research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

ISSN: 2052-2525

The nature of halogen bonding: insights from interacting quantum atoms and source function studies

crossmark logo

aDepartment of Pharmaceutical Sciences, Università degli Studi di Milano, via Mangiagalli 25, 20133 Milano, Italy, bDepartment of Chemistry, Università degli Studi di Milano, via Golgi 19, 20133 Milano, Italy, and cCNR-SCITEC, Institute of Chemical Sciences and Technologies `Giulio Natta' and INSTM RU, via Golgi 19, 20133 Milano, Italy
*Correspondence e-mail: alessandra.forni@cnr.it, stefano.pieraccini@unimi.it, maurizio.sironi@unimi.it

Edited by A. Martín Pendás, Universidad de Oviedo, Spain (Received 3 October 2024; accepted 15 January 2025; online 27 January 2025)

This article is part of a collection of articles on Quantum Crystallography, and commemorates the 100th anniversary of the development of Quantum Mechanics. We would like to dedicate it to our colleague and friend Dr Carlo Gatti on the occasion of his 70th birthday, to recognise the passion for research he has constantly transmitted to his co-workers.

A detailed study of the X⋯N (X = I, Br) halogen bonds in complexes formed by an extended set of substituted pyridines with DX molecules (D = X, CN) is reported here. The nature of these interactions has been investigated at different (MP2 and DFT) levels of theory through Bader's quantum theory of atoms in molecules (QTAIM) and Pendás' interacting quantum atoms (IQA) scheme, focusing on the role of the local environment (i.e. the substituent on the pyridine ring and the halogenated residue) on the halogen bond features. We found that the exchange-correlation energy represents a substantial contribution to the IQA total energy, in some cases comparable to (I2 complexes) or even dominating (ICN complexes) the electrostatic term. Meaningful information is provided by the source function, indicating that the major contribution to the electron density at the bond critical point of the X⋯N interaction is derived from the halogen atom, while a much lower contribution comes from the nitro­gen atom, which acts as either source or sink for electron density. A relevant contribution from distal atoms, including the various electron-donor and electron-withdrawing substituents in different positions of the pyridine ring, is also determined, highlighting the non-local character of the electron density. The existence of possible relationships between binding energies, interaction energies according to IQA, and QTAIM descriptors such as delocalization indices and source function, has been inspected. In general, good correlations are only found when the local environment, external to the directly involved halogen and nitro­gen atoms, plays a minor role in the interaction.

1. Introduction

Non-covalent interactions (NCIs) represent a subject of ever-growing interest not only in view of the synthesis and characterization of new functional materials (Haque et al., 2023[Haque, A., Alenezi, K. M., Khan, M. S., Wong, W.-Y. & Raithby, P. R. (2023). Chem. Soc. Rev. 52, 454-472.]; Molina et al., 2017[Molina, P., Zapata, F. & Caballero, A. (2017). Chem. Rev. 117, 9907-9972.]; Xiao & Fu, 2019[Xiao, L. & Fu, H. (2019). Chem. A Eur. J. 25, 714-723.]; Bulfield & Huber, 2016[Bulfield, D. & Huber, S. M. (2016). Chem. A Eur. J. 22, 14434-14450.]; Moghadasnia et al., 2024[Moghadasnia, M. P., Eckstein, B. J., Martin, H. R., Paredes, J. U. & McGuirk, C. M. (2024). Cryst. Growth Des. 24, 2304-2321.]) as well as understanding biomolecular structure and function (Jena et al., 2022[Jena, S., Dutta, J., Tulsiyan, K. D., Sahu, A. K., Choudhury, S. S. & Biswal, H. S. (2022). Chem. Soc. Rev. 51, 4261-4286.]; Xu et al., 2014[Xu, Z., Yang, Z., Liu, Y., Lu, Y., Chen, K. & Zhu, W. (2014). J. Chem. Inf. Model. 54, 69-78.]; Kojasoy & Tantillo, 2022[Kojasoy, V. & Tantillo, D. J. (2022). Org. Biomol. Chem. 21, 11-23.]; Walker et al., 2023[Walker, M. G., Mendez, C. G. & Ho, P. S. (2023). Chem. Asian J. 18, e202300026.]; Verteramo et al., 2024[Verteramo, M. L., Ignjatović, M. M., Kumar, R., Wernersson, S., Ekberg, V., Wallerstein, J., Carlström, G., Chadimová, V., Leffler, H., Zetterberg, F., Logan, D. T., Ryde, U., Akke, M. & Nilsson, U. J. (2024). iScience 27, 109636.]), but also for the increasingly in-depth knowledge gained through theoretical and computational investigations (Brammer et al., 2023[Brammer, L., Peuronen, A. & Roseveare, T. M. (2023). Acta Cryst. C79, 204-216.]; Kolář & Hobza, 2016[Kolář, M. H. & Hobza, P. (2016). Chem. Rev. 116, 5155-5187.]; Wolters et al., 2014[Wolters, L. P., Schyman, P., Pavan, M. J., Jorgensen, W. L., Bickelhaupt, F. M. & Kozuch, S. (2014). WIREs Comput. Mol. Sci. 4, 523-540.]; Phan Dang et al., 2023[Phan Dang, C.-T., Tam, N. M., Huynh, T.-N. & Trung, N. T. (2023). RSC Adv. 13, 31507-31517.]; Scheiner et al., 2020[Scheiner, S., Michalczyk, M. & Zierkiewicz, W. (2020). Coord. Chem. Rev. 405, 213136.]; Grabowski, 2021[Grabowski, S. J. (2021). Molecules, 26, 4939.]), where supramolecular systems often represent a test bench for assessing new methods and protocols (Jiménez-Grávalos et al., 2021[Jiménez-Grávalos, F., Gallegos, M., Martín Pendás, Á. & Novikov, A. S. (2021). J. Comput. Chem. 42, 676-687.]; Guerra et al., 2024[Guerra, C., Burgos, J., Ayarde-Henríquez, L. & Chamorro, E. (2024). J. Phys. Chem. A, 128, 6158-6166.]; Forni et al., 2016[Forni, A., Pieraccini, S., Franchini, D. & Sironi, M. (2016). J. Phys. Chem. A, 120, 9071-9080.], 2014[Forni, A., Pieraccini, S., Rendine, S. & Sironi, M. (2014). J. Comput. Chem. 35, 386-394.], 2012[Forni, A., Pieraccini, S., Rendine, S., Gabas, F. & Sironi, M. (2012). ChemPhysChem, 13, 4224-4234.]). The fields of crystal and cocrystal engineering (Frontera & Bauzá, 2021[Frontera, A. & Bauzá, A. (2021). Crystals (Basel), 11, 1205.]; Nemec et al., 2021[Nemec, V., Lisac, K., Bedeković, N., Fotović, L., Stilinović, V. & Cinčić, D. (2021). CrystEngComm, 23, 3063-3083.]; Saha et al., 2018[Saha, S., Mishra, M. K., Reddy, C. M. & Desiraju, G. R. (2018). Acc. Chem. Res. 51, 2957-2967.]; Hajji et al., 2023[Hajji, M., Haouas, A., Abad, N. & Guerfel, T. (2023). ChemistrySelect, 8, e202302624.]; Siddiqui et al., 2024[Siddiqui, R., Rani, J., Titi, H. M. & Patra, R. (2024). Coord. Chem. Rev. 517, 215994.]), eventually supplemented by experimental charge density studies (Forni et al., 2019[Forni, A., Franchini, D., Dapiaggi, F., Pieraccini, S., Sironi, M., Scilabra, P., Pilati, T., Petko, K. I., Resnati, G. & Yagupolkii, Y. L. (2019). Cryst. Growth Des. 19, 1621-1631.]; Otte et al., 2021[Otte, F., Kleinheider, J., Hiller, W., Wang, R., Englert, U. & Strohmann, C. (2021). J. Am. Chem. Soc. 143, 4133-4137.]; Eraković et al., 2019[Eraković, M., Cinčić, D., Molčanov, K. & Stilinović, V. (2019). Angew. Chem. Int. Ed. 58, 15702-15706.]; Shukla et al., 2018[Shukla, R., Claiser, N., Souhassou, M., Lecomte, C., Balkrishna, S. J., Kumar, S. & Chopra, D. (2018). IUCrJ, 5, 647-653.]; Thomas et al., 2022[Thomas, S. P., Dikundwar, A. G., Sarkar, S., Pavan, M. S., Pal, R., Hathwar, V. R. & Row, T. N. G. (2022). Molecules, 27, 3690.]), have greatly contributed to our knowledge and the potential of NCIs. Owing to the increasing body of structural and theoretical information, apparently well assessed definitions and concepts inherent to consolidated interactions, such as hydrogen bonds (Arunan et al., 2011[Arunan, E., Desiraju, G. R., Klein, R. A., Sadlej, J., Scheiner, S., Alkorta, I., Clary, D. C., Crabtree, R. H., Dannenberg, J. J., Hobza, P., Kjaergaard, H. G., Legon, A. C., Mennucci, B. & Nesbitt, D. J. (2011). Pure Appl. Chem. 83, 1637-1641. ]) and halogen bonds (XBs) (Desiraju et al., 2013[Desiraju, G. R., Ho, P. S., Kloo, L., Legon, A. C., Marquardt, R., Metrangolo, P., Politzer, P., Resnati, G. & Rissanen, K. (2013). Pure Appl. Chem. 85, 1711-1713.]), need to be continuously upgraded and/or revisited (Grabowski, 2024[Grabowski, S. J. (2024). Chem. Commun. 60, 6239-6255.]; Scheiner, 2023[Scheiner, S. (2023). Phys. Chem. Chem. Phys. 25, 7184-7194.]; Varadwaj et al., 2024[Varadwaj, P. R., Varadwaj, A., Marques, H. M. & Yamashita, K. (2024). Cryst. Growth Des. 24, 5494-5525.]).

