Next Article in Journal
Bulk Cyclotron Resonance in the Topological Insulator Bi2Te3
Next Article in Special Issue
Halogen Bonds Fabricate 2D Molecular Self-Assembled Nanostructures by Scanning Tunneling Microscopy
Previous Article in Journal
Preparation and Electrochemical Properties of Co3O4 Supercapacitor Electrode Materials
Previous Article in Special Issue
N/N Bridge Type and Substituent Effects on Chemical and Crystallographic Properties of Schiff-Base (Salen/Salphen) Niii Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

σ- and π-Hole Interactions

Departament de Química, Universitat de les Illes Baleares, Crta de Valldemossa km 7.5, 07122 Palma de Mallorca, Baleares, Spain
Crystals 2020, 10(9), 721; https://doi.org/10.3390/cryst10090721
Submission received: 17 August 2020 / Accepted: 18 August 2020 / Published: 19 August 2020
(This article belongs to the Special Issue σ- and π-Hole Interactions)
Supramolecular chemistry is a very active research field that was initiated in the last century [1,2,3,4]. It was defined as chemistry beyond the molecule, and the word supermolecule was invented by Lehn [3]. The chemistry beyond the molecule refers to organized entities of higher complexity resulting from the association of molecules that are held together by noncovalent interactions [5]. The organized supramolecular entities are built by the formation of various noncovalent forces, which are frequently working synergistically in the same supramolecular assembly. Therefore, precise control of the noncovalent interactions is needed to succeed in this field, as exemplified by many regulation processes in nature.
A deep understanding of noncovalent interactions is necessary to advance in many fields, especially in crystal growth and crystal engineering [6]. Theoreticians have demonstrated that the distribution of the electron density around covalently bonded atoms is not isotropic, revealing that the use of point charges to define the properties of an atom (electron-rich or electron-poor) is not valid [7]. That is, a single atom presents regions of higher and lower electron density, where the electrostatic potential can be negative and positive, respectively, in some cases. The positive area is usually defined as a σ- or π-hole, depending on its location. These holes of electron density are responsible for the formation of attractive interactions with any electron-rich site (anion, Lewis base, π-system, etc.). The halogen bond can be considered as the prototypical example of σ-hole interaction [8]. After the emergence of the halogen bond, the interest in σ- and π-hole interactions embracing elements of groups 12–16 [9,10,11,12,13] and 18 [14,15,16,17] of the Periodic Table has grown exponentially. Halogen and chalcogen bonding interactions have already been defined by the IUPAC [18,19]. They are well-recognized interactions that are used regularly by the scientific community in crystal engineering, supramolecular chemistry, and catalysis [20]. However, more experimental and theoretical work is probably needed to extend such a statement to the elements of groups 12–15, acting as Lewis acids.
This issue gathers nine excellent contributions. In reference [21], Alkorta et al. combined theoretical calculations and a search in the Cambridge Structural Database (CSD) to investigate the interaction of dinuclear Ag(I) pyrazolates with Lewis bases, as examples of regium bonding [22]. They studied the effect of the substituents and ligands on the aromaticity and found an interesting relationship between the intramolecular Ag–Ag distance and stability.
