Next Article in Journal / Special Issue
Assessing Antipsoriatic Effects of Bitter Pu’er Tea and Its Three Major Compounds, Strictinin, Theacrine and Epigallocatechin Gallate, in Imiquimod-Treated Mice
Previous Article in Journal / Special Issue
Traditional Chinese Medicines and Prescriptions Brought from China to Japan by a Monk (Jianzhen, Japanese: Ganjin): A Historical Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Secondary Interactions on Structural Diversity between a Pair of Halogen-Bonded Co-Crystals Containing Isosteric Donors

1
Department of Biological Sciences, Webster University, St. Louis, MO 63119, USA
2
Department of Chemistry and Biochemistry, Missouri State University, Springfield, MO 65897, USA
3
Department of Chemistry and