Among NCIs, XBs represent, together with hydrogen bonds, the most widely studied interactions (Metrangolo & Resnati, 2012[Metrangolo, P. & Resnati, G. (2012). Cryst. Growth Des. 12, 5835-5838.], 2008[Metrangolo, P. & Resnati, G. (2008). Science, 321, 918-919.]; Fourmigué, 2009[Fourmigué, M. (2009). Curr. Opin. Solid State Mater. Sci. 13, 36-45.]; Politzer & Murray, 2013[Politzer, P. & Murray, J. S. (2013). ChemPhysChem, 14, 278-294.]; Cavallo et al., 2016[Cavallo, G., Metrangolo, P., Milani, R., Pilati, T., Priimagi, A., Resnati, G. & Terraneo, G. (2016). Chem. Rev. 116, 2478-2601.]), as testified by applications in different fields, spanning from biological systems (Parisini et al., 2011[Parisini, E., Metrangolo, P., Pilati, T., Resnati, G. & Terraneo, G. (2011). Chem. Soc. Rev. 40, 2267-2278.]; Wilcken et al., 2013[Wilcken, R., Zimmermann, M. O., Lange, A., Joerger, A. C. & Boeckler, F. M. (2013). J. Med. Chem. 56, 1363-1388.]; Vargas Jentzsch & Matile, 2013[Vargas Jentzsch, A. & Matile, S. (2013). J. Am. Chem. Soc. 135, 5302-5303.]; Auffinger et al., 2004[Auffinger, P., Hays, F. A., Westhof, E. & Ho, P. S. (2004). Proc. Natl Acad. Sci. USA, 101, 16789-16794.]; Lu et al., 2010[Lu, Y., Wang, Y. & Zhu, W. (2010). Phys. Chem. Chem. Phys. 12, 4543-4551.]; Cavallo et al., 2016[Cavallo, G., Metrangolo, P., Milani, R., Pilati, T., Priimagi, A., Resnati, G. & Terraneo, G. (2016). Chem. Rev. 116, 2478-2601.]) to drug design (Jiang et al., 2006[Jiang, Y., Alcaraz, A. A., Chen, J.-M., Kobayashi, H., Lu, Y. J. & Snyder, J. P. (2006). J. Med. Chem. 49, 1891-1899.]; Bissantz et al., 2010[Bissantz, C., Kuhn, B. & Stahl, M. (2010). J. Med. Chem. 53, 5061-5084.]; Lu et al., 2009[Lu, Y., Shi, T., Wang, Y., Yang, H., Yan, X., Luo, X., Jiang, H. & Zhu, W. (2009). J. Med. Chem. 52, 2854-2862.]), and the development of smart materials such as polymers (Lauher et al., 2008[Lauher, J. W., Fowler, F. W. & Goroff, N. S. (2008). Acc. Chem. Res. 41, 1215-1229.]), liquid crystals (Nguyen et al., 2004[Nguyen, H. L., Horton, P. N., Hursthouse, M. B., Legon, A. C. & Bruce, D. W. (2004). J. Am. Chem. Soc. 126, 16-17.]), solid-state materials with peculiar electronic properties including superconductors (Kato et al., 2002[Kato, R., Imakubo, T., Yamamoto, H., Maeda, R., Fujiwara, M., Yamaura, J.-I. & Sawa, H. (2002). Mol. Cryst. Liq. Cryst. 380, 61-68.]), and porous (Metrangolo et al., 2009[Metrangolo, P., Carcenac, Y., Lahtinen, M., Pilati, T., Rissanen, K., Vij, A. & Resnati, G. (2009). Science, 323, 1461-1464.]) and phospho­rescent organic materials (Bolton et al., 2011[Bolton, O., Lee, K., Kim, H.-J., Lin, K. Y. & Kim, J. (2011). Nat. Chem. 3, 415.]).

An XB is the interaction between a covalently bonded halogen atom X and a nucleophilic acceptor A, according to the scheme DXA. The nucleophilic site is usually a lone pair on a heteroatom, such as oxygen (Kosmas, 2007[Kosmas, A. M. (2007). Bioinorg. Chem. Appl. 2007, 046393.]; Kratzer et al., 2015[Kratzer, P., Ramming, B., Römisch, S. & Maas, G. (2015). Cryst­EngComm, 17, 4486-4494.]; Nelyubina et al., 2011[Nelyubina, Y., Antipin, M. & Lyssenko, K. (2011). Mendeleev Commun. 21, 250-252.]), nitro­gen (Hakkert & Erdélyi, 2015[Hakkert, S. B. & Erdélyi, M. (2015). J. Phys. Org. Chem. 28, 226-233.]; Bartashevich et al., 2014[Bartashevich, E. V., Troitskaya, E. A. & Tsirelson, V. G. (2014). Chem. Phys. Lett. 601, 144-148.]) or sulfur (Ford et al., 2017[Ford, M. C., Saxton, M. & Ho, P. S. (2017). J. Phys. Chem. Lett. 8, 4246-4252.]; Eccles et al., 2014[Eccles, K., Morrison, R., Sinha, A., Maguire, A. & Lawrence, S. (2014). Acta Cryst. A70, C651.]; Hauchecorne et al., 2011[Hauchecorne, D., Moiana, A., van der Veken, B. J. & Herrebout, W. A. (2011). Phys. Chem. Chem. Phys. 13, 10204-10213.]); an anion (Decato et al., 2021[Decato, D. A., Riel, A. M. S., May, J. H., Bryantsev, V. S. & Berryman, O. B. (2021). Angew. Chem. Int. Ed. 60, 3685-3692.]; Fotović et al., 2023[Fotović, L., Bedeković, N. & Stilinović, V. (2023). Cryst. Growth Des. 23, 3384-3392.]; Mínguez Espallargas et al., 2009[Mínguez Espallargas, G., Zordan, F., Arroyo Marín, L., Adams, H., Shankland, K., van de Streek, J. & Brammer, L. (2009). Chem. A Eur. J. 15, 7554-7568.], 2006[Mínguez Espallargas, G., Brammer, L. & Sherwood, P. (2006). Angew. Chem. Int. Ed. 45, 435-440.]); or a π-system such as a benzene ring (Luo et al., 2022[Luo, J., Dai, H., Zeng, C., Wu, D. & Cao, M. (2022). Molecules, 27. https://doi.org/10.3390/molecules27228078.]; Forni et al., 2016[Forni, A., Pieraccini, S., Franchini, D. & Sironi, M. (2016). J. Phys. Chem. A, 120, 9071-9080.], 2014[Forni, A., Pieraccini, S., Franchini, D. & Sironi, M. (2016). J. Phys. Chem. A, 120, 9071-9080.], 2012[Forni, A., Pieraccini, S., Rendine, S., Gabas, F. & Sironi, M. (2012). ChemPhysChem, 13, 4224-4234.]). Moiety D may be both organic or inorganic in nature and diverse in size, ranging from another halogen atom (Schneider et al., 2017[Schneider, F., Caramori, G., Parreira, R., Lippolis, V., Arca, M. & Ciancaleoni, G. (2017). Eur. J. Inorg. Chem. 2018, 1007-1015.]; Alkorta et al., 1998[Alkorta, I., Rozas, I. & Elguero, J. (1998). J. Phys. Chem. A, 102, 9278-9285.]) to large biological residues (Borozan & Stojanović, 2013[Borozan, S. Z. & Stojanović, S. Đ. (2013). Comput. Biol. Chem. 47, 231-239.]; Scholfield et al., 2013[Scholfield, M. R., Zanden, C. M., Carter, M. & Ho, P. S. (2013). Protein Sci. 22, 139-152.]).

The ability of a highly electronegative atom to interact with a nucleophilic site was first rationalized by Mulliken (1950[Mulliken, R. S. (1950). J. Am. Chem. Soc. 72, 600-608.]). He interpreted an XB as deriving from the orbital interaction between the lone pair on A and the DX antibonding orbital, thus assigning charge transfer (CT) character to the interaction, in addition to some degree of covalency. This model was subsequently confirmed by hybrid valence bond/molecular orbital (VB/MO) calculations, based on the block-localized wavefunction (BLW) (Mo et al., 2007[Mo, Y., Song, L. & Lin, Y. (2007). J. Phys. Chem. A, 111, 8291-8301.], 2011[Mo, Y., Bao, P. & Gao, J. (2011). Phys. Chem. Chem. Phys. 13, 6760-6775.]) method. BLW calculations on a series of XB complexes demonstrated that, except for weak interactions dominated by dispersion forces, CT is non-negligible or even the dominant contribution of most XBs (Wang et al., 2014[Wang, C., Danovich, D., Mo, Y. & Shaik, S. (2014). J. Chem. Theory Comput. 10, 3726-3737.]).

Much later, Politzer et al. (2010[Politzer, P., Murray, J. S. & Clark, T. (2010). Phys. Chem. Chem. Phys. 12, 7748-7757.], 2007[Politzer, P., Lane, P., Concha, M. C., Ma, Y. & Murray, J. S. (2007). J. Mol. Model. 13, 305-311.]) explained that an XB can be attributed to the anisotropic electron density distribution around a covalently bonded halogen atom, whose valence electronic structure can be roughly described as ns2 npx2 npy2 npz1, z being the direction of the bond. In agreement with such description, a concentration of electron density is found in the region orthogonal to the bond direction, and a depletion of electron density is present along the extension of the DX bond. According to this model, the configuration of the bonded halogen explains the formation, in the electrostatic potential (ESP), of a positive region along the extension of the DX bond, called a σ-hole, and a negative region around it (Politzer et al., 2007[Politzer, P., Lane, P., Concha, M. C., Ma, Y. & Murray, J. S. (2007). J. Mol. Model. 13, 305-311.], 2013[Politzer, P., Murray, J. S. & Clark, T. (2013). Phys. Chem. Chem. Phys. 15, 11178-11189.]; Murray et al., 2009[Murray, J. S., Lane, P. & Politzer, P. (2009). J. Mol. Model. 15, 723-729.]; Clark et al., 2007[Clark, T., Hennemann, M., Murray, J. S. & Politzer, P. (2007). J. Mol. Model. 13, 291-296.]; Clark, 2013[Clark, T. (2013). WIREs Comput. Mol. Sci. 3, 13-20.]; Kolář & Hobza, 2016[Kolář, M. H. & Hobza, P. (2016). Chem. Rev. 116, 5155-5187.]). More recently, the existence of the σ-hole has been proven on a rigorous ab initio basis by means of spin-coupled (SC) VB calculations, and associated with a contraction of the SC orbitals describing the pz lone pair, while the negative belt around the halogen atom, observed only when bonded to electron-withdrawing groups of medium strength, is to be ascribed to a reduced contraction of the SC orbitals corresponding to the px and py lone pairs (Forni et al., 2024[Forni, A., Genoni, A., Pieraccini, S. & Sironi, M. (2024). Comprehensive Computational Chemistry, 1st ed., edited by M. Yáñez & R. J. Boyd, pp. 533-551. Oxford: Elsevier.]; Franchini et al., 2020[Franchini, D., Forni, A., Genoni, A., Pieraccini, S., Gandini, E. & Sironi, M. (2020). ChemistryOpen, 9, 445-450.]). The features of the σ-hole depend on both the nature of the halogen atom [i.e. its polarizability and electronegativity (Messina et al., 1998[Messina, M. T., Metrangolo, P., Panzeri, W., Ragg, E. & Resnati, G. (1998). Tetrahedron Lett. 39, 9069-9072.]; Metrangolo et al., 2002[Metrangolo, P., Panzeri, W., Recupero, F. & Resnati, G. (2002). J. Fluor. Chem. 114, 27-33.]; Politzer et al., 2010[Politzer, P., Murray, J. S. & Clark, T. (2010). Phys. Chem. Chem. Phys. 12, 7748-7757.]; Clark et al., 2007[Clark, T., Hennemann, M., Murray, J. S. & Politzer, P. (2007). J. Mol. Model. 13, 291-296.])] and the chemical environment, in particular the electron-withdrawing ability of D and the charge-donor propensity of A (Lo et al., 2012[Lo, R., Ballabh, A., Singh, A., Dastidar, P. & Ganguly, B. (2012). CrystEngComm, 14, 1833-1841.]; Riley et al., 2011[Riley, K. E., Murray, J. S., Fanfrlík, J., Řezáč, J., Solá, R. J., Concha, M. C., Ramos, F. M. & Politzer, P. (2011). J. Mol. Model. 17, 3309-3318.]; Kolář et al., 2014[Kolář, M., Hostaš, J. & Hobza, P. (2014). Phys. Chem. Chem. Phys. 16, 9987-9996.]). The strength of XB interaction has been proven to correlate with the ESP value on the σ-hole (Politzer et al., 2007[Politzer, P., Lane, P., Concha, M. C., Ma, Y. & Murray, J. S. (2007). J. Mol. Model. 13, 305-311.]; Murray et al., 2009[Murray, J. S., Lane, P. & Politzer, P. (2009). J. Mol. Model. 15, 723-729.]; Riley et al., 2009[Riley, K. E., Murray, J. S., Politzer, P., Concha, M. C. & Hobza, P. (2009). J. Chem. Theory Comput. 5, 155-163.], 2011[Riley, K. E., Murray, J. S., Fanfrlík, J., Řezáč, J., Solá, R. J., Concha, M. C., Ramos, F. M. & Politzer, P. (2011). J. Mol. Model. 17, 3309-3318.]). According to SCVB calculations, different descriptors associated with the interaction, such as the overlap between the involved SC orbitals, their shapes and the Chirgwin–Coulson weights of the SC structures point to a VB picture of halogen bonding as due to a shift/delocalization of the Lewis base lone pair towards the halogen atom (Franchini et al., 2019[Franchini, D., Genoni, A., Dapiaggi, F., Pieraccini, S. & Sironi, M. (2019). Int. J. Quantum Chem. 119, e25946.]; Forni et al., 2024[Forni, A., Genoni, A., Pieraccini, S. & Sironi, M. (2024). Comprehensive Computational Chemistry, 1st ed., edited by M. Yáñez & R. J. Boyd, pp. 533-551. Oxford: Elsevier.]). Moreover, it was reported that one of the electrons of the Lewis base lone pair is localized on the halogen atom in the direction pointing towards the D group, highlighting the importance of the CT contribution to the interaction, in agreement with the Mulliken model and BLW calculations (Wang et al., 2014[Wang, C., Danovich, D., Mo, Y. & Shaik, S. (2014). J. Chem. Theory Comput. 10, 3726-3737.]).

