. Feynman, R.P. Forces in molecules. Phys. Rev. 1939, 56, 340–343. [Google Scholar] [CrossRef]
  80. Politzer, P.; Murray, J.S. Halogen bonding and beyond: Factors influencing the nature of CN-R and SiN-R complexes with F-Cl and Cl2. Chem. Acc. 2012, 131, 1114. [Google Scholar] [CrossRef]
  81. Slater, J.C. Hellmann-Feynman and virial theorems in the Xα method. J. Chem. Phys. 1972, 57, 2389–2396. [Google Scholar] [CrossRef]
  82. Bader, R.F.W. Bond paths are not chemical bonds. J. Phys. Chem. A 2009, 113, 10391–10396. [Google Scholar] [CrossRef] [Green Version]
  83. Rahm, M.; Hoffmann, R. Distinguishing bonds. J. Am. Chem. Soc. 2016, 138, 3731–3744. [Google Scholar] [CrossRef]
  84. Politzer, P.; Murray, J.S. A look at bonds and bonding. Struct. Chem. 2019, 30, 1153–1157. [Google Scholar] [CrossRef]
  85. Riley, K.E.; Murray, J.S.; Fanfrlίk, J.; Řezáč, J.; Solá, R.J.; Concha, M.C.; Ramos, F.M.; Politzer, P. Halogen bond tunability I: The effects of aromatic fluorine substitution on the strengths of halogen-bonding interactions involving chlorine, bromine and iodine. J. Mol. Model. 2011, 17, 3309–3318. [Google Scholar] [CrossRef]
  86. Metrangolo, P.; Murray, J.S.; Pilati, T.; Politzer, P.; Resnati, G.; Terraneo, G. Fluorine-centered halogen bonding: A factor in recognition phenomena and reactivity. Cryst. Growth Des. 2011, 11, 4238–4246. [Google Scholar] [CrossRef]
Figure 1. Computed electrostatic potentials on the 0.001 au molecular surface of SeClBr. In (a), the selenium is in the foreground, the chlorine is in the rear left, and the bromine in the rear right. In (b), the chlorine is in the foreground left and the bromine in the foreground right. Color ranges, in kcal/mol: red, more positive than 24; yellow, between 24 and 12; green, between 12 and 0; blue, negative. Black hemispheres indicate the most positive potentials (VS,max) corresponding to the two σ-holes on the selenium, on the extensions of the Cl–Se and Br–Se bonds, and the σ-holes on the chlorine and bromine, on the extensions of the Se–Cl and Se–Br bonds.
Figure 1. Computed electrostatic potentials on the 0.001 au molecular surface of SeClBr. In (a), the selenium is in the foreground, the chlorine is in the rear left, and the bromine in the rear right. In (b), the chlorine is in the foreground left and the bromine in the foreground right. Color ranges, in kcal/mol: red, more positive than 24; yellow, between 24 and 12; green, between 12 and 0; blue, negative. Black hemispheres indicate the most positive potentials (VS,max) corresponding to the two σ-holes on the selenium, on the extensions of the Cl–Se and Br–Se bonds, and the σ-holes on the chlorine and bromine, on the extensions of the Se–Cl and Se–Br bonds.
Crystals 10 00076 g001
Figure 2. Computed electrostatic potential on the 0.001 au molecular surface of SeO3. Selenium is in the center. Color ranges, in kcal/mol: red, more positive than 36; yellow, between 36 and 18; green, between 18 and 0; blue, negative. Black hemisphere indicates the most positive potential (VS,max) corresponding to the π-holes above and below selenium.
