A Combined Theoretical and Cambridge Structural Database Study of

Mar 28, 2014 - hole along with an analysis the Cambridge Structural Database. We have found a large number of hits that provide strong experimental su...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCA

A Combined Theoretical and Cambridge Structural Database Study of π‑Hole Pnicogen Bonding Complexes between Electron Rich Molecules and Both Nitro Compounds and Inorganic Bromides (YO2Br, Y = N, P, and As) Antonio Bauzá, Rafael Ramis, and Antonio Frontera* Departament de Química, Universitat de les Illes Balears, Crta. de Valldemossa km 7.5, 07122 Palma de Mallorca (Baleares), Spain S Supporting Information *

ABSTRACT: Quantum calculations at the DFT-D3/def2-TZVPD level of theory have been used to examine complexes between O2YBr (YN, P, and As) molecules and several Lewis bases, that is, NH3, H2O, and HF. The interactions of the lone pair of the ammonia, water, and hydrogen fluoride with the σ-hole and π-hole of O2YBr molecules have been considered. In general, the complexes where the Lewis base lone pair interacts with the πhole are more favorable than those with σ-hole. The nature of the interactions has been characterized with the Bader theory of atoms in molecules (AIM). We have also studied the ability of trifluoronitromethane and nitromethane to interact with anions using their πhole along with an analysis the Cambridge Structural Database. We have found a large number of hits that provide strong experimental support for ability of the nitryl (−NO2) group to interact with anions and Lewis bases. In some X-ray structures, the π-hole interaction is crucial in the crystal packing and has a strong influence in the solid state architecture of the complexes. Finally, due to the relevance in atmospheric chemistry, we have studied noncovalent σ/π-hole complexes of nitryl bromide with ozone.



INTRODUCTION Noncovalent interactions dominate many aspects of contemporary chemistry and biology. Nowadays the construction of any host to bind virtually any guest is possible thanks to modern synthetic chemistry methods even in competitive media.1,2 The progress in understanding many biological functions and drug design has been done mostly due to the chemical and physics insights acquired from the study of synthetic host−guest complexes3−5 and the development of supramolecular chemistry. Clear examples are cation−π6,7 and anion−π8−13 interactions that previously to the demonstration of their prominent role in living systems they first became apparent in artificial complexes. This also applies for other noncovalent interactions involving aromatic moieties,14,15 weak C−H hydrogen bonds16−18 or the interaction between halogen atoms and Lewis bases.19−22 Definitely the most important aspect of host−guest chemistry is the possibility not only to elucidate the mechanism of all contributions to molecular recognition, but also to clarify geometric constraints and to assign discrete energy values to them.23 This can help to develop energy scoring functions for many purposes like crystal engineering, catalysis, and drug design. A positive π-hole24 has been defined by Murray et al.25 as a region of positive electrostatic potential that is perpendicular to a portion of a molecular framework. It is the counterpart of a σhole, which is along the extension of a covalent bond to an atom.26 In a given group of the periodic table, σ/π-holes become more positive on going from the lighter to the heavier © 2014 American Chemical Society

atoms. In addition, the presence of electron withdrawing groups in the rest of the molecule strongly affects the σ/π-hole magnitude.27−29 Positive σ/π-holes interact in a highly directional manner with concentrations of negative charge, for example, anions or the lone pairs of Lewis bases. Recently, several theoretical works have studied competition and interplay effects between σ-hole, π-hole, and hydrogen bond interactions. In some of them, nitryl halides (NO2X) molecules have been used as simultaneous σ/π-hole donors.30−33 The interest in these molecules arises from their importance in atmospheric chemistry34,35 since they are involved in stratospheric ozone depletion cycles.36,37 Moreover, recent field observations have demonstrated the importance of nitryl bromide in chemical cycles involving halogen species and heterogeneous reactions in atmospheric plumes of quiescently degassing volcanoes.38 It should be also mentioned that Alkorta and co-workers have recently analyzed orthogonal interactions between nitryl derivatives and electron donors.39 Moreover, they have studied pnicogen bonds involving sp2 hybridized phosphorus atoms.40 Finally, monomeric phosphenic bromide PO2Br is produced by a photochemical reaction between the high-temperature molecule OPBr and O3 in solid Ar. It has been demonstrated by IR spectroscopy and 15O NMR that it is a planar molecule with C2v symmetry.41 Received: March 6, 2014 Revised: March 28, 2014 Published: March 28, 2014 2827