Most computational studies on halogen bonding, and in general on NCIs, are focused on global quantities, such as interaction energies and their decompositions, dipole moments, CT and so on, while very little is known about atomic contributions, in particular those coming from the donor and acceptor atoms directly involved, to determine these quantities. Such knowledge would provide important insights into the tunability of NCIs.

In this context, invaluable information can be acquired through the quantum theory of atoms in molecules (QTAIM) (Bader, 1991[Bader, R. F. W. (1991). Chem. Rev. 91, 893-928.], 1990[Bader, R. F. W. (1990). Atoms in Molecules: a Quantum Theory. International Series of Monographs on Chemistry. Clarendon Press.]; Popelier et al., 2000[Popelier, P., Aicken, F. & O'Brien, S. (2000). Chemical Modelling: Applications and Theory, Vol. 1, pp. 143-198.]; Popelier, 2000[Popelier, P. L. A. (2000). Atoms in Molecules: an Introduction. London: Prentice Hall.]) and the related interacting quantum atoms (IQA) approaches (Blanco et al., 2005[Blanco, M. A., Martín Pendás, A. & Francisco, E. (2005). J. Chem. Theory Comput. 1, 1096-1109.]; Francisco et al., 2006[Francisco, E., Martín Pendás, A. & Blanco, M. A. (2006). J. Chem. Theory Comput. 2, 90-102.]; Guevara-Vela et al., 2020[Guevara-Vela, J. M., Francisco, E., Rocha-Rinza, T. & Martín Pendás, Á. (2020). Molecules, 25, 4028.]), which provide a partitioning scheme for the charge distribution of a system (and related properties) and its total energy, respectively, allowing us to gain insights into the specific role of interacting atoms in the formation of an XB (Alkorta et al., 2020[Alkorta, I., Silva, A. F. & Popelier, P. L. A. (2020). Molecules, 25, 2674.]). Within QTAIM, a powerful tool is provided by the source function (SF) concept (Bader, 1990[Bader, R. F. W. (1990). Atoms in Molecules: a Quantum Theory. International Series of Monographs on Chemistry. Clarendon Press.]; Bader & Gatti, 1998[Bader, R. F. W. & Gatti, C. (1998). Chem. Phys. Lett. 287, 233-238.]; Gatti, Saleh & Lo Presti, 2016[Gatti, C., Saleh, G. & Lo Presti, L. (2016). Acta Cryst. B72, 180-193.]; Tantardini et al., 2016[Tantardini, C., Ceresoli, D. & Benassi, E. (2016). J. Comput. Chem. 37, 2133-2139.]), describing any local value of a scalar function (e.g. the electron density) as due to the contributions from all other points of the molecular space, therefore providing valuable chemical insights into covalent and non-covalent bonding.

In this work, we report the results of an investigation on the X⋯N XB performed through QTAIM, SF and IQA approaches. Inspired by the study of Bartashevich et al. (2014[Bartashevich, E. V., Troitskaya, E. A. & Tsirelson, V. G. (2014). Chem. Phys. Lett. 601, 144-148.]) on the I⋯N XB in substituted pyridines, we considered, for XB acceptors, the same set of pyridines, while for XB donors we examined, aside from the previously investigated I2 case, Br2, ICN and BrCN molecules, to evaluate the effect of the nature of both the halogen and the attached substituent on the XB interaction. Several relationships between binding energies, interaction energies (according to the IQA scheme) and QTAIM descriptors have been examined, based on different (MP2 and DFT) levels of theory.

2. Methods

According to QTAIM, atoms are viewed as disjointed and exhaustive regions of space, or basins (Ω), bound by zero-flux surfaces of the electron density gradient. The interatomic surfaces, for each point r on the surface S(r), follow the relation

[\nabla \rho({\bf r}) \cdot n({\bf r}) = 0, \eqno (1)]

ρ(r) being the gradient of the electron density, ρ(r); and n(r) being the unitary vector normal to the surface at point r. The points where the first derivatives of ρ(r) vanish are defined as critical points, and the lines of maximum electron density connecting two critical points are known as bond paths.

The electron-pair sharing between two basins ΩA and ΩB connected by a bond path can be described in terms of the delocalization index (Bader & Stephens, 1975[Bader, R. F. W. & Stephens, M. E. (1975). J. Am. Chem. Soc. 97, 7391-7399.]; Fradera et al., 1999[Fradera, X., Austen, M. A. & Bader, R. F. W. (1999). J. Phys. Chem. A, 103, 304-314.]; Bader et al., 1996[Bader, R. F. W., Streitwieser, A., Neuhaus, A., Laidig, K. E. & Speers, P. (1996). J. Am. Chem. Soc. 118, 4959-4965.]):

[\delta \left(\Omega_A,\Omega_B \right) = - 2\mathop \int \limits_{\Omega_A} {\rm d}{{\bf r}_{1}}\mathop \int \limits_{\Omega_B} {\rm d}{{\bf r}_2}\left [{\rho \left({{{\bf r}_{1}},{{\bf r}_{2}}} \right) - \rho \left({{{\bf r}_1}} \right)\rho \left({{{\bf r}_2}} \right)} \right], \eqno (2)]

where ρ(r1, r2) is the pair density (i.e. the probability density of finding a pair of electrons at the volume elements dr1 and dr2); and [\rho({\bf r}_1)\rho ({\bf r}_2)] is the uncorrelated component of the pair density, which provides the probability of concurrently finding two independent electrons in positions r1 and r2. The delocalization index is therefore associated with the magnitude of the exchange of the electrons in the basin of atom A with those in the basin of atom B. This has been demonstrated to reflect the bond order (Outeiral et al., 2018[Outeiral, C., Vincent, M. A., Martín Pendás, Á. & Popelier, P. L. A. (2018). Chem. Sci. 9, 5517-5529.]).

A useful tool to quantify how distant atomic interactions affect the delocalization between the two interacting basins is the SF (Bader, 1990[Bader, R. F. W. (1990). Atoms in Molecules: a Quantum Theory. International Series of Monographs on Chemistry. Clarendon Press.]; Bader & Essén, 1984[Bader, R. F. W. & Essén, H. (1984). J. Chem. Phys. 80, 1943-1960.]; Bader & Gatti, 1998[Bader, R. F. W. & Gatti, C. (1998). Chem. Phys. Lett. 287, 233-238.]; Gatti, Saleh & Lo Presti, 2016[Gatti, C., Saleh, G. & Lo Presti, L. (2016). Acta Cryst. B72, 180-193.]; Tantardini et al., 2016[Tantardini, C., Ceresoli, D. & Benassi, E. (2016). J. Comput. Chem. 37, 2133-2139.]). This function is derived from the consideration that the total electron density at any reference point r can be seen as determined by local contributions from each point r′ in space, according to

[\rho ({\bf r}) = - {1 \over 4} \int {{\nabla^2\rho({\bf r}')} \over {|{\bf r} - {\bf r}'|}}\,{\rm d}{\bf r}'. \eqno (3)]

Partitioning the whole space into atomic basins allows us to replace the integration in equation (3)[link] with a sum of integrations over atomic basins Ωi, each of them providing the contribution SF(r, Ωi) to the total electron density deriving from that atom:

[\rho ({\bf r}) = - {1 \over 4}\mathop \sum \limits_i \mathop \int \limits_{{\Omega _i}} {{{\nabla ^2}\rho \left({{\bf r}'} \right)} \over {\left| {{\bf r} - {\bf r}'} \right|}}\,{\rm d}{\bf r}' = \mathop \sum \limits_i {\rm SF}\left({{\bf r},{\Omega _{i}}} \right). \eqno (4)]

An atomic basin can yield a positive or a negative contribution to ρ(r), therefore behaving as either a `source' or a `sink' of electron density, respectively.

Starting from QTAIM, the IQA (Blanco et al., 2005[Blanco, M. A., Martín Pendás, A. & Francisco, E. (2005). J. Chem. Theory Comput. 1, 1096-1109.]; Francisco et al., 2006[Francisco, E., Martín Pendás, A. & Blanco, M. A. (2006). J. Chem. Theory Comput. 2, 90-102.]; Guevara-Vela et al., 2020[Guevara-Vela, J. M., Francisco, E., Rocha-Rinza, T. & Martín Pendás, Á. (2020). Molecules, 25, 4028.]) provides a general energy partition scheme to decompose the total energy of a system into atomic and interatomic contributions:

[E_{\rm IQA} = \mathop \sum \limits_A E_{\rm IQA}^{{\rm Intra}(A)} + \mathop \sum \limits_{A \gt B} E_{\rm IQA}^{{\rm Inter}(A,B)}. \eqno (5)]

Atomic contributions account for the kinetic energy and the electron–nucleus and electron–electron interactions for particles belonging to the same atom:

[E_{\rm IQA}^{{\rm Intra}(A)} = {T^A} + V_{\rm en}^{AA} + V_{\rm ee}^{AA}. \eqno (6)]

Interatomic energy, on the other hand, includes all the interactions between each atom pair (i.e. nucleus–nucleus, nucleus–electron and electron–electron interactions):

[E_{\rm IQA}^{{\rm Inter}(A,B)} = V_{\rm ne}^{AB} + V_{\rm en}^{AB} + V_{\rm ee}^{AB} + V_{\rm nn}^{AB}, \eqno (7)]

where the electron–electron term comprises a Coulomb part [{\rm VC}_{\rm IQA}^{{\rm Inter}(A,B)}] and an exchange-correlation part [{\rm VX}_{\rm IQA}^{{\rm Inter}(A,B)}]. The interatomic term [E_{\rm IQA}^{{\rm Inter}(A,B)}] therefore represents only a contribution to the total binding energy, [E_{\rm IQA}^{\rm BIND}], between two interacting molecular fragments F1 and F2. This latter term, in fact, accounts not only for all the interactions between atom pairs belonging to each molecular fragment, [E_{\rm IQA}^{{\rm Inter}({\rm F}1\cdot\cdot\cdot{\rm F}2)}], but also for both atomic, [\Delta E_{\rm IQA}^{{\rm Intra}({\rm F})}], and interatomic, [\Delta E_{\rm IQA}^{{\rm Inter}({\rm F})}], energy variations upon complexation inside each fragment (Syzgantseva et al., 2013[Syzgantseva, O. A., Tognetti, V. & Joubert, L. (2013). J. Phys. Chem. A, 117, 8969-8980.]):