Figure 2. Computed electrostatic potential on the 0.001 au molecular surface of SeO3. Selenium is in the center. Color ranges, in kcal/mol: red, more positive than 36; yellow, between 36 and 18; green, between 18 and 0; blue, negative. Black hemisphere indicates the most positive potential (VS,max) corresponding to the π-holes above and below selenium.
Crystals 10 00076 g002
Figure 3. ΔE predicted with Equation (4) plotted against the computed ΔE, in kcal/mol, for the σ-hole interactions in Table 1. R2 = 0.938.
Figure 3. ΔE predicted with Equation (4) plotted against the computed ΔE, in kcal/mol, for the σ-hole interactions in Table 1. R2 = 0.938.
Crystals 10 00076 g003
Figure 4. ΔE predicted with Equation (5) plotted against the computed ΔE, in kcal/mol, for the π-hole interactions in Table 2. R2 = 0.952.
Figure 4. ΔE predicted with Equation (5) plotted against the computed ΔE, in kcal/mol, for the π-hole interactions in Table 2. R2 = 0.952.
Crystals 10 00076 g004
Figure 5. ΔE predicted with Equation (4) plotted against the computed ΔE, in kcal/mol, for all of the σ- and π-hole interactions in Table 1 and Table 2. R2 = 0.927.
Figure 5. ΔE predicted with Equation (4) plotted against the computed ΔE, in kcal/mol, for all of the σ- and π-hole interactions in Table 1 and Table 2. R2 = 0.927.
Crystals 10 00076 g005
Figure 6. Computed ΔE plotted against ΔE(pol) obtained with Equation (6), in kcal/mol, for the σ-hole interactions in Table 1.
Figure 6. Computed ΔE plotted against ΔE(pol) obtained with Equation (6), in kcal/mol, for the σ-hole interactions in Table 1.
Crystals 10 00076 g006
Figure 7. Computed ΔE plotted against ΔE(pol) obtained with Equation (6), in kcal/mol, for the π-hole interactions in Table 2.
Figure 7. Computed ΔE plotted against ΔE(pol) obtained with Equation (6), in kcal/mol, for the π-hole interactions in Table 2.
Crystals 10 00076 g007
Table 1. Computed data for σ-hole interactions. The σ-hole is on the extension of the bond between atoms, and it is indicated in bold. Dashed lines show interactions. ΔE is interaction energy, as shown in Equation (2). VS,max is the most positive electrostatic potential of the σ-hole on the 0.001 au molecular surface before interaction. R is equilibrium interaction separation. V(R) and ε(R) are the electrostatic potential and electric field of the σ-hole molecule at the approximate position of the negative site. ΔE(pol) is the energy of polarization of the negative site, as shown in Equation (6).
Table 1. Computed data for σ-hole interactions. The σ-hole is on the extension of the bond between atoms, and it is indicated in bold. Dashed lines show interactions. ΔE is interaction energy, as shown in Equation (2). VS,max is the most positive electrostatic potential of the σ-hole on the 0.001 au molecular surface before interaction. R is equilibrium interaction separation. V(R) and ε(R) are the electrostatic potential and electric field of the σ-hole molecule at the approximate position of the negative site. ΔE(pol) is the energy of polarization of the negative site, as shown in Equation (6).