dx.doi.org/10.1021/jp502301n | J. Phys. Chem. A 2014, 118, 2827−2834

The Journal of Physical Chemistry A

Article

N) where the bromine has been substituted by a −CF3 group presents also a very positive value of electrostatic potential due to its strong electron-withdrawing effect. Finally, compound 5 (Figure 2E) also exhibits a large positive region that it is originated from the influence of the nitrogen atom and the acidic hydrogen atoms. The most positive region is found close to the hydrogen atoms. Therefore, in this compound a hydrogen bonding interaction using the C−H bond is favored with respect to a π-hole interaction. The MEPS of compounds 1−3 also show a σ-hole along the extension of the Br−Y covalent bond that ranges 30−34 kcal/mol. Therefore, the MEP analysis indicates that π-hole interactions are favored with respect to σ-hole halogen bond in compounds 2 and 3 and a competitive σ/π-hole interaction is expected in 1. Neutral Complexes. The σ-hole halogen and π-hole pnicogen bonding complexes considered in this study are shown in Figure 3. We have used three Lewis bases that range from moderately strong (NH3) to very weak (HF). The energetic results are gathered in Table 1, and from the inspection of the results some considerations can be made. First, the π-hole complexes are more favorable than the σ-hole ones apart from the complexes of ammonia and water with nitryl bromide. Second, the π-hole complexes involving the heavier pnicogens exhibit very large and negative interaction energies, in agreement with the MEP results commented above. The short equilibrium distances observed in the π-hole complexes of the heavier pnicogens indicate a partial covalent bonding, especially in complexes 7 and 8. This aspect is further analyzed below in the AIM analysis. The optimized geometries of some representative complexes and their corresponding AIM distribution of critical points and bond paths are shown in Figure 4. The planarity of the −YO2 group is lost upon complexation, resulting in a slightly pyramidalization of the pnicogen atom. A common feature of all compounds (see Figure 4) upon complexation is the formation of a bond critical point located along the line connecting the donor atom with the halogen/pnicogen atom. As aforementioned, complexes 7 and 8 have large interaction energies and short equilibrium distances that suggest some degree of covalent bonding. The values of the Laplacian (∇2ρ(r)) for complexes 6−23 are also included in Table 1, and all complexes apart from 7 have positive values as is common in closed shell noncovalent interactions. Interestingly, complex 7 exhibits a modest but negative value of ∇2ρ(r) that confirms the partial covalence of the N···P interaction in this system. Furthermore, it has been previously demonstrated that the value of the charge density measured at the bond critical point can be used as a measure of the strength of the interaction in a variety of noncovalent interactions.50−52 The values of the charge density computed at the bond critical point for complexes 6−23 are gathered in Table 1. Inspection of the results indicates that there is a strong correlation between the interaction energy (EBSSE) and the electron charge density at the bond critical point ρ(3,−1). In Figure 5, we represent the plot of the regression between ρ(3,−1) and EBSSE for complexes 6−23, with a regression coefficient r = 0.983. It is worth stressing the importance of this relationship, since it includes both σ-hole halogen bond and π-hole pnicogen bonding complexes and allows for dealing simultaneously with both interactions. Anionic Complexes. In addition to the neutral complexes described above we have evaluated the anion binding ability of trifluoronitromethane and nitromethane by means of their π-

In this Article, we report a theoretical DFT study (BP86-D3/ def2-TZVP) where we analyze the energetic and geometric features of a series of complexes between several π-hole donor molecules, including BrYO2 (Y = N, P, and As) and several lone pair donor molecules (NH3, H2O, and HF). Moreover, we have studied the ability of CF3NO2, and CH3NO2, to interact with anions using the π-hole. We also provide experimental evidence from the Cambridge Structural Database42 that demonstrates the importance of π-hole interaction in the solid state. It should be mentioned that the nitro group is often used to modify the electronic nature or aromatic rings or aliphatic chains. However, their ability to participate in π-hole interactions is normally not considered by researchers and it may have a strong influence in the crystal packing of the compounds, as demonstrated herein. Moreover, the BrYO2 (Y = N, P, and As) molecules are ditopic σ-hole halogen bond donors and π-hole pnicogen bond donors. We have analyzed energetically and using the Bader’s theory of atoms in molecules43−45 the relative importance of both interactions than can be useful to explain their reactivity.



THEORETICAL METHODS The geometries of all complexes were optimized at the BP86D3/def2TZVPD level of theory by means of the Turbomole 6.4 software.46 For the theoretical study, we have used the BP86-D3 DFT method that includes the latest available correction for dispersion effects. The interaction energies were calculated with correction for the basis set superposition error (BSSE) by using the Boys−Bernardi counterpoise technique.47 The Cs symmetry constrain has been imposed in the optimizations of all complexes, unless otherwise noted. The Bader “Atoms in molecules” (AIM) theory43 was successfully applied to the study of the noncovalent interactions discussed herein. The AIM analysis was carried out by means of the AIMall calculation package.48 The MEPS analysis was performed at the BP86/def2TZVP level of theory by using the Gaussian 09 package.49



RESULTS AND DISCUSSION MEP Study. First of all we have computed the molecular electrostatic potential surfaces (MEPS) of the Lewis acids studied in this work (see Figure 1). The representation of the

Figure 1. σ/π-Hole donor and acceptor molecules and anions used in this work.