[ \eqalign { E_{\rm IQA}^{\rm BIND} = & \ E_{\rm IQA}^{{\rm Inter}({\rm F}1\cdot \cdot \cdot {\rm F}2)} + \Delta E_{\rm IQA}^{{\rm Intra}({\rm F}1)} + \Delta E_{\rm IQA}^{{\rm Intra}({\rm F}2)} + \Delta E_{\rm IQA}^{{\rm Inter}({\rm F}1)} \cr & + \Delta E_{\rm IQA}^{{\rm Inter}({\rm F}2)}. }\eqno (8)]

Based on these formulae, IQA provides both the `two-centre' XB energy [i.e. the interaction energy between the XB donor (X) and acceptor (A)], [E_{\rm IQA}^{{\rm Inter}(X,A)}], and the `total' XB energy, [E_{\rm IQA}^{\rm BIND}]. The latter is expected to recover the `conventional' binding energy, [E^{\rm BIND}], computed as the difference between the energies of the complex and the isolated fragments, in particular when dispersion contributions are small (Syzgantseva et al., 2013[Syzgantseva, O. A., Tognetti, V. & Joubert, L. (2013). J. Phys. Chem. A, 117, 8969-8980.]).

3. Computational details

Ab initio calculations were performed using the Gaussian16 software (Frisch et al., 2016[Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Petersson, G. A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A. V., Bloino, J., Janesko, B. G., Gomperts, R., Mennucci, B., Hratchian, H. P., Ortiz, J. V., Izmaylov, A. F., Sonnenberg, J. L., Williams-Young, D., Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone, A., Henderson, T., Ranasinghe, D., Zakrzewski, V. G., Gao, J., Rega, N., Zheng, G., Liang, W., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M. J., Heyd, J. J., Brothers, E. N., Kudin, K. N., Staroverov, V. N., Keith, T. A., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A. P., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Millam, J. M., Klene, M., Adamo, C., Cammi, R., Ochterski, J. W., Martin, R. L., Morokuma, K., Farkas, O., Foresman, J. B. & Fox, D. J. (2016). Gaussian 16. Revision C.01. Gaussian Inc., Wallingford, Connecticut, USA.]). The systems were investigated at both MP2 (Head-Gordon & Head-Gordon, 1994[Head-Gordon, M. & Head-Gordon, T. (1994). Chem. Phys. Lett. 220, 122-128.]; Saebø & Almlöf, 1989[Saebø, S. & Almlöf, J. (1989). Chem. Phys. Lett. 154, 83-89.]; Head-Gordon et al., 1988[Head-Gordon, M., Pople, J. A. & Frisch, M. J. (1988). Chem. Phys. Lett. 153, 503-506.]; Frisch et al., 1990a[Frisch, M. J., Head-Gordon, M. & Pople, J. A. (1990a). Chem. Phys. Lett. 166, 275-280.],b[Frisch, M. J., Head-Gordon, M. & Pople, J. A. (1990b). Chem. Phys. Lett. 166, 281-289.]) and DFT levels of theory, employing the ubiquitously used B3LYP functional (Vosko et al., 1980[Vosko, S. H., Wilk, L. & Nusair, M. (1980). Can. J. Phys. 58, 1200-1211.]; Lee et al., 1988[Lee, C., Yang, W. & Parr, R. G. (1988). Phys. Rev. B, 37, 785-789.]; Becke, 1993[Becke, A. D. (1993). J. Chem. Phys. 98, 5648-5652.]) and four other functionals: M06-2X (Zhao & Truhlar, 2008[Zhao, Y. & Truhlar, D. G. (2008). Theor. Chem. Acc. 120, 215-241.]), M11 (Peverati & Truhlar, 2011[Peverati, R. & Truhlar, D. G. (2011). J. Phys. Chem. Lett. 2, 2810-2817.]), ωB97X (Chai & Head-Gordon, 2008b[Chai, J.-D. & Head-Gordon, M. (2008b). J. Chem. Phys. 128, 084106.]) and ωB97XD (Chai & Head-Gordon, 2008a[Chai, J.-D. & Head-Gordon, M. (2008a). Phys. Chem. Chem. Phys. 10, 6615-6620.]), which gave a better performance according to our previous QTAIM investigation on XBs (Forni et al., 2016[Forni, A., Pieraccini, S., Franchini, D. & Sironi, M. (2016). J. Phys. Chem. A, 120, 9071-9080.]). All calculations were performed with the 6-311++G(d,p) basis set. For the iodine atom, this basis set, not internally stored in Gaussian16, was built up by downloading the 6-311G(d,p) one from the basis set exchange site (Feller, 1996[Feller, D. (1996). J. Comput. Chem. 17, 1571-1586.]; Schuchardt et al., 2007[Schuchardt, K. L., Didier, B. T., Elsethagen, T., Sun, L., Gurumoorthi, V., Chase, J., Li, J. & Windus, T. L. (2007). J. Chem. Inf. Model. 47, 1045-1052.]) and retrieving the diffuse functions from the literature (Glukhovtsev et al., 1995[Glukhovtsev, M. N., Pross, A., McGrath, M. P. & Radom, L. (1995). J. Chem. Phys. 103, 1878-1885. ]).

Geometry optimizations were carried out at the DFT level using an ultrafine grid with 99 radial shells and 590 angular points per shell. All minima have been confirmed by vibrational frequency analysis. Due to the high computational cost of the MP2 method, we have used the complexes at fixed geometry obtained at the B3LYP/6-1++G(d,p) level. On each minimum-energy structure, for every above-mentioned method, a single-point calculation was performed to introduce the counterpoise correction and obtain the binding energy, EBIND, computed as the difference in energy between the complex and the isolated non-relaxed fragments.

The wavefunctions obtained by optimization at the DFT level, and by single-point calculations without the counterpoise correction at the MP2 level, have been used for QTAIM and IQA analyses by means of the AIMAll software (Todd, 2019[Todd, A. K. (2019). AIMAll, version 19.10.12. TK Gristmill Software, Overland Park KS, USA. https://aim.tkgristmill.com.]). In particular, we computed the electron density values, ρ(rBCP), and the atomic contributions to the source function, SF(Ω), at the bond critical point (BCP) of the X⋯N XB, the latter quantity also expressed as percentage contributions, SF%(Ω), to the total density in the reference point, ρ(rBCP). The delocalization indices δ(ΩX, ΩN) have also been evaluated.

The two-centre XB energy, [E_{\rm IQA}^{{\rm Inter}(X,{\rm N})}], and its exchange-correlation contribution, [{\rm VX}_{\rm IQA}^{{\rm Inter}({X,\rm N})}], have been computed at the B3LYP, M06-2X and MP2 levels of theory, all of them supported in AIMAll. Note that quantities derived from the second-order density matrix, such as delocalization indices and the two-atomic exchange energies in IQA, are not rigorously defined within DFT, preventing, in principle, their exact evaluation. However, suitable approximations have been implemented in AIMAll allowing, in particular, the total energy of the system to be recovered from the IQA components (Maxwell et al., 2016[Maxwell, P., Pendás, M. & Popelier, P. L. A. (2016). Phys. Chem. Chem. Phys. 18, 20986-21000.]). Concerning the MP2 decomposition, AIMAll implementation of IQA analysis uses natural orbitals of the one-electron density matrix to compute the two-electron density matrix through the Müller approximation (Müller, 1984[Müller, A. M. K. (1984). Phys. Lett. A, 105, 446-452.]). The extent of this approximation has been previously tested (Tognetti et al., 2018[Tognetti, V., Silva, A. F., Vincent, M. A., Joubert, L. & Popelier, P. L. A. (2018). J. Phys. Chem. A, 122, 7748-7756.]) by comparison with that implemented in Morphy, which uses Hartree–Fock orbitals and provides the exact MP2 IQA partitioning (McDonagh et al., 2016[McDonagh, J. L., Vincent, M. A. & Popelier, P. L. A. (2016). Chem. Phys. Lett. 662, 228-234.]; Popelier, 1996[Popelier, P. L. A. (1996). Comput. Phys. Commun. 93, 212-240.]). Evaluation of the three (kinetic, electrostatic and exchange-correlation) physical contributions to atomic energies through the two approaches revealed that, for light-atom (i.e. C, N, O and F) containing molecules, the Müller approximation provides systematically more positive kinetic energies and more negative electrostatic and exchange-correlation energies (Tognetti et al., 2018[Tognetti, V., Silva, A. F., Vincent, M. A., Joubert, L. & Popelier, P. L. A. (2018). J. Phys. Chem. A, 122, 7748-7756.]). It is therefore argued that relating DFT and MP2 properties obtained through different IQA implementations is completely meaningless, while reliable conclusions can be drawn when comparing quantities obtained for a series of compounds within a given approach.

The 2D contour diagrams of the integrand of the SF have been generated with Multiwfn (version 3.8; Lu & Chen, 2012[Lu, T. & Chen, F. (2012). J. Comput. Chem. 33, 580-592.]) using rBCP as the reference point. Finally, molecular representations have been obtained with the Gaussview software (Dennington et al., 2019[Dennington, R., Keith, T. A. & Millam, J. M. (2019). GaussView. Version 6. Gaussian Inc., Wallingford, CT, USA. https://gaussian.com/gaussview6/.]).

4. Results and discussion

Based on the work of Bartashevich et al. (2014[Bartashevich, E. V., Troitskaya, E. A. & Tsirelson, V. G. (2014). Chem. Phys. Lett. 601, 144-148.]), we have considered, as XB acceptors, the same set of substituted pyridines (R-py, i.e. 15 pyridine-based compounds bearing different electron-withdrawing and electron-donor substituents) located in different positions on the ring. The presence of substituents in position 2 allows the evaluation of combined effects arising from electronic factors and steric hindrance with the approaching halogen atom. As XB donors, in addition to I2 (Bartashevich et al., 2014[Bartashevich, E. V., Troitskaya, E. A. & Tsirelson, V. G. (2014). Chem. Phys. Lett. 601, 144-148.]), other halogenated molecules (i.e. Br2, ICN and BrCN) have been considered to evaluate the effects of both the nature of the halogen and the attached substituent on the properties of the σ-hole and the XB features. In line with the investigation of Bartashevich et al. (2014[Bartashevich, E. V., Troitskaya, E. A. & Tsirelson, V. G. (2014). Chem. Phys. Lett. 601, 144-148.]), the B3LYP functional has been adopted throughout. This functional was demonstrated to provide the most accurate IQA energies, using coupled cluster singles and doubles (CCSD) data as a reference, in a previous investigation on intramolecular interactions in glycol conformers (Cukrowski, 2019[Cukrowski, I. (2019). Phys. Chem. Chem. Phys. 21, 10244-10260.]). Selected calculations have also been performed at the MP2 and DFT levels, using four other functionals besides B3LYP for the latter, to highlight the differences between energies, QTAIM properties and IQA results obtained at different levels of theory.