InteractionΔE (kcal/mol)VS,max (kcal/mol)R (Å)V(R) (kcal/mol)ε(R) (au)ΔE(pol) (kcal/mol)
FF---OH2−1.6814.52.602.70.002549−0.02
FF---O=CH2−2.0414.52.542.80.002781−0.05
H3C(Cl)Se---NCH−2.588.83.191.80.001770−0.02
HF2CBr---NCH−2.9716.53.123.10.003319−0.06
HF2CBr---O=CH2−3.4716.53.023.50.003856−0.09
HCCH---NCH−3.5830.72.309.10.006834−0.26
H2NH---NH3−3.6225.52.287.70.005442−0.13
NC-Cl---OH2−3.6535.32.8813.60.007823−0.19
HF2CBr---NH3−3.6716.53.063.30.003621−0.06
H3C(Cl)Se---NH3−3.918.83.131.70.001990−0.02
F3CBr---NCH−3.9224.13.047.10.005416−0.16
F3CBr---O=CH2−4.0424.12.947.80.006187−0.22
F3CH---NCH−4.1331.22.3312.60.007259−0.29
F3CH---O=CH2−4.3231.22.3112.80.007431−0.32
HOH---NCH−4.4143.42.1115.20.01097−0.66
F2(NC)P---NCH−4.6036.52.9812.00.007417−0.30
H3(F)Si---NCH−4.6734.12.9112.20.008482−0.39
HCCH---NH3−4.7130.72.269.50.007218−0.23
F3CBr---NH3−4.9724.12.977.60.005923−0.16
HF2CBr---NH2CH3−5.0516.52.983.70.004098−0.13
NC-Br---OH2−5.1542.72.8718.30.01112−0.40
F3CH---NH3−5.3331.22.2813.10.007667−0.26
HCCH---NH2CH3−5.5130.72.1710.30.008089−0.52
NC-H---OH2−5.6650.02.0522.80.01439−0.66
FCl---OH2−5.9340.32.5316.50.01396−0.62
FCl---NCH−5.9640.32.6015.40.01256−0.87
F3As---NCH−6.0136.22.9114.60.01044−0.60
HF2CBr---NH(CH3)2−6.1716.52.904.10.004658−0.25
HCCH---NH(CH3)2−6.2630.72.0911.10.009099−0.95
F3CH---NH2CH3−6.2931.22.2413.40.007994−0.51
F3CBr---NH2CH3−6.4224.12.888.20.006701−0.36
Cl-Cl---NH3−6.4824.12.558.90.009112−0.37
(NC)2S---NCH−6.6543.82.9217.00.01033−0.59
HOH---NH3−6.9643.41.9717.10.01325−0.78
F2(NC)P---NH3−6.9836.52.7314.70.01032−0.48
H3FSi---NH3−7.2634.12.5616.80.01416−0.89
F3CBr---NH(CH3)2−7.7924.12.808.90.007504−0.65
HOH---NH2CH3−8.1343.41.9118.10.01442−1.65
(NC)2S---O=CH2−8.2743.82.8218.30.01156−0.79
FBr---OH2−8.4347.72.5322.30.01908−1.16
F3As---O=CH2−8.4436.22.6817.80.01407−1.16
(NC)2S---NH3−8.5643.82.8617.80.01104−0.54
F2(NC)P---NH2CH3−9.4536.52.5616.90.01305−1.35
(NC)2S---NH2CH3−10.1043.82.7419.50.01267−1.28
F3As---NH3−10.1336.22.5919.40.01611−1.16
H3FSi---NH2CH3−10.3534.12.3421.10.02059−3.37
FCl---NH3−12.1840.32.2721.80.02138−2.03
F2(NC)P---NH(CH3)2−12.2636.52.3920.00.01713−3.38
(NC)2S---NH(CH3)2−12.3943.82.6620.70.01393−2.24
F3As---NH2CH3−13.1036.22.5121.10.01825−2.65
H3FSi---NH(CH3)2−13.2234.12.2723.00.02358−6.41
F3As---NH(CH3)2−16.2936.22.4123.50.02156−5.36
FCl---NH2CH3−17.2840.32.1525.30.02697−5.78
FCl---NH(CH3)2−21.9640.32.0927.40.03012−10.46
Table 2. Computed data for π-hole interactions. Atoms with π-holes are indicated in bold. Interactions are with the first atom listed for the negative site. ΔE is interaction energy, as shown in Equation (2). VS,max is the most positive electrostatic potential of the π-hole on the 0.001 au molecular surface before interaction. R is the equilibrium interaction separation. V(R) and ε(R) are the electrostatic potential and electric field of the π-hole molecule at the approximate position of the negative site. ΔE(pol) is the energy of polarization of the negative site, as shown in Equation (6).