MEPS is shown in Figure 2 where we have also indicated the value of the energy potential at the π-hole. Interestingly, the value of the energy potential at the π-hole in compounds 2 and 3 (Y = P and As) is around 20 kcal/mol more positive than that in 1 (Figure 2A). Therefore, a considerably higher ability of 2 and 3 (Figure 2B and C) to interact with Lewis is anticipated. Moreover, compound 4 (Figure 2D) that is similar to 1 (Y 2828

dx.doi.org/10.1021/jp502301n | J. Phys. Chem. A 2014, 118, 2827−2834

The Journal of Physical Chemistry A

Article

Figure 2. MEPS computed for compounds 1−5 at the BP86-D3/def2-TZVP level of theory.

29) present more modest binding energies, which include contribution from both hydrogen bonding and π-hole interactions. The optimized geometries of two representative complexes and their corresponding AIM distribution of critical points and bond paths are shown in Figure 6. The π-hole interaction in all complexes is characterized by the presence of a bond critical point connecting the halide with the nitrogen atom (see Figure 6A). In addition, a second bond critical point emerges upon complexation in nitrometane as a consequence of the X−···HC interaction. Therefore, complexes 27−29 are further characterized by the presence of a ring critical point (see Figure 6B). The Laplacian is positive in all complexes and the value of the charge density measured at the bond critical point is also a good indicator of the strength of the interaction for the anionic complexes (see Table 2). CSD Analysis. We have performed several searches in the Cambridge Structural Database (CSD) in order to provide experimental support to the anion/Lewis base binding ability of the nitro group by means of the π-hole. It is well-known that the CSD is a convenient and reliable tool for analyzing geometrical parameters53 and often reveals aspects that have not been noticed by the original authors. We have performed three separate searches of nitro-derivatives establishing short contacts (N···X distance lesser than the van der Waals radii sum) with F, Cl, and Br. Interestingly, we have found 178 X-ray structures for π-hole N···F contacts, 56 for N···Cl, and 12 for N···Br. This large number of X-ray structures shows that this interaction is common in the solid state. We have selected several structures where the interaction plays a crucial role determining the 3D architecture of the crystal packing. In particular, we show in Figure 7 the selected hits for neutral complexes and their CSD reference codes. It can be observed that the X···N interaction is responsible for the formation of infinite 1D chains in the solid state of ALAYIW (Figure 7A),54 FUBLAQ 55 (Figure 7B), and YUQJIE 56 (Figure 7C) compounds. Moreover, in GENSUO57 structure (Figure 7D), the double F···N interaction between the para-fluoro substituent and the 2-nitroiminoimidazolidine moiety controls

Figure 3. σ/π-Hole complexes 6−29 studied in this work.

hole on the nitrogen atom. The MEP surface plot of Figure 2 clearly shows a well-defined π-hole in trifluoronitromethane and, conversely, for nitromethane the MEP surface shows a large blue area that encompasses a large region including the methyl hydrogen atoms and the nitrogen atom. The global minimum in nitromethane is located close to one hydrogen atom indicating that this molecule has more ability to form hydrogen bonds than π-hole pnicogen bonds. As a matter of fact, all attempts to optimize π-hole complexes of nitromethane with halides always end to the formation of a hydrogen bonded complex X−···HCH2NO2. Therefore, for these complexes, we have evaluated the π-hole interaction using counterpoisecorrected equilibrium geometries computed within the frozenmonomer approximation by computing single point energies with the halide located above the nitrogen atom at 0.05 Å increments. The results are summarized in Table 2. As expected, the trifluoromethane complexes 24−26 present very favorable interaction energies in agreement with the large positive potential at the π-hole of CF3NO2. The fluoride complex is the most favorable, and chloride and bromide complexes present similar interaction energies in agreement with the equilibrium distances. Nitromethane complexes (27− 2829

dx.doi.org/10.1021/jp502301n | J. Phys. Chem. A 2014, 118, 2827−2834

The Journal of Physical Chemistry A

Article

Table 1. Interaction Energies at the BP86-D3/def2-TZVPD Level of Theory without and with the BSSE Correction (E and EBSSE in kcal/mol, respectively), Equilibrium Distances (Re in Å), and Electron Charge Density and Its Laplacian (ρ(r) and ∇2ρ(r) in a.u., respectively) at the Bond Critical Point That Emerges upon Complexation in Compounds 6−23 complex

E

EBSSE

Re

102 × ρ(r)

102 × ∇2ρ(r)