In Tables S1–S6 of the supporting information, the computed interaction energies (EBIND) obtained at the B3LYP, MP2, M06-2X, M11, ωB97X and ωB97XD levels for the four DX⋯(R-)py series of complexes (DX = I2, Br2, ICN and BrCN) are reported. The substituted pyridines are sorted according to the scale of experimental values of pKBI2 (Table S7), where pKBI2 = log10[Kc], Kc being the equilibrium constant for the reaction py + I2 → py⋯I—I in hexane at 298 K (Laurence et al., 2011[Laurence, C., Graton, J., Berthelot, M. & El Ghomari, M. J. (2011). Chem. A Eur. J. 17, 10431-10444.]). Binding energies, approximately increasing in the same order as the pKBI2 values, denote the formation of XBs of medium strength (Cavallo et al., 2016[Cavallo, G., Metrangolo, P., Milani, R., Pilati, T., Priimagi, A., Resnati, G. & Terraneo, G. (2016). Chem. Rev. 116, 2478-2601.]). Interestingly, considering the MP2 reference values, the EBIND of complexes with I2 (−6.3 to −12.1 kcal mol−1) fall into approximately the same range as those of Br2 (−6.0 to −12.3 kcal mol−1), differing from the generally reported increase of XB strength with halogen weight. However, note that most computational studies on XBs involving X2 dihalogens focus on the lighter atoms: F2, Cl2 and Br2 (Alkorta et al., 1998[Alkorta, I., Rozas, I. & Elguero, J. (1998). J. Phys. Chem. A, 102, 9278-9285.]; Wang et al., 2014[Wang, C., Danovich, D., Mo, Y. & Shaik, S. (2014). J. Chem. Theory Comput. 10, 3726-3737.]; Karpfen, 2003[Karpfen, A. (2003). Theoretical Chemistry Accounts: Theory, Computation, and Modeling (Theoretica Chimica Acta), 110, 1-9.]) with only few exceptions including I2 (Bartashevich et al., 2015[Bartashevich, E., Troitskaya, E., Pendás, M. & Tsirelson, V. (2015). Comput. Theor. Chem. 1053, 229-237.]; McAllister et al., 2023[McAllister, L. J., Taylor, J., Pridmore, N. E., McEllin, A. J., Whitwood, A. C., Karadakov, P. B. & Bruce, D. W. (2023). CrystEngComm, 25, 1683-1692.]). These results suggest that further assessment of the XB strength involving the full series of dihalogens using, in particular, well calibrated basis sets, is highly desirable. Replacement of the non-interacting halogen with the electron-withdrawing CN group implies a decrease of EBIND in almost all cases, which is much more pronounced for bromine complexes, restoring the expected I > Br energy trend. MP2 results therefore show that EBIND decreases in the order I2 ≃ Br2 > ICN > BrCN. Note, however, that XCN molecules are characterized by larger and higher σ-holes than those of the corresponding X2 dihalogens (Fig. 1[link]), as expected based on the different electronegativity of the involved species. More precisely, the maximum ESP values at the σ-holes are +0.031, +0.034, +0.056 and +0.055 a.u. for I2, Br2, ICN and BrCN, respectively.

[Figure 1]
Figure 1
ESPs on the isosurface of electron density (0.001 electrons bohr−3) calculated at the B3LYP/6-311++G(d,p) level for (a) I2, (b) ICN, (c) Br2 and (d) BrCN.

These results provide evidence that halogen bonding is not (always) dominated by electrostatics, because the concept of the σ-hole cannot be considered as the single parameter to be associated with the strength of the interaction (Tognetti & Joubert, 2016[Tognetti, V. & Joubert, L. (2016). Applications of Topological Methods in Molecular Chemistry, edited by R. Chauvin, C. Lepetit, B. Silvi & E. Alikhani, pp. 435-459. Cham: Springer International Publishing.]). As recognized by Tognetti and coworkers (Syzgantseva et al., 2013[Syzgantseva, O. A., Tognetti, V. & Joubert, L. (2013). J. Phys. Chem. A, 117, 8969-8980.]): `at long-range, electrostatics clearly dominate (as expected from the σ-hole model) and are responsible for the initiation of the bond formation process'. At short range, covalence contributions can become important, as demonstrated by the VB/MO (Wang et al., 2014[Wang, C., Danovich, D., Mo, Y. & Shaik, S. (2014). J. Chem. Theory Comput. 10, 3726-3737.]), SCVB (Franchini et al., 2019[Franchini, D., Genoni, A., Dapiaggi, F., Pieraccini, S. & Sironi, M. (2019). Int. J. Quantum Chem. 119, e25946.]; Forni et al., 2024[Forni, A., Genoni, A., Pieraccini, S. & Sironi, M. (2024). Comprehensive Computational Chemistry, 1st ed., edited by M. Yáñez & R. J. Boyd, pp. 533-551. Oxford: Elsevier.]) and MP4/IQA calculations (Alkorta et al., 2020[Alkorta, I., Silva, A. F. & Popelier, P. L. A. (2020). Molecules, 25, 2674.]).

Looking at the DFT results obtained with the different examined functionals, the better agreement with MP2 is provided by M06-2X calculations, confirming this functional as one of the best for modelling XB complexes at the DFT level (Kozuch & Martin, 2013[Kozuch, S. & Martin, J. M. L. (2013). J. Chem. Theory Comput. 9, 1918-1931.]; Forni et al., 2016[Forni, A., Pieraccini, S., Franchini, D. & Sironi, M. (2016). J. Phys. Chem. A, 120, 9071-9080.], 2014[Forni, A., Pieraccini, S., Rendine, S. & Sironi, M. (2014). J. Comput. Chem. 35, 386-394.], 2012[Forni, A., Pieraccini, S., Rendine, S., Gabas, F. & Sironi, M. (2012). ChemPhysChem, 13, 4224-4234.]), though all the other functionals display an acceptable performance.

IQA studies have been performed on MP2 and DFT wavefunctions (within the approximations adopted in AIMAll, see Computational details[link]) using, for the latter, the B3LYP and M06-2X functionals. In particular, owing to its relatively affordable computational costs, B3LYP has been used to determine the total binding energy ([E_{\rm IQA}^{\rm BIND}]) according to equation (8)[link], and ascertain the error, due to the adopted approximations, in recovering the exact B3LYP [E^{\rm BIND}] value (see Computational details[link]). The relative errors, reported in Table S9, are always below 10% for the complexes with bromine and below 15% for those with iodine, though no systematic trends can be individuated. Note that the rather good agreement between the two energy values (see Fig. S1 of the supporting information) is due to the relatively high interaction energies of the investigated systems, where dispersion contributions are low (Syzgantseva et al., 2013[Syzgantseva, O. A., Tognetti, V. & Joubert, L. (2013). J. Phys. Chem. A, 117, 8969-8980.]).

The two-centre XB energies [E_{\rm IQA}^{{\rm Inter}({X,{\rm N}})}] [see equation (7[link])] obtained at the B3LYP, MP2 and M06-2X levels are reported in Table S10. Analysis of a possible correlation between [E_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] and EBIND for each DX⋯(R-)py set indicates that, in general, these descriptors are not correlated for the present systems, except for only a few fortuitous cases (Fig. S2). This result indicates that the formation of medium-strength XBs implies a strong rearrangement within the two interacting fragments, with associated variations of atomic energies which depend on both the DX and the R-py species (Tognetti & Joubert, 2016[Tognetti, V. & Joubert, L. (2016). Applications of Topological Methods in Molecular Chemistry, edited by R. Chauvin, C. Lepetit, B. Silvi & E. Alikhani, pp. 435-459. Cham: Springer International Publishing.]). Note that, in the case of strong XB interactions (e.g. involving anions), the intra-fragment energy variations become less relevant with respect to the interaction energy, resulting in good agreement between EBIND and [E_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] (Syzgantseva et al., 2013[Syzgantseva, O. A., Tognetti, V. & Joubert, L. (2013). J. Phys. Chem. A, 117, 8969-8980.]).

Inspection of the percentage contribution of the exchange-correlation energy, [{\rm VX}_{\rm IQA}^{{\rm Inter}(X,{\rm N})}], to [E_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] (Table S10) indicates that this term always has a remarkable weight, in some cases comparable to or even greater than the electrostatic [{\rm VC}_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] one. We note that:

(i) For each DX donor, such contribution shows only a slight increase with the strength of the interaction, from about 3 to 6% according to the method and the DX donor, in spite of a much greater increase of [E_{\rm IQA}^{{\rm Inter}(X, {\rm N})}], allowing us to consider an average contribution for each set of complexes, as reported in Table 1[link]. Here, it can be observed that the B3LYP functional systematically provides the greater exchange-correlation contributions, followed by MP2 and then M06-2X.

Table 1
Average percentage contribution of the exchange-correlation term, [{\rm VX}_{\rm IQA}^{{\rm Inter}({X,{\rm N}})}], to the two-centre XB energy, [E_{\rm IQA}^{{\rm Inter}({X,{\rm N}})}], computed at the B3LYP, MP2 and M06-2X levels of theory

  B3LYP MP2 M06-2X
I2 51.60 49.01 44.09
Br2 28.37 27.59 25.13
ICN 62.86 57.98 55.97
BrCN 32.14 30.05 29.05

(ii) Complexes with iodo-derivatives have much greater [{\rm VX}_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] contributions than the corresponding bromo-derivative complexes.

(iii) For systems containing XCN, [{\rm VX}_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] is significantly lower with respect to systems containing X2, in agreement with the electron-withdrawing power of the CN group, which imparts greater electrostatic character in the interaction. Accordingly, the average [{\rm VX}_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] values are found to increase in the order Br2 < BrCN < I2 < ICN or, equivalently, the electrostatic contribution, [{\rm VC}_{\rm IQA}^{{\rm Inter}(X,{\rm N})}], to the [E_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] interaction energy decreases in the same order from ICN to Br2, as qualitatively inferred from the electronegativity of the atomic species involved.

The existence of possible relationships between the two-atomic energy contributions and local properties of the electron density has been then tested. The exchange-correlation contribution was found to correlate very well, at each level of theory, with both the electron density at the BCP of the X⋯N bond, ρX⋯N(rBCP), and the delocalization index δ(ΩX, ΩN). The relationships are found to depend only on the nature of the halogen atom. The correlations of ρX⋯N(rBCP) with [{\rm VX}_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] obtained at the B3LYP level are reported in Fig. 2[link] [see Figs. S3 and S4 for all results obtained at the B3LYP, MP2 and M06-2X levels for the correlations of ρX⋯N(rBCP) and δ(ΩX, ΩN), respectively, with [{\rm VX}_{\rm IQA}^{{\rm Inter}(X,{\rm N})}]].

[Figure 2]
Figure 2
Relationships between electron density at the X⋯N BCP, ρ(rBCP) and exchange-correlation contribution to the XB energy obtained at the B3LYP level of theory for complexes formed with (a) iodine and (b) bromine derivatives.

Moreover, a good correlation of [E_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] with both ρX⋯N(rBCP) and δ(ΩX, ΩN) was also found at each level of theory but, in this case, four different relationships have been identified, one for each set of dimers. In Fig. 3[link] the correlations of [E_{\rm IQA}^{{\rm Inter}(X,{\rm N})}] with δ(ΩX, ΩN) obtained at the B3LYP level are reported (see Figs. S5 and S6 for the whole set of results). Note that, for both I⋯N and Br⋯N interactions, complexes formed with dihalogens display better correlation than those with XCN.