In reference [23], Varadwaj et al. studied theoretically the CH3Cl molecule and its complexes with ten Lewis bases to demonstrate that CH3Cl is a genuine halogen bond donor. They have evidenced that the electronic charge density distribution around the Cl is anisotropic. The negative belt is able to participate in halogen, chalcogen, or hydrogen bonding interactions. Moreover, they show that the positive σ-hole on the Cl atom in CH3Cl is not induced by the electric field of the interacting species, as previously suggested in the literature. Instead, it is an inherent property of chlorine in this molecule.
In reference [24], Belmont-Sanchez et al. reported the synthesis and X-ray characterization of several out-of-sphere cadmium complexes with 2,6-diaminopurine. The crystal packing of these compounds is mostly dominated by H-bonds, which were analyzed by using DFT calculations. Interestingly, the results were in clear contrast with those previously reported for similar complexes with adenine instead of 2,6-diaminopurine [25]. The factors contributing to such differences are discussed and rationalized on the basis of the additional exocyclic 2-amino group in 2,6-diaminopurine compared to adenine.
In reference [26], Kletsov et al. synthesized and X-ray characterized four N-substituted 1,3,5-triazinanes and focused on the crucial role of C–H···π and C–H···O H-bonding interactions determining their solid-state architecture. Quite remarkable is the fact that the XRD analysis demonstrated an unprecedented feature of the crystalline structure. That is, the symmetrically substituted 1,3,5-triazacyclohexanes have two chemically identical sulfonamide N-atoms in different sp2 and sp3 hybridizations.
In reference [27], Zhang et al. reported the synthesis and X-ray characterization of a cocrystal formed by hexamethylbenzene (HMB) combined with 1,3-diiodotetrafluorobenzene (1,3-DITFB) founding an unexpected sandwiched-layer structure. The formation of the alternating layer was further studied using DFT calculations showing that dispersion forces are very important in the formation of the HMB layer. In contrast, the formation of the 1,3-DITFB layer is induced by weak but cooperative C–I···F halogen bonds.
In reference [28], Yannacone et al. studied the nature of π-hole interaction in several fluorinated aromatic systems focusing on the effect of the substituents and the presence/absence of heteroatoms in the arene on the strength of the π-hole interaction. Moreover, the authors have also analyzed cooperativity effects with other interactions like hydrogen bonding.
In reference [29], Novoa-Ramírez et al. have used thirteen ligands (N,N’-bis(5-R-salicylidene)ethylenediamine (where R = MeO, Me, OH, H, Cl, Br, NO2) and (N,N’-bis(5-R-salicylidene)-1,2-phenylenediamine (where R = MeO, Me, OH, H, Cl, Br) to synthesize and X-ray characterize thirteen nickel complexes. By using Hirshfeld surface analysis, they showed that their packaging was favored by C∙∙∙H/H∙∙∙C interactions, C–H∙∙∙O hydrogen, and π-stacking interactions.
This Special Issue also includes two reviews, one written by Tiekink [30], who elegantly describes the results of a survey of X-ray structures of main group element compounds (M = Sn, Pb As, Sb, Bi and Te) exhibiting intermolecular M···Se noncovalent interactions. The second review written by Alkorta, Elguero, and I [31], provides a consistent description of noncovalent interactions, covering most groups of the Periodic Table. The interactions are described and discussed using their trivial names. That is, apart from hydrogen bonds, the following noncovalent interactions are described: alkali, alkaline earth, regium, spodium, triel, tetrel, pnictogen, chalcogen, halogen, and aerogen, thus covering a wide range of interactions. In this review, the possibility of extending the Cahn-Ingold-Prelog priority rules to noncovalent interactions is suggested.