6 (H3N:NO2Br) 7 (H3N:PO2Br) 8 (H3N:AsO2Br) 9 (H3N:BrNO2) 10 (H3N:BrPO2) 11 (H3N:BrAsO2) 12 (H2O:NO2Br) 13 (H2O:PO2Br) 14 (H2O:AsO2Br) 15 (H2O:BrNO2) 16 (H2O:BrPO2) 17 (H2O:BrAsO2) 18 (HF:NO2Br) 19 (HF:PO2Br) 20 (HF:AsO2Br) 21 (HF:BrNO2) 22 (HF:BrPO2) 23 (HF:BrAsO2)

−3.5 −27.9 −24.9 −9.0 −5.7 −6.6 −2.2 −13.5 −13.4 −3.9 −2.9 −3.2 −1.0 −3.7 −4.0 −0.5 −1.2 −1.1

−3.2 −27.4 −24.3 −8.7 −5.5 −6.3 −2.1 −13.1 −13.0 −3.8 −2.8 −3.1 −0.9 −3.5 −3.8 −0.5 −1.1 −1.0

2.967 1.962 2.163 2.581 2.861 2.775 2.888 2.136 2.294 2.791 2.980 2.948 2.973 2.644 2.687 3.319 3.248 3.267

0.98 10.60 8.15 3.78 2.18 2.61 0.85 6.76 5.50 2.15 1.42 1.56 0.60 1.99 1.99 0.42 0.49 0.48

3.46 −8.28 4.90 8.81 5.82 6.62 3.95 1.83 7.73 6.69 4.88 5.17 2.95 5.88 6.15 2.01 2.42 2.33

Figure 4. Distribution of critical points and bond paths in complexes 6, 9, 13, 16, 20, and 23.

Table 2. Interaction Energies at the BP86-D3/def2-TZVP Level of Theory without and with the BSSE Correction (E and EBSSE in kcal/mol, respectively), Equilibrium Distances (Re in Å), and Electron Charge Density and Its Laplacian (ρ(r) and ∇2ρ(r) in a.u., respectively) at the Bond Critical Point That Emerges upon Complexation in Compounds 24− 29 complex 24 25 26 27 28 29

Figure 5. Plot of the regression between the interaction energy and the charge density at the bond critical point for complexes 6−23.

the formation of an interesting self-assembled dimer in the solid state. Some examples involving anions are also gathered in Figure 8 including BF 4 − (QASBES, 58 Figure 8A), Cl − (MAZZAP,59 Figure 8B), and Br− (HIQYAH,60 Figure 8C) that give strong support to the theoretical calculations explained above regarding the anion binding ability of the



(F :CF3NO2) (Cl−:CF3NO2) (Br−:CF3NO2) (F−:CH3NO2) (Cl−:CH3NO2) (Br−:CH3NO2)

E

EBSSE

Re

102 × ρ(r)

102 × ∇2ρ(r)

−27.3 −12.4 −10.5 −11.6 −6.7 −6.0

−27.1 −12.2 −10.3 −11.5 −6.7 −5.9

2.064 2.922 3.155 2.550 3.250 3.450

5.91 1.80 1.40 1.95 0.84 0.69

19.10 4.40 3.06 7.06 2.91 2.33

nitro group using the π-hole. In addition, we also represent in Figure 8 (bottom) some X-ray fragments where nitromethane is involved in π-hole interaction with BF4−, PF6−, and SbF6− anions that correspond to FEBXAL61 (Figure 8D), FIWMUU62 2830

dx.doi.org/10.1021/jp502301n | J. Phys. Chem. A 2014, 118, 2827−2834

The Journal of Physical Chemistry A

Article

bromine (Figure 9F) or the nitrogen atom (Figure 9G). In the bifurcated complex 32, the interaction is further characterized by the presence of a ring critical point (Figure 9H). The π-hole complexes are characterized by the presence of a bond critical point connecting the oxygen atom of ozone with the nitrogen atom (Figure 9I). Since the interaction energies of the complexes and the differences among them are very small, we have also optimized complexes 30−34 at the ab initio RI-MP2/ def2-TZVP level of theory and computed the interaction energies. In addition, we have also performed single point calculations (using the RI-MP2 geometries) at the CCSD(T)/ def2-TZVP level of theory in order to validate the DFT energetic study. The interaction energies and equilibrium distances at the three levels of theory are summarized in Table 3. At the ab initio levels, the energetic differences between the different binding modes are very small in agreement with the results obtained at DFT/def2-TZVP level of theory method. The interaction energies at the CCSD(T)/def2-TZVP//RIMP2/def2-TZVP level of theory indicate that σ-hole and π-hole complexes 30−34 are isoenergetic.

Figure 6. Distribution of bond (red) and ring (yellow) critical points and bond paths in complexes 25 and 27.