[Figure 3]
Figure 3
Relationships between the delocalization index and the IQA XB energy obtained at the B3LYP level.

On the other hand, no correlation has been found between [E^{\rm BIND}] and δ(ΩX, ΩN) (Fig. S7). More precisely, for each dataset, only 9 out of 15 complexes show a linear trend. The dimers deviating from this trend correspond to the ortho-substituted pyridines, whose deviations increase with the size of the substituent. Moreover, it is greater for complexes with XCN, suggesting the presence of interactions between the substituent on the pyridine ring and both the halogen and its attached CN group. This result underlines the importance of secondary interactions on the stability of the complexes, which cannot be predicted by simply analysing the properties exclusively connected to the interacting X and N atoms (Syzgantseva et al., 2013[Syzgantseva, O. A., Tognetti, V. & Joubert, L. (2013). J. Phys. Chem. A, 117, 8969-8980.]).

Finally, halogen-dependent relationships have also been found between ρX⋯N(rBCP) and δ(ΩX, ΩN). Fig. 4[link] illustrates how the delocalization index between the two basins is proportional to the electron density at the X⋯N BCP. All results are collected in Fig. S8.

[Figure 4]
Figure 4
Relationships between the δ(ΩX, ΩN) delocalization index and the electron density at the X⋯N BCP obtained at the B3LYP level for complexes formed with moieties containing (a) iodine and (b) bromine atoms.

Further information on the nature of the X⋯N XB is provided by the SF in the corresponding rBCP. Looking at Tables S1–S6, reporting the absolute (SF) and percentage (SF%) values of the SF for halogen and nitro­gen atoms, the main contribution to ρX⋯N(rBCP) as a source of electron density (i.e. with positive SF) comes from the halogen, with SF% values up to 50%, and slightly increasing with the strength of the interaction and decreasing in the order I2 ≃ Br2 > ICN > BrCN. The nitro­gen atom displays a greater variability in its SF% contribution, acting as either a sink (in XCN complexes) or a source (in most X2 ones) to ρX⋯N(rBCP). Again, the stronger the interaction, the greater the nitro­gen SF% value, the maximum value being 16% (B3LYP and MP2 results). Remarkably, the sum of SF% on the halogen and nitro­gen atoms accounts for far less than 100% of the electron density at the X⋯N BCP, their cumulative contributions amounting at most to 65% (in Br2 complexes, B3LYP and MP2 results). The remaining density mainly comes from the atoms directly bonded to the interacting X and N atoms and gradually decreases moving away from them, therefore evidencing how the X⋯N BCP topological properties are strongly influenced by the distal atoms.

The different atomic contributions to electron density at the ρX⋯N(rBCP) can of course be visually inspected by looking at the integrand of the SF, [L({\bf r'}) = {{\nabla^2\rho({\bf r'})} / {{\bf r} - {\bf r'}}}]. The 2D contour diagrams of L(r) for the complexes of unsubstituted pyridine with I2, ICN, Br2 and BrCN using the B3LYP wavefunctions are reported in Fig. 5[link], focusing on the XB region. By comparing the L(r) contour diagrams of the complexes with I2 and Br2 [Figs. 5[link](a) and 5[link](c)], it is clear that the bromine atom contributes more significantly than the iodine atom to ρX⋯N(rBCP), explaining the higher interaction energy associated with the (Br-)Br⋯N bond in this complex. When the halogen bonded to bromine is substituted with a CN group, the bromine contribution at the BCP decreases [Fig. 5[link](d)]. A comparable effect is observed for its iodo-derivative analogue, though the extent of this decrease is less pronounced [Fig. 5[link](b)], reflecting in the higher interaction energy for the complex with ICN with respect to that with BrCN.

[Figure 5]
Figure 5
2D contour diagrams of L(r) for the complexes of pyridine with (a) I2, (b) ICN, (c) Br2 and (d) BrCN using the B3LYP wavefunctions. Solid-red and dashed-blue lines correspond to positive and negative contours, respectively.

When pyridine bears substituents in position 2, a drop in the SF% contribution for both the halogen and the nitro­gen atoms is observed. This result can be explained by looking at the corresponding L(r) contour diagrams. Comparing the plot for the complex 2-chloro­pyridine·I2 (Fig. 6[link]) with that of pyridine·I2 [Fig. 5[link](a)], we can see that the presence of the chlorine atom, besides determining a deviation of the C– I⋯N bond angle from linearity, influences the shape and the extension of both the nitro­gen and iodine basins, reducing their contribution to L(r). Concomitantly, a high SF% value (8%) is obtained for the chlorine atom, much greater than that of the hydrogen atom (4%) placed in the same position in pyridine·I2. Analogous results have been obtained replacing chlorine with different groups such as -F, -CH3, -CH2CH3 and -CH(CH3)2.

[Figure 6]
Figure 6
2D contour diagrams of L(r) for the complex 2-chloro­pyridine·I2 using the B3LYP wavefunction. Solid-red and dashed-blue lines correspond to positive and negative contours, respectively.

The lower extension of nitro­gen and halogen basins in the 2-substituted complexes is associated with a lower delocalization index between them. It was previously demonstrated that, for the (I-)I⋯N XB (Bartashevich et al., 2014[Bartashevich, E. V., Troitskaya, E. A. & Tsirelson, V. G. (2014). Chem. Phys. Lett. 601, 144-148.]), the corresponding delocalization index is quantitatively correlated to the sum of the atomic contributions of the halogen and nitro­gen atoms to ρX⋯N(rBCP). We can generalize this finding by expressing [\delta(\Omega_{X},\Omega_{\rm N})] as

[\delta\left(\Omega_{X},\Omega_{\rm N} \right) = \alpha {\rm SF}\left(\Omega_{X} \right) + \beta {\rm SF}\left(\Omega_{\rm N} \right) + \gamma, \eqno (9)]

with different α, β and γ values according to both the different set of complexes and the level of theory, as summarized in Table S11. A very good agreement is observed between the δ(ΩX, ΩN) values (Tables S1–S6) and those calculated with equation (9)[link], reported in Table S12.

It is also interesting to observe how substituents in positions 3 and 4 influence the electron density at the BCP. Halogen and nitro­gen contributions to the SF are greater, compared with those obtained with non-substituted pyridine, for systems containing electron-donor groups, while a drop is observed for those containing electron-withdrawing groups. Moreover, the electron-donation effect increases with the size of the substituent, but it appears that increasing the number of substituents is preferable to increasing their size. All these variations in the atomic contributions to ρX⋯N(rBCP) clearly reflect in different interaction energies, allowing us to quantitatively relate the strength of the interaction to the specific SF% values of both the interacting atomic pair and the distal atoms.

5. Conclusions

The X⋯N (X = I, Br) halogen-bonding interactions in complexes formed by substituted pyridines and XCN/X2 molecules have been investigated in the QTAIM/IQA framework. IQA analysis underlines the importance of the exchange-correlation contribution to the total interaction energy, in some cases dominating the electrostatic term. Quantitative relationships between IQA energies and QTAIM descriptors, such as the electron density at the X⋯N BCP and delocalization index, have been found. Such topological properties generally show a good correlation with the two-centre IQA interaction energy between the two main atoms involved in the interaction. The obtained relationships only depend on the nature of the halogen atom if the sole exchange-correlation contribution is considered. On the other hand, no correlations with local topological properties at the X⋯N BCP are in general found if the total binding energies are considered, underlying the important role of the local environment on the stability of the complexes. This point is quantified by analysis of the atomic contributions to the electron density at the X⋯N BCP, as provided by the SF. This analysis allows us to show the essentially non-local character of electron density and quantitatively relates the strength of the interaction with specific atomic contributions coming from, not only the directly interacting pair, but also the distal atoms including the different substituents on the pyridine ring.

Supporting information


Conflict of interest

The authors declare that there are no conflicts of interest.

Data availability

The data supporting the results reported in this article can be accessed within the article and as published supporting material.