In summary, this Special Issue gathers an interesting group of manuscripts devoted to the study of several types of σ- and π-hole noncovalent interactions and their importance in the solid-state of different compounds, including biologically relevant ones like diaminopurines, good halogen bond donors like 1,3-diiodotetrafluorobenzene, and several theoretical investigations devoted to π-hole interactions in arenes and regium bonds in Ag(I) derivatives. Moreover, two excellent and comprehensive reviews are published in this collection with the latest advances in noncovalent interactions that I believe make this Special Issue even more special.
To finish, I wish to thank all authors who have submitted their excellent papers to this Special Issue and also the reviewers who carefully read them, providing constructive and helpful suggestions and corrections on all manuscripts. I am especially thankful to the editorial staff at Crystals for their incredibly fast and professional work, dealing with all manuscripts and the selection of suitable referees.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Ariga, K.; Kunitake, T. Supramolecular Chemistry: Fundamentals and Applications; Springer: Berlin/Heidelberg, Germany, 2006. [Google Scholar]
  2. Pedersen, C.J. The Discovery of Crown Ethers (Noble Lecture). Angew. Chem. Int. Ed. Engl. 1988, 27, 1021–1027. [Google Scholar] [CrossRef]
  3. Lehn, J.-M. Supramolecular Chemistry—Scope and Perspectives Molecules, Supermolecules, and Molecular Devices (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 1988, 27, 89–112. [Google Scholar] [CrossRef]
  4. Cram, D.J. The Design of Molecular Hosts, Guests, and Their Complexes (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 1988, 27, 1009–1020. [Google Scholar] [CrossRef]
  5. Lehn, J.-M. Supramolecular chemistry: Receptors, catalysts, and carriers. Science 1985, 227, 849–856. [Google Scholar] [CrossRef]
  6. Desiraju, G.R. Crystal Engineering: From Molecule to Crystal. J. Am. Chem. Soc. 2013, 135, 9952–9967. [Google Scholar] [CrossRef]
  7. 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]
  8. Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, T. The Halogen Bond. Chem. Rev. 2016, 116, 2478–2601. [Google Scholar] [CrossRef] [Green Version]
  9. Bauzá, A.; Alkorta, I.; Elguero, J.; Mooibroek, T.J.; Frontera, A. Spodium Bonds: Noncovalent Interactions Involving Group 12 Elements. Angew. Chem. Int. Ed. 2020. [Google Scholar] [CrossRef]
  10. Grabowski, S.J. Boron and other triel lewis acid centers: From hypovalency to hypervalency. ChemPhysChem. 2014, 15, 2985–2993. [Google Scholar] [CrossRef]
  11. Bauzá, A.; Mooibroek, T.J.; Frontera, A. Tetrel-bonding interaction: Rediscovered supramolecular force? Angew. Chem. Int. Ed. 2013, 52, 12317–12321. [Google Scholar] [CrossRef]
  12. Mahmudov, K.T.; Gurbanov, A.V.; Aliyeva, V.A.; Resnati, G.; Pombeiro, A.J.L. Pnictogen bonding in coordination chemistry. Coord. Chem. Rev. 2020, 418, 213381. [Google Scholar] [CrossRef]
  13. Scilabra, P.; Terraneo, G.; Resnati, G. The Chalcogen Bond in Crystalline Solids: A World Parallel to Halogen Bond. Acc. Chem. Res. 2019, 52, 1313–1324. [Google Scholar] [CrossRef] [PubMed]
  14. Bauzá, A.; Frontera, A. Aerogen bonding interaction: A new supramolecular force? Angew. Chem. Int. Ed. 2015, 54, 7340–7343. [Google Scholar] [CrossRef] [PubMed]
  15. Bauzá, A.; Frontera, A. σ/π-Hole noble gas bonding interactions: Insights from theory and experiment. Coord. Chem. Rev. 2020, 404, 213112. [Google Scholar] [CrossRef]
  16. Gomila, R.M.; Frontera, A. Covalent and Non-covalent Noble Gas Bonding Interactions in XeFn Derivatives (n = 2–6): A Combined Theoretical and ICSD Analysis. Front. Chem. 2020, 8, 395. [Google Scholar] [CrossRef] [PubMed]
  17. Frontera, A. Noble Gas Bonding Interactions Involving Xenon Oxides and Fluorides. Molecules 2020, 25, 3419. [Google Scholar] [CrossRef]
  18. 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]