Table 2. Computed data for π-hole interactions. Atoms with π-holes are indicated in bold. Interactions are with the first atom listed for the negative site. ΔE is interaction energy, as shown in Equation (2). VS,max is the most positive electrostatic potential of the π-hole on the 0.001 au molecular surface before interaction. R is the equilibrium interaction separation. V(R) and ε(R) are the electrostatic potential and electric field of the π-hole molecule at the approximate position of the negative site. ΔE(pol) is the energy of polarization of the negative site, as shown in Equation (6).
π-Hole MoleculeNegative SiteΔE (kcal/mol)VS,max (kcal/mol)R (Å)V(R) (kcal/mol)ε(R) (au)ΔE(pol) (kcal/mol)
SeO2OCO−2.3635.43.1910.50.007622−0.36
SeO2CO−2.4235.43.448.50.005766−0.14
Cl2C=OCO−2.5522.83.163.90.003483−0.05
Cl2C=OOCO−2.7022.82.885.30.005104−0.16
Cl2C=ONCH−3.9922.82.895.20.005042−0.14
FPO2OCO−4.0058.42.7120.50.01795−1.99
SO2NCH−4.0432.93.0310.20.007796−0.33
SeO2NCH−4.1435.43.0811.60.008652−0.41
FNO2NCH−4.3832.82.808.50.007202−0.28
Cl3BNCH−4.4624.02.846.70.006555−0.24
F2C=ONCH−4.5840.92.779.00.007798−0.33
Cl2C=ONH3−4.6422.82.855.50.005332−0.13
FNO2NH3−5.5232.82.798.50.007301−0.24
FPO2CO−5.9158.42.5324.90.02326−2.24
SO2NH3−6.0032.92.8312.30.01004−0.45
F2C=ONH3−6.1440.92.6710.00.008952−0.36
F3BNCH−6.9448.82.3517.90.01836−1.85
H3CPO2NCH−7.5947.62.6814.80.01361−1.02
SeO2NH3−7.6235.42.8714.00.01123−0.56
SO2NH2CH3−9.5232.92.5915.60.01398−1.55
H2Si=ONCH−11.5543.42.1830.70.03635−7.25
FPO2NCH−11.8858.42.2136.30.03868−8.21
SeO2NH2CH3−12.2335.42.5818.80.01693−2.28
SO2NH(CH3)2−12.9332.92.4318.70.01793−3.71
F2Si=OCO−13.5266.72.1245.90.05101−10.76
SeO2NH(CH3)2−16.6635.42.4222.40.02177−5.47
H3CPO2NH3−18.4147.62.0532.90.04145−7.65
F2Si=ONCH−22.5866.72.0153.40.06293−21.74
H2Si=ONH3−26.6943.42.0338.20.04925−10.80
H2Si=ONH2CH3−31.9143.42.0039.70.05195−21.46
FPO2NH3−32.3558.41.9551.50.06246−17.37
H2Si=ONH(CH3)2−36.3943.41.9443.70.05916−40.36
F2Si=ONH3−42.3866.71.9558.20.07063−22.21
Table 3. Computed VS,min and experimental polarizabilities of negative sites.
Table 3. Computed VS,min and experimental polarizabilities of negative sites.
MoleculeVS,min (kcal/mol) a Polarizability, α (Å3) b
CO2−12.62.911
CO−14.0 1.953
H2C=O −30.2 2.77
HCN−32.22.593
NH(CH3)2 −39.15.447
H2O−39.6 1.501
NH2CH3−42.93.754
NH3−46.22.103
a The VS,min are on the nitrogen and oxygen atoms except for CO, for which it is on the carbon atom. b Reference [72].

Share and Cite

MDPI and ACS Style

Murray, J.S.; Politzer, P. Interaction and Polarization Energy Relationships in σ-Hole and π-Hole Bonding. Crystals 2020, 10, 76. https://doi.org/10.3390/cryst10020076

AMA Style

Murray JS, Politzer P. Interaction and Polarization Energy Relationships in σ-Hole and π-Hole Bonding. Crystals. 2020; 10(2):76. https://doi.org/10.3390/cryst10020076

Chicago/Turabian Style

Murray, Jane S., and Peter Politzer. 2020. "Interaction and Polarization Energy Relationships in σ-Hole and π-Hole Bonding" Crystals 10, no. 2: 76. https://doi.org/10.3390/cryst10020076

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

Article Metrics

Back to TopTop