References

First citationAlkorta, I., Rozas, I. & Elguero, J. (1998). J. Phys. Chem. A, 102, 9278–9285.  Web of Science CrossRef CAS Google Scholar
First citationAlkorta, I., Silva, A. F. & Popelier, P. L. A. (2020). Molecules, 25, 2674.  CrossRef PubMed Google Scholar
First citationArunan, E., Desiraju, G. R., Klein, R. A., Sadlej, J., Scheiner, S., Alkorta, I., Clary, D. C., Crabtree, R. H., Dannenberg, J. J., Hobza, P., Kjaergaard, H. G., Legon, A. C., Mennucci, B. & Nesbitt, D. J. (2011). Pure Appl. Chem. 83, 1637–1641.   CrossRef CAS Google Scholar
First citationAuffinger, P., Hays, F. A., Westhof, E. & Ho, P. S. (2004). Proc. Natl Acad. Sci. USA, 101, 16789–16794.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBader, R. F. W. (1990). Atoms in Molecules: a Quantum Theory. International Series of Monographs on Chemistry. Clarendon Press.  Google Scholar
First citationBader, R. F. W. (1991). Chem. Rev. 91, 893–928.  CrossRef CAS Web of Science Google Scholar
First citationBader, R. F. W. & Essén, H. (1984). J. Chem. Phys. 80, 1943–1960.  CrossRef CAS Web of Science Google Scholar
First citationBader, R. F. W. & Gatti, C. (1998). Chem. Phys. Lett. 287, 233–238.  Web of Science CrossRef CAS Google Scholar
First citationBader, R. F. W. & Stephens, M. E. (1975). J. Am. Chem. Soc. 97, 7391–7399.  CrossRef CAS Web of Science Google Scholar
First citationBader, R. F. W., Streitwieser, A., Neuhaus, A., Laidig, K. E. & Speers, P. (1996). J. Am. Chem. Soc. 118, 4959–4965.  CrossRef CAS Web of Science Google Scholar
First citationBartashevich, E., Troitskaya, E., Pendás, M. & Tsirelson, V. (2015). Comput. Theor. Chem. 1053, 229–237.  CrossRef CAS Google Scholar
First citationBartashevich, E. V., Troitskaya, E. A. & Tsirelson, V. G. (2014). Chem. Phys. Lett. 601, 144–148.  CrossRef CAS Google Scholar
First citationBecke, A. D. (1993). J. Chem. Phys. 98, 5648–5652.  CrossRef CAS Web of Science Google Scholar
First citationBissantz, C., Kuhn, B. & Stahl, M. (2010). J. Med. Chem. 53, 5061–5084.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBlanco, M. A., Martín Pendás, A. & Francisco, E. (2005). J. Chem. Theory Comput. 1, 1096–1109.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBolton, O., Lee, K., Kim, H.-J., Lin, K. Y. & Kim, J. (2011). Nat. Chem. 3, 415.  CrossRef Google Scholar
First citationBorozan, S. Z. & Stojanović, S. Đ. (2013). Comput. Biol. Chem. 47, 231–239.  CrossRef CAS PubMed Google Scholar
First citationBrammer, L., Peuronen, A. & Roseveare, T. M. (2023). Acta Cryst. C79, 204–216.  Web of Science CrossRef IUCr Journals Google Scholar
First citationBulfield, D. & Huber, S. M. (2016). Chem. A Eur. J. 22, 14434–14450.  CrossRef CAS Google Scholar
First citationCavallo, G., Metrangolo, P., Milani, R., Pilati, T., Priimagi, A., Resnati, G. & Terraneo, G. (2016). Chem. Rev. 116, 2478–2601.  Web of Science CrossRef CAS PubMed Google Scholar
First citationChai, J.-D. & Head-Gordon, M. (2008a). Phys. Chem. Chem. Phys. 10, 6615–6620.  Web of Science CrossRef PubMed CAS Google Scholar
First citationChai, J.-D. & Head-Gordon, M. (2008b). J. Chem. Phys. 128, 084106.  Web of Science CrossRef PubMed Google Scholar
First citationClark, T. (2013). WIREs Comput. Mol. Sci. 3, 13–20.  CrossRef CAS Google Scholar
First citationClark, T., Hennemann, M., Murray, J. S. & Politzer, P. (2007). J. Mol. Model. 13, 291–296.  Web of Science CrossRef PubMed CAS Google Scholar
First citationCukrowski, I. (2019). Phys. Chem. Chem. Phys. 21, 10244–10260.  CrossRef CAS PubMed Google Scholar
First citationDecato, D. A., Riel, A. M. S., May, J. H., Bryantsev, V. S. & Berryman, O. B. (2021). Angew. Chem. Int. Ed. 60, 3685–3692.  Web of Science CSD CrossRef CAS Google Scholar
First citationDennington, R., Keith, T. A. & Millam, J. M. (2019). GaussView. Version 6. Gaussian Inc., Wallingford, CT, USA. https://gaussian.com/gaussview6/Google Scholar
First citationDesiraju, G. R., Ho, P. S., Kloo, L., Legon, A. C., Marquardt, R., Metrangolo, P., Politzer, P., Resnati, G. & Rissanen, K. (2013). Pure Appl. Chem. 85, 1711–1713.  Web of Science CrossRef CAS Google Scholar
First citationEccles, K., Morrison, R., Sinha, A., Maguire, A. & Lawrence, S. (2014). Acta Cryst. A70, C651.  CrossRef IUCr Journals Google Scholar
First citationEraković, M., Cinčić, D., Molčanov, K. & Stilinović, V. (2019). Angew. Chem. Int. Ed. 58, 15702–15706.  Google Scholar
First citationFeller, D. (1996). J. Comput. Chem. 17, 1571–1586.  CrossRef CAS Google Scholar
First citationFord, M. C., Saxton, M. & Ho, P. S. (2017). J. Phys. Chem. Lett. 8, 4246–4252.  CrossRef CAS PubMed Google Scholar
First citationForni, A., Franchini, D., Dapiaggi, F., Pieraccini, S., Sironi, M., Scilabra, P., Pilati, T., Petko, K. I., Resnati, G. & Yagupolkii, Y. L. (2019). Cryst. Growth Des. 19, 1621–1631.  Web of Science CSD CrossRef CAS Google Scholar
First citationForni, A., Genoni, A., Pieraccini, S. & Sironi, M. (2024). Comprehensive Computational Chemistry, 1st ed., edited by M. Yáñez & R. J. Boyd, pp. 533–551. Oxford: Elsevier.  Google Scholar
First citationForni, A., Pieraccini, S., Franchini, D. & Sironi, M. (2016). J. Phys. Chem. A, 120, 9071–9080.  CrossRef CAS PubMed Google Scholar
First citationForni, A., Pieraccini, S., Rendine, S., Gabas, F. & Sironi, M. (2012). ChemPhysChem, 13, 4224–4234.  Web of Science CrossRef CAS PubMed Google Scholar
First citationForni, A., Pieraccini, S., Rendine, S. & Sironi, M. (2014). J. Comput. Chem. 35, 386–394.  CrossRef CAS PubMed Google Scholar
First citationFotović, L., Bedeković, N. & Stilinović, V. (2023). Cryst. Growth Des. 23, 3384–3392.  PubMed Google Scholar
First citationFourmigué, M. (2009). Curr. Opin. Solid State Mater. Sci. 13, 36–45.  Google Scholar
First citationFradera, X., Austen, M. A. & Bader, R. F. W. (1999). J. Phys. Chem. A, 103, 304–314.  Web of Science CrossRef CAS Google Scholar
First citationFranchini, D., Forni, A., Genoni, A., Pieraccini, S., Gandini, E. & Sironi, M. (2020). ChemistryOpen, 9, 445–450.  CrossRef CAS PubMed Google Scholar
First citationFranchini, D., Genoni, A., Dapiaggi, F., Pieraccini, S. & Sironi, M. (2019). Int. J. Quantum Chem. 119, e25946.  CrossRef Google Scholar
First citationFrancisco, E., Martín Pendás, A. & Blanco, M. A. (2006). J. Chem. Theory Comput. 2, 90–102.  Web of Science CrossRef CAS PubMed Google Scholar
First citationFrisch, M. J., Head-Gordon, M. & Pople, J. A. (1990a). Chem. Phys. Lett. 166, 275–280.  CrossRef CAS Google Scholar
First citationFrisch, M. J., Head-Gordon, M. & Pople, J. A. (1990b). Chem. Phys. Lett. 166, 281–289.  CrossRef CAS Google Scholar
First citationFrisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Petersson, G. A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A. V., Bloino, J., Janesko, B. G., Gomperts, R., Mennucci, B., Hratchian, H. P., Ortiz, J. V., Izmaylov, A. F., Sonnenberg, J. L., Williams-Young, D., Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone, A., Henderson, T., Ranasinghe, D., Zakrzewski, V. G., Gao, J., Rega, N., Zheng, G., Liang, W., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M. J., Heyd, J. J., Brothers, E. N., Kudin, K. N., Staroverov, V. N., Keith, T. A., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A. P., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Millam, J. M., Klene, M., Adamo, C., Cammi, R., Ochterski, J. W., Martin, R. L., Morokuma, K., Farkas, O., Foresman, J. B. & Fox, D. J. (2016). Gaussian 16. Revision C.01. Gaussian Inc., Wallingford, Connecticut, USA.  Google Scholar
First citationFrontera, A. & Bauzá, A. (2021). Crystals (Basel), 11, 1205.  CrossRef Google Scholar
First citationGatti, C., Saleh, G. & Lo Presti, L. (2016). Acta Cryst. B72, 180–193.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGlukhovtsev, M. N., Pross, A., McGrath, M. P. & Radom, L. (1995). J. Chem. Phys. 103, 1878–1885.   CrossRef CAS Google Scholar
First citationGrabowski, S. J. (2021). Molecules, 26, 4939.  Web of Science CrossRef PubMed Google Scholar
First citationGrabowski, S. J. (2024). Chem. Commun. 60, 6239–6255.  CrossRef CAS Google Scholar
First citationGuerra, C., Burgos, J., Ayarde-Henríquez, L. & Chamorro, E. (2024). J. Phys. Chem. A, 128, 6158–6166.  CrossRef CAS PubMed Google Scholar
First citationGuevara-Vela, J. M., Francisco, E., Rocha-Rinza, T. & Martín Pendás, Á. (2020). Molecules, 25, 4028.  PubMed Google Scholar
First citationHajji, M., Haouas, A., Abad, N. & Guerfel, T. (2023). ChemistrySelect, 8, e202302624.  CSD CrossRef Google Scholar
First citationHakkert, S. B. & Erdélyi, M. (2015). J. Phys. Org. Chem. 28, 226–233.  Web of Science CrossRef CAS Google Scholar
First citationHaque, A., Alenezi, K. M., Khan, M. S., Wong, W.-Y. & Raithby, P. R. (2023). Chem. Soc. Rev. 52, 454–472.  CrossRef CAS PubMed Google Scholar
First citationHauchecorne, D., Moiana, A., van der Veken, B. J. & Herrebout, W. A. (2011). Phys. Chem. Chem. Phys. 13, 10204–10213.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHead-Gordon, M. & Head-Gordon, T. (1994). Chem. Phys. Lett. 220, 122–128.  CAS Google Scholar
First citationHead-Gordon, M., Pople, J. A. & Frisch, M. J. (1988). Chem. Phys. Lett. 153, 503–506.  CAS Google Scholar
First citationJena, S., Dutta, J., Tulsiyan, K. D., Sahu, A. K., Choudhury, S. S. & Biswal, H. S. (2022). Chem. Soc. Rev. 51, 4261–4286.  Web of Science CrossRef CAS PubMed Google Scholar
First citationJiang, Y., Alcaraz, A. A., Chen, J.-M., Kobayashi, H., Lu, Y. J. & Snyder, J. P. (2006). J. Med. Chem. 49, 1891–1899.  CSD CrossRef PubMed CAS Google Scholar
First citationJiménez-Grávalos, F., Gallegos, M., Martín Pendás, Á. & Novikov, A. S. (2021). J. Comput. Chem. 42, 676–687.  Web of Science PubMed Google Scholar
First citationKarpfen, A. (2003). Theoretical Chemistry Accounts: Theory, Computation, and Modeling (Theoretica Chimica Acta), 110, 1–9.  Google Scholar
First citationKato, R., Imakubo, T., Yamamoto, H., Maeda, R., Fujiwara, M., Yamaura, J.-I. & Sawa, H. (2002). Mol. Cryst. Liq. Cryst. 380, 61–68.  CrossRef CAS Google Scholar
First citationKojasoy, V. & Tantillo, D. J. (2022). Org. Biomol. Chem. 21, 11–23.  CrossRef PubMed Google Scholar
First citationKolář, M., Hostaš, J. & Hobza, P. (2014). Phys. Chem. Chem. Phys. 16, 9987–9996.  PubMed Google Scholar
First citationKolář, M. H. & Hobza, P. (2016). Chem. Rev. 116, 5155–5187.  Web of Science PubMed Google Scholar
First citationKosmas, A. M. (2007). Bioinorg. Chem. Appl. 2007, 046393.  CrossRef Google Scholar