  19. Aakeroy, C.B.; Bryce, D.L.; Desiraju, G.R.; Frontera, A.; Legon, A.C.; Nicotra, F.; Rissanen, K.; Scheiner, S.; Terraneo, G.; Metrangolo, P.; et al. Definition of the chalcogen bond (IUPAC Recommendations 2019). Pure Appl. Chem. 2019, 91, 1889–1892. [Google Scholar] [CrossRef]
  20. Taylor, M.S. Anion recognition based on halogen, chalcogen, pnictogen and tetrel bonding. Coord. Chem. Rev. 2020, 413, 213270. [Google Scholar] [CrossRef]
  21. Alkorta, I.; Trujillo, C.; Sánchez-Sanz, G.; Elguero, J. Regium Bonds between Silver(I) Pyrazolates Dinuclear Complexes and Lewis Bases (N2, OH2, NCH, SH2, NH3, PH3, CO and CNH). Crystals 2020, 10, 137. [Google Scholar] [CrossRef] [Green Version]
  22. Frontera, A.; Bauzá, A. Regium–π bonds: An Unexplored Link between Noble Metal Nanoparticles and Aromatic Surfaces. Chem. Eur. J. 2018, 24, 7228–7234. [Google Scholar] [CrossRef] [PubMed]
  23. Varadwaj, P.R.; Varadwaj, A.; Marques, H.M. Does Chlorine in CH3Cl Behave as a Genuine Halogen Bond Donor? Crystals 2020, 10, 146. [Google Scholar] [CrossRef] [Green Version]
  24. Belmont-Sánchez, J.C.; Ruiz-González, N.; Frontera, A.; Matilla-Hernández, A.; Castiñeiras, A.; Niclós-Gutiérrez, J. Anion–Cation Recognition Pattern, Thermal Stability and DFT-Calculations in the Crystal Structure of H2dap[Cd(HEDTA)(H2O)] Salt (H2dap = H2(N3,N7)-2,6-Diaminopurinium Cation). Crystals 2020, 10, 304. [Google Scholar] [CrossRef] [Green Version]
  25. Serrano-Padial, E.; Choquesillo-Lazarte, D.; Bugella-Altamirano, E.; Castineiras, A.; Carballo, R. Niclos-Gutierrez, New copper(II) compound having protonated forms of ethylenediaminetetraacetate(4−) ion (EDTA) and adenine (AdeH): Synthesis, crystal structure, molecular recognition and physical properties of (AdeH2)[Cu(HEDTA)(H2O)]·2H2O. J. Polyhedron. 2002, 21, 1451. [Google Scholar] [CrossRef]
  26. Kletskov, A.V.; Gil, D.M.; Frontera, A.; Zaytsev, V.P.; Merkulova, N.L.; Beltsova, K.R.; Sinelshchikova, A.A.; Grigoriev, M.S.; Grudova, M.V.; Zubkov, F.I. Intramolecular sp2–sp3 Disequalization of Chemically Identical Sulfonamide Nitrogen Atoms: Single Crystal X-Ray Diffraction Characterization, Hirshfeld Surface Analysis and DFT Calculations of N-Substituted Hexahydro-1,3,5-Triazines. Crystals 2020, 10, 369. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Wang, J.-G.; Wang, W. Unexpected Sandwiched-Layer Structure of the Cocrystal Formed by Hexamethylbenzene with 1,3-Diiodotetrafluorobenzene: A Combined Theoretical and Crystallographic Study. Crystals 2020, 10, 379. [Google Scholar] [CrossRef]
  28. Yannacone, S.; Freindorf, M.; Tao, Y.; Zou, W.; Kraka, E. Local Vibrational Mode Analysis of π–Hole Interactions between Aryl Donors and Small Molecule Acceptors. Crystals 2020, 10, 556. [Google Scholar] [CrossRef]
  29. Novoa-Ramírez, C.S.; Silva-Becerril, A.; Olivera-Venturo, F.L.; García-Ramos, J.C.; Flores-Alamo, M.; Ruiz-Azuara, L. N/N Bridge Type and Substituent Effects on Chemical and Crystallographic Properties of Schiff-Base (Salen/Salphen) Niii Complexes. Crystals 2020, 10, 616. [Google Scholar] [CrossRef]
  30. Tiekink, E.R.T. A Survey of Supramolecular Aggregation Based on Main Group Element⋯Selenium Secondary Bonding Interactions—A Survey of the Crystallographic Literature. Crystals 2020, 10, 503. [Google Scholar] [CrossRef]
  31. Alkorta, I.; Elguero, J.; Frontera, A. Not Only Hydrogen Bonds: Other Noncovalent Interactions. Crystals 2020, 10, 180. [Google Scholar] [CrossRef] [Green Version]

Share and Cite

MDPI and ACS Style

Frontera, A. σ- and π-Hole Interactions. Crystals 2020, 10, 721. https://doi.org/10.3390/cryst10090721

AMA Style

Frontera A. σ- and π-Hole Interactions. Crystals. 2020; 10(9):721. https://doi.org/10.3390/cryst10090721

Chicago/Turabian Style

Frontera, Antonio. 2020. "σ- and π-Hole Interactions" Crystals 10, no. 9: 721. https://doi.org/10.3390/cryst10090721

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