(Figure 8E), and XELVIU63 (Figure 8F) codes, respectively. In these structures, solvent molecules of nitromethane are incorporated in the crystal structure of organic cations (not shown in Figure 8), and interestingly they are located in the proximity of the counterion (anions) establishing, among others, stabilizing F···N interactions. Ozone Complexes. Due to the importance of both molecules in atmospheric chemistry, we have also analyzed the interaction modes of ozone with nitryl bromide. We have computed several σ-hole halogen (see Figure 9A-C) and π-hole pnicogen bonding complexes (see Figure 9D,E) in order to investigate the most favorable orientation. The interaction energies are more modest than those previously described for the σ/π-hole complexes of nitryl bromide with water (complexes 12 and 15, see Table 1). At the BP86-D3 level, the σ-hole halogen bonding complexes 31 (Cs symmetry, Figure 9B) and 32 (C2v symmetry, Figure 9C) are slightly more favorable than the π-hole pnicogen bonding. The σ-hole complexes are characterized by the presence of a bond critical point connecting the oxygen atom of ozone with either the



CONCLUDING REMARKS In conclusion, we have studied the σ/π-hole interaction competition in complexes of BrYO2 (Y = N, P, and As) compounds with Lewis bases. The σ-hole halogen bond interaction is more favorable only for nitryl bromide interacting with NH3, H2O and O3. For the rest of the compounds, the πhole pnicogen bonding is more favorable. The AIM analysis demonstrates that the value of the density at the bond critical point that emerges upon complexation is a good measure of bond order and interaction strength in σ/π-hole complexes. We have also studied the energetic and geometric features of

Figure 7. Partial view of several crystal structures retrieved from the CSD. The reference codes are indicated. Distances in Å. 2831

dx.doi.org/10.1021/jp502301n | J. Phys. Chem. A 2014, 118, 2827−2834

The Journal of Physical Chemistry A

Article

Figure 8. Partial view of several crystal structures retrieved from the CSD. The reference codes are indicated. Distances in Å.

Figure 9. BP86-D3/def2-TZVP optimized geometries of ozone−nitryl bromide complexes 30−34 are shown along with the interaction energies (BSSE corrected) and the distribution of critical points and bond paths. Distances in Å.

is certainly important to gain knowledge in the intricate mechanism that governs the molecular recognition and crystal packing. To this respect, the assignment of energy values to them can help to understand these mechanistic contributions to the supramolecular chemistry and crystal engineering community.

Table 3. Interaction Energies at the BP86-D3/def2-TZVP, RI-MP2/def2-TZVP, and CCSD(T)/def2-TZVP//RI-MP2/ def2-TZVP Levels of theory with the BSSE Correction (EBSSE in kcal/mol) and Equilibrium Distances (Re in Å) for Complexes 30−34 EBSSE

Re

complex

BP86-D3

RI-MP2

CCSD(T)

BP86-D3

RI-MP2

30 31 32 33 34

−1.2 −1.9 −1.9 −1.3 −0.8

−1.2 −1.2 −1.1 −1.7 −1.3

−0.9 −1.1 −1.1 −1.1 −1.0

3.030 2.906 3.177 3.099 3.066

3.134 3.177 3.385 2.953 3.023



ASSOCIATED CONTENT

S Supporting Information *

Cartesian coordinated of all compounds and complexes (1−29) are given. The list of CSD reference codes is also provided. This material is available free of charge via the Internet at http://pubs.acs.org.



anion···π-hole complexes in trifluoronitromethane and nitromethane. To this respect, we have analyzed the CSD to find experimental examples of such interaction, and we have found a large number of structures that confirm the existence and importance of the interaction of Lewis bases and anions with nitroderivatives, including nitromethane. This theoretical study

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +34 971173498. Fax: +34 971 173426. 2832