First citationKozuch, S. & Martin, J. M. L. (2013). J. Chem. Theory Comput. 9, 1918–1931.  Web of Science CrossRef CAS PubMed Google Scholar
First citationKratzer, P., Ramming, B., Römisch, S. & Maas, G. (2015). Cryst­EngComm, 17, 4486–4494.  Web of Science CSD CrossRef CAS Google Scholar
First citationLauher, J. W., Fowler, F. W. & Goroff, N. S. (2008). Acc. Chem. Res. 41, 1215–1229.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLaurence, C., Graton, J., Berthelot, M. & El Ghomari, M. J. (2011). Chem. A Eur. J. 17, 10431–10444.  CrossRef CAS Google Scholar
First citationLee, C., Yang, W. & Parr, R. G. (1988). Phys. Rev. B, 37, 785–789.  CrossRef CAS Web of Science Google Scholar
First citationLo, R., Ballabh, A., Singh, A., Dastidar, P. & Ganguly, B. (2012). CrystEngComm, 14, 1833–1841.  CrossRef CAS Google Scholar
First citationLu, T. & Chen, F. (2012). J. Comput. Chem. 33, 580–592.  Web of Science CrossRef PubMed Google Scholar
First citationLu, Y., Shi, T., Wang, Y., Yang, H., Yan, X., Luo, X., Jiang, H. & Zhu, W. (2009). J. Med. Chem. 52, 2854–2862.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLu, Y., Wang, Y. & Zhu, W. (2010). Phys. Chem. Chem. Phys. 12, 4543–4551.  Web of Science CrossRef CAS PubMed Google Scholar
First citationLuo, J., Dai, H., Zeng, C., Wu, D. & Cao, M. (2022). Molecules, 27. https://doi.org/10.3390/molecules27228078Google Scholar
First citationMaxwell, P., Pendás, M. & Popelier, P. L. A. (2016). Phys. Chem. Chem. Phys. 18, 20986–21000.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMcAllister, L. J., Taylor, J., Pridmore, N. E., McEllin, A. J., Whitwood, A. C., Karadakov, P. B. & Bruce, D. W. (2023). CrystEngComm, 25, 1683–1692.  CSD CrossRef CAS Google Scholar
First citationMcDonagh, J. L., Vincent, M. A. & Popelier, P. L. A. (2016). Chem. Phys. Lett. 662, 228–234.  CrossRef CAS Google Scholar
First citationMessina, M. T., Metrangolo, P., Panzeri, W., Ragg, E. & Resnati, G. (1998). Tetrahedron Lett. 39, 9069–9072.  CrossRef CAS Google Scholar
First citationMetrangolo, P., Carcenac, Y., Lahtinen, M., Pilati, T., Rissanen, K., Vij, A. & Resnati, G. (2009). Science, 323, 1461–1464.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationMetrangolo, P., Panzeri, W., Recupero, F. & Resnati, G. (2002). J. Fluor. Chem. 114, 27–33.  CrossRef CAS Google Scholar
First citationMetrangolo, P. & Resnati, G. (2008). Science, 321, 918–919.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMetrangolo, P. & Resnati, G. (2012). Cryst. Growth Des. 12, 5835–5838.  Web of Science CrossRef CAS Google Scholar
First citationMínguez Espallargas, G., Brammer, L. & Sherwood, P. (2006). Angew. Chem. Int. Ed. 45, 435–440.  Google Scholar
First citationMínguez Espallargas, G., Zordan, F., Arroyo Marín, L., Adams, H., Shankland, K., van de Streek, J. & Brammer, L. (2009). Chem. A Eur. J. 15, 7554–7568.  Google Scholar
First citationMo, Y., Bao, P. & Gao, J. (2011). Phys. Chem. Chem. Phys. 13, 6760–6775.  CrossRef CAS PubMed Google Scholar
First citationMo, Y., Song, L. & Lin, Y. (2007). J. Phys. Chem. A, 111, 8291–8301.  CrossRef PubMed CAS Google Scholar
First citationMoghadasnia, M. P., Eckstein, B. J., Martin, H. R., Paredes, J. U. & McGuirk, C. M. (2024). Cryst. Growth Des. 24, 2304–2321.  CrossRef CAS Google Scholar
First citationMolina, P., Zapata, F. & Caballero, A. (2017). Chem. Rev. 117, 9907–9972.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMüller, A. M. K. (1984). Phys. Lett. A, 105, 446–452.  Google Scholar
First citationMulliken, R. S. (1950). J. Am. Chem. Soc. 72, 600–608.  CrossRef CAS Web of Science Google Scholar
First citationMurray, J. S., Lane, P. & Politzer, P. (2009). J. Mol. Model. 15, 723–729.  Web of Science CrossRef PubMed CAS Google Scholar
First citationNelyubina, Y., Antipin, M. & Lyssenko, K. (2011). Mendeleev Commun. 21, 250–252.  CrossRef CAS Google Scholar
First citationNemec, V., Lisac, K., Bedeković, N., Fotović, L., Stilinović, V. & Cinčić, D. (2021). CrystEngComm, 23, 3063–3083.  Web of Science CrossRef CAS Google Scholar
First citationNguyen, H. L., Horton, P. N., Hursthouse, M. B., Legon, A. C. & Bruce, D. W. (2004). J. Am. Chem. Soc. 126, 16–17.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationOtte, F., Kleinheider, J., Hiller, W., Wang, R., Englert, U. & Strohmann, C. (2021). J. Am. Chem. Soc. 143, 4133–4137.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationOuteiral, C., Vincent, M. A., Martín Pendás, Á. & Popelier, P. L. A. (2018). Chem. Sci. 9, 5517–5529.  CrossRef CAS PubMed Google Scholar
First citationParisini, E., Metrangolo, P., Pilati, T., Resnati, G. & Terraneo, G. (2011). Chem. Soc. Rev. 40, 2267–2278.  Web of Science CrossRef CAS PubMed Google Scholar
First citationPeverati, R. & Truhlar, D. G. (2011). J. Phys. Chem. Lett. 2, 2810–2817.  Web of Science CrossRef CAS Google Scholar
First citationPhan Dang, C.-T., Tam, N. M., Huynh, T.-N. & Trung, N. T. (2023). RSC Adv. 13, 31507–31517.  CrossRef CAS PubMed Google Scholar
First citationPolitzer, P., Lane, P., Concha, M. C., Ma, Y. & Murray, J. S. (2007). J. Mol. Model. 13, 305–311.  Web of Science CrossRef PubMed CAS Google Scholar
First citationPolitzer, P. & Murray, J. S. (2013). ChemPhysChem, 14, 278–294.  Web of Science CrossRef CAS PubMed Google Scholar
First citationPolitzer, P., Murray, J. S. & Clark, T. (2010). Phys. Chem. Chem. Phys. 12, 7748–7757.  Web of Science CrossRef CAS PubMed Google Scholar
First citationPolitzer, P., Murray, J. S. & Clark, T. (2013). Phys. Chem. Chem. Phys. 15, 11178–11189.  Web of Science CrossRef CAS PubMed Google Scholar
First citationPopelier, P., Aicken, F. & O'Brien, S. (2000). Chemical Modelling: Applications and Theory, Vol. 1, pp. 143–198.  Google Scholar
First citationPopelier, P. L. A. (1996). Comput. Phys. Commun. 93, 212–240.  CrossRef CAS Web of Science Google Scholar
First citationPopelier, P. L. A. (2000). Atoms in Molecules: an Introduction. London: Prentice Hall.  Google Scholar
First citationRiley, K. E., Murray, J. S., Fanfrlík, J., Řezáč, J., Solá, R. J., Concha, M. C., Ramos, F. M. & Politzer, P. (2011). J. Mol. Model. 17, 3309–3318.  Web of Science CrossRef CAS PubMed Google Scholar
First citationRiley, K. E., Murray, J. S., Politzer, P., Concha, M. C. & Hobza, P. (2009). J. Chem. Theory Comput. 5, 155–163.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSaebø, S. & Almlöf, J. (1989). Chem. Phys. Lett. 154, 83–89.  Google Scholar
First citationSaha, S., Mishra, M. K., Reddy, C. M. & Desiraju, G. R. (2018). Acc. Chem. Res. 51, 2957–2967.  Web of Science CrossRef CAS PubMed Google Scholar
First citationScheiner, S. (2023). Phys. Chem. Chem. Phys. 25, 7184–7194.  CrossRef CAS PubMed Google Scholar
First citationScheiner, S., Michalczyk, M. & Zierkiewicz, W. (2020). Coord. Chem. Rev. 405, 213136.  Web of Science CrossRef Google Scholar
First citationSchneider, F., Caramori, G., Parreira, R., Lippolis, V., Arca, M. & Ciancaleoni, G. (2017). Eur. J. Inorg. Chem. 2018, 1007–1015.  CrossRef Google Scholar
First citationScholfield, M. R., Zanden, C. M., Carter, M. & Ho, P. S. (2013). Protein Sci. 22, 139–152.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSchuchardt, K. L., Didier, B. T., Elsethagen, T., Sun, L., Gurumoorthi, V., Chase, J., Li, J. & Windus, T. L. (2007). J. Chem. Inf. Model. 47, 1045–1052.  Web of Science CrossRef PubMed CAS Google Scholar
First citationShukla, R., Claiser, N., Souhassou, M., Lecomte, C., Balkrishna, S. J., Kumar, S. & Chopra, D. (2018). IUCrJ, 5, 647–653.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationSiddiqui, R., Rani, J., Titi, H. M. & Patra, R. (2024). Coord. Chem. Rev. 517, 215994.  CrossRef Google Scholar
First citationSyzgantseva, O. A., Tognetti, V. & Joubert, L. (2013). J. Phys. Chem. A, 117, 8969–8980.  CrossRef CAS PubMed Google Scholar
First citationTantardini, C., Ceresoli, D. & Benassi, E. (2016). J. Comput. Chem. 37, 2133–2139.  CrossRef CAS PubMed Google Scholar
First citationThomas, S. P., Dikundwar, A. G., Sarkar, S., Pavan, M. S., Pal, R., Hathwar, V. R. & Row, T. N. G. (2022). Molecules, 27, 3690.  Web of Science CrossRef PubMed Google Scholar
First citationTodd, A. K. (2019). AIMAll, version 19.10.12. TK Gristmill Software, Overland Park KS, USA. https://aim.tkgristmill.comGoogle Scholar
First citationTognetti, V. & Joubert, L. (2016). Applications of Topological Methods in Molecular Chemistry, edited by R. Chauvin, C. Lepetit, B. Silvi & E. Alikhani, pp. 435–459. Cham: Springer International Publishing.  Google Scholar
First citationTognetti, V., Silva, A. F., Vincent, M. A., Joubert, L. & Popelier, P. L. A. (2018). J. Phys. Chem. A, 122, 7748–7756.  CrossRef CAS PubMed Google Scholar
First citationVaradwaj, P. R., Varadwaj, A., Marques, H. M. & Yamashita, K. (2024). Cryst. Growth Des. 24, 5494–5525.  CrossRef CAS Google Scholar
First citationVargas Jentzsch, A. & Matile, S. (2013). J. Am. Chem. Soc. 135, 5302–5303.  Web of Science CrossRef CAS PubMed Google Scholar
First citationVerteramo, M. L., Ignjatović, M. M., Kumar, R., Wernersson, S., Ekberg, V., Wallerstein, J., Carlström, G., Chadimová, V., Leffler, H., Zetterberg, F., Logan, D. T., Ryde, U., Akke, M. & Nilsson, U. J. (2024). iScience 27, 109636.  CrossRef PubMed Google Scholar
First citationVosko, S. H., Wilk, L. & Nusair, M. (1980). Can. J. Phys. 58, 1200–1211.  Web of Science CrossRef CAS Google Scholar
First citationWalker, M. G., Mendez, C. G. & Ho, P. S. (2023). Chem. Asian J. 18, e202300026.  CrossRef PubMed Google Scholar
First citationWang, C., Danovich, D., Mo, Y. & Shaik, S. (2014). J. Chem. Theory Comput. 10, 3726–3737.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWilcken, R., Zimmermann, M. O., Lange, A., Joerger, A. C. & Boeckler, F. M. (2013). J. Med. Chem. 56, 1363–1388.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWolters, L. P., Schyman, P., Pavan, M. J., Jorgensen, W. L., Bickelhaupt, F. M. & Kozuch, S. (2014). WIREs Comput. Mol. Sci. 4, 523–540.  Web of Science CrossRef CAS Google Scholar
First citationXiao, L. & Fu, H. (2019). Chem. A Eur. J. 25, 714–723.  CrossRef CAS Google Scholar
First citationXu, Z., Yang, Z., Liu, Y., Lu, Y., Chen, K. & Zhu, W. (2014). J. Chem. Inf. Model. 54, 69–78.  Web of Science CrossRef CAS PubMed Google Scholar
First citationZhao, Y. & Truhlar, D. G. (2008). Theor. Chem. Acc. 120, 215–241.  Web of Science CrossRef CAS Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

ISSN: 2052-2525