dx.doi.org/10.1021/jp502301n | J. Phys. Chem. A 2014, 118, 2827−2834

The Journal of Physical Chemistry A

Article

Author Contributions

(21) Erdelyi, M. Halogen Bonding in Solution. Chem. Soc. Rev. 2012, 41, 3547−3557. (22) Parisini, E.; Metrangolo, P.; Pilati, T.; Resnati, G.; Terraneo, G. Halogen Bonding in Halocarbon−Protein Complexes: a Structural Survey. Chem. Soc. Rev. 2010, 40, 2267−2278. (23) Schneider, H.−J. Binding Mechanisms in Supramolecular Complexes. Angew. Chem., Int. Ed. Engl. 2009, 48, 3924−3977. (24) Politzer, P.; Murray, J. S.; Clark, T. Halogen Bonding and Other σ-Hole Interactions: a Perspective. Phys. Chem. Chem. Phys. 2010, 12, 7748−7757. (25) Murray, J. S.; Lane, P.; Clark, T.; Riley, K. E.; Politzer, P. σHoles, π-Holes and Electrostatically Driven Interactions. J. Mol. Model. 2012, 18, 541−548. (26) Clark, T.; Henneman, M.; Murray, J. S.; Politzer, P. Halogen Bonding: the Sigma Hole. J. Mol. Model. 2007, 13, 291−296. (27) (a) Murray, J. S.; Riley, K. E.; Politzer, P.; Clark, T. Directional Weak Intermolecular Interactions: Sigma-Hole Bonding. Aust. J. Chem. 2010, 63, 1598−1607. (28) Bauzá, A.; Mooibroek, T. J.; Frontera, A. Tetrel Bonding: A Rediscovered Supramolecular Force? Angew. Chem., Int. Ed. 2013, 52, 12317−12321. (29) Grabowski, S. J. Tetrel Bond−σ-Hole Bond as a Preliminary Stage of the SN2 Reaction. Phys. Chem. Chem. Phys. 2014, 16, 1824− 1834. (30) Solimannejad, M.; Nassirinia, N.; Amani, S. A computational study of 1:1 and 1:2 complexes of nitryl halides (O2NX) with HCN and HNC. Struct. Chem. 2013, 24, 651−659. (31) Esrafili, M. D.; Mohammadian−Sabet, F.; Solimannejad, M. A Theoretical Evidence for Mutual Influencie Between S···N(C) and Hydrogen/Lithium/Halogen Bond: Competition and Interplay Between π-Hole and σ-Hole Interactions. Struct Chem. 2014, DOI: 10.1007/s11224-014-0392-8. (32) Solimannejad, M.; Ramezani, V.; Trujillo, C.; Alkorta, I.; Sánchez−Sanz, G.; Elguero, J. Competition and Interplay between σHole and π-Hole Interactions: a Computational Study of 1:1 and 1:2 Complexes of Nitryl Halides (O2NX) with Ammonia. J. Phys. Chem. A 2012, 116, 5199−5206. (33) del Bene, J. E.; Alkorta, I.; Elguero, J. Properties of Complexes H2C(X)P:PXH2, for X = F, Cl, OH, CN, NC, CCH, H, CH3, and BH2: P···P Pnicogen Bonding at σ-Holes and π-Holes. J. Phys. Chem. A 2013, 117, 11592−11604. (34) Fickert, S.; Helleis, F.; Adams, J. W.; Moortgat, G. K.; Crowley, J. N. Reactive Uptake of ClNO2 on Aqueous Bromide Solutions. J. Phys. Chem. A 1998, 102, 10689−10696. (35) Lesclaux, R.; Caralp, F.; Dognon, A. M.; Cariolle, D. The Rate of Formation of Halomethyl Peroxy Nitrates in the Stratosphere and their Possible Role as Temporary Reservoirs for ClOx and NOX Species. Geophys. Res. Lett. 1986, 13, 933−936. (36) Molina, M. J.; Rowland, F. S. Stratospheric Sink for Chlorofluoromethanes: Chlorine Atom−Catalysed Destruction of Ozone. Nature 1974, 249, 810−812. (37) Wayne, R. P. Chemistry of Atmospheres; Oxford University Press: Oxford, U.K., 1991. (38) von Glasow, R. Atmospheric Chemistry in Volcanic Plumes. Proc. Nat. Acad. Sci. 2010, 107, 6594−6599. (39) Sánchez-Sanz, G.; Trujillo, C.; Solimannejad, M.; Alkorta, I.; Elguero, J. Orthogonal Interactions between Nitryl Derivatives and Electron Donors: Pnictogen Bonds. Phys. Chem. Chem. Phys. 2013, 15, 14310−14318. (40) del Bene, J. E.; Alkorta, I.; Elguero, J. Characterizing Complexes with Pnicogen Bonds Involving sp2 Hydridized Phosphorus Atoms: (H2CPX2) with X = F, Cl, OH, CN, NC, CCH, H,CH3 and BH2. J. Phys. Chem. A 2013, 117, 6893−6903. (41) Schunck, S.; Goecke, H. J.; Koeppe, R.; Schnoeckel, H. Molekulares PO2Br Matrix−IR−Untersuchungen und ab initio SCF− Rechnungen. ZAAC 1989, 579, 66−74. (42) Allen, F. H. The Cambridge Structural Database: a Quarter of a Million Crystal Structures and Rising. Acta Crystallogr. 2002, B58, 380−388.

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank the DGICYT of Spain (Projects CTQ2011-27512/ BQU and CONSOLIDER INGENIO 2010 CSD2010-00065, FEDER funds) and the Direcció General de Recerca i Innovació del Govern Balear (Project 23/2011, FEDER funds) for financial support.



REFERENCES

(1) Oshovsky, G. V.; Reinhoudt, D. N.; Verboom, W. Supramolecular Chemistry in Water. Angew. Chem., Int. Ed. 2007, 46, 2366−2393. (2) Dsouza, R. N.; Pischel, U.; Nau, W. M. Fluorescent Dyes and their Supramolecular Host−Guest Complexes with Macrocycles in Aqueous Solution. Chem. Rev. 2011, 111, 7941−7980. (3) Zürcher, M.; Diederich, F. Structure−Based Drug Design: Exploring the Proper Filling of Apolar Pockets at Enzyme Active Sites. J. Org. Chem. 2008, 73, 4345−4361. (4) Smith, D. K. A Supramolecular Approach to Medicinal Chemistry: Medicine Beyond Molecule. J. Chem. Educ. 2005, 82, 393−400. (5) Uhlenheuer, D. A.; Petkau, K.; Brunsveld, L. Combining Supramolecular Chemistry with Biology. Chem. Soc. Rev. 2010, 39, 2817−2826. (6) Ma, J. C.; Dougherty, D. A. The Cation−π Interaction. Chem. Rev. 1997, 97, 1303−1324. (7) Zacharias, N.; Dougherty, D. A. Cation−π Interactions in Ligand Recognition and Analysis. Trends Pharmacol. Sci. 2002, 23, 281−287. (8) Frontera, A.; Gamez, P.; Mascal, M.; Mooibroek, T. J.; Reedijk, J. Putting Anion−π Interactions into Perspective. Angew. Chem. Int. Ed. 2011, 50, 9564−9583. (9) Schottel, B. L.; Chifotides, H. T.; Dunbar, K. R. Supramolecular Chemistry of Anions. Chem. Soc. Rev. 2008, 37, 68−83. (10) Hay, B. P.; Bryantsev, V. S. Anion Arene Adducts: C−H Hydrogen Bonding, Anion−π Interaction, and Carbon Motifs Bonding. Chem. Commun. 2008, 2417−2428. (11) Frontera, A. Encapsulation of Anions: Macrocyclic Receptors Based on Metal Coordination and Anion−π Interactions. Coord. Chem. Rev. 2013, 257, 1716−1727. (12) Bauzá, A.; Quiñonero, D.; Deyà, P. M.; Frontera, A. Anion−π Interactions in [S4N3]+ Rings. New J. Chem. 2013, 37, 2636−2641. (13) Bauzá, A.; Quiñonero, D.; Deyà, P. M.; Frontera, A. Theoretical ab initio Study of Anion−π Interactions in Inorganic Rings. Chem. Phys. Lett. 2012, 530, 145−150. (14) Meyer, E. A.; Castellano, R. K.; Diederich, F. Interactions with Aromatic Rings in Chemical and Biological Recognition. Angew. Chem., Int. Ed. Engl. 2003, 42, 1210−1250. (15) Salonen, L. M.; Ellermann, M.; Diederich, F. Aromatic Rings in Chemical and Biological Recognition: Energetics and Structures. Angew. Chem., Int. Ed. 2011, 50, 4808−4842. (16) Desiraju, G. R.; Steiner, T. The Weak Hydrogen Bond; Oxford University Press: Oxford, 1999. (17) Steiner, T. The Hydrogen Bond in the Solid State. Angew. Chem., Int. Ed. 2002, 41, 48−76. (18) Castellano, R. K. Progress toward Understanding the Nature and Function of CH···O Interactions. Curr. Org. Chem. 2004, 8, 845−865. (19) Legon, A. C. The Halogen Bond: an Interim Perspective. Phys. Chem. Chem. Phys. 2010, 12, 7736−7747. (20) Scholfield, M. R.; Van der Zanden, C. M.; Carter, M.; Ho, P. S. Halogen Bonding (X-Bonding): a Biological Perspective. Protein Sci. 2013, 22, 139−152. 2833

dx.doi.org/10.1021/jp502301n | J. Phys. Chem. A 2014, 118, 2827−2834

The Journal of Physical Chemistry A

Article

(63) Schottel, B. L.; Chifotides, H. T.; Shatruk, M.; Chouai, A.; Perez, L. M.; Bacsa, J.; Dunbar, K. R. Anion−π Interactions as Controlling Elements in Self-Assembly Reactions of Ag(I) Complexes with πAcidic Aromatic Rings. J. Am. Chem. Soc. 2006, 128, 5895−5912.

(43) Bader, R. F. W. A Quantum Theory of Molecular Structure and its Applications. Chem. Rev. 1991, 91, 893−928. (44) Bader, R. F. W. Atoms in Molecules, A Quantum Theory; Oxford University Press: Oxford, UK, 1990. (45) Popelier, P. L. A. Atoms In Molecules. An Introduction; Prentice Hall: Harlow, England, 2000. (46) Ahlrichs, R.; Bär, M.; Hacer, M.; Horn, H.; Kömel, C. Electronic Structure Calculation on Workstation Computers: The Program System Turbomole. Chem. Phys. Lett. 1989, 162, 165−169. (47) Boys, S. B.; 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. (48) Todd, A.; Keith, T. K. AIMall, version 13.05.06; Gristmill Software: Overland Park, KS, 2013. (49) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A. et al. Gaussian 09, Revision B.01; Gaussian, Inc.: Wallingford, CT, 2009. (50) Bauzá, A.; Quiñonero, D.; Frontera, A.; Deyà, P. M. Substituent Effects in Halogen Bonding Complexes between Aromatic Donors and Acceptors: A Comprehensive ab initio Study. Phys. Chem. Chem. Phys. 2011, 13, 20371−20379. (51) Lucas, X.; Frontera, A.; Quiñonero, D.; Deyà, P. M. Substituent Effects in Ion−pi Interactions: Fine−Tuning via the Ethynyl Group. J. Phys. Chem. A 2010, 114, 1926−1930. (52) Escudero, D.; Estarellas, C.; Frontera, A.; Quiñonero, D.; Deyà, P. M. Theoretical and Crystallographic Study of Edge−To−Face Aromatic Interactions between Pyridine Moieties and Benzene. Chem. Phys. Lett. 2009, 468, 280−285. (53) Nangia, A.; Biradha, K.; Desiraju, G. R. Correlation of biological activity in β−lactam antibiotics with Woodward and Cohen structural parametersa Cambridge database study. J. Chem. Soc., Perkin Trans. 2 1996, 943−953. (54) Cresswell, A. J.; Davies, S. G.; Lee, J. A.; Roberts, P. M.; Russell, A. J.; Thomson, J. E.; Tyte, M. J. β-Fluoroamphetamines via the Stereoselective Synthesis of Benzylic Fluorides. Org. Lett. 2010, 12, 2936−2939. (55) Clapham, K. M.; Batsanov, A. S.; Bryce, M. R.; Tarbit, B. Trifluoromethyl-Substituted Pyrydyl- and Pyrazolylboronic Acids and Esters: Synthesis and Suzuki−Miyaura Cross-Coupling Reactions. Org. Biomol. Chem. 2009, 7, 2155−2161. (56) Fun, H.-K.; Goh, J. H.; Chandrakantha, B.; Isloor, A. M.; Shetty, P. 2-Chloro-4-nitro-1H-imidazole. Acta Crystallogr. 2010, E66, o1828− o1829. (57) Ravikumar, K.; Sridhar, B.; Mahesh, M.; Reddy, V. V. N. 1-[5(4-Fluorophenyl)-1,2-oxazol-3-ylmethyl]-2-nitroiminoimidazolidine. Acta Crystallogr. 2006, E62, o3850−o3851. (58) Manivannan, S.; Dhanuskodi, S.; Kirschbaum, K.; Tiwari, S. K. Design of an Efficient Solution Grown Semiorganic NLO Crystal for Short Wavelenght Generation: 2-Amino-5-nitropyridinium Tetrafluoroborate. Cryst. Growth Des. 2005, 5, 1463−1468. (59) Rathore, R. S.; Narasimhamurthy, T.; Vijay, T.; Yathirajan, H. S.; Nagaraja, P. 5-Chlorocarbonyl-10,11-dihydro-5H-dibenz[b,f ] azepine. Acta Crystallogr. 2006, C62, o36−o38. (60) Zaleski, J.; Daszkiewicz, Z.; Kyziol, J. 1-Methyl-4-(Nmethylnitramino) Pyridinium Bromide. Acta Crystallogr. 1999, C55, 691−693. (61) Blake, A. J.; Casabo, J.; Devillanova, F. A.; Escriche, L.; Garau, A.; Isaia, F.; Lippolis, V.; Kivekas, R.; Muns, V.; Schroder, M.; et al. Aza−Thia Crowns Containing the 1,10-Phenanthroline Sub-Unit. Substitution Reactions in [Nil(MeCN)][BF4]2 {L=2,5,8-Trithia[9](2,9)-1,10-phenanthrolinophane}. J. Chem. Soc., Dalton Trans. 1999, 1085−1902. (62) Grant, G. J.; Chen, W.; Goforth, A. M.; Baucom, C. L.; Patel, K.; Repovic, P.; van Derveer, D. G.; Pennington, W. T. Anion and Solvent Effects Upon the Structure of Platinum (II)Complexes with Thiacrown Ligands: the Crystal Structures of [Pt(9S3)2](PF6)22 CH3NO2, [Pt(9S3)2](BF4)2·2CH3NO2, [Pt(9S3)2](Otf)·2CH3NO2, And [Pt(18S6)](BF4)2. Eur. J. Inorg. Chem. 2005, 479−485. 2834

dx.doi.org/10.1021/jp502301n | J. Phys. Chem. A 2014, 118, 2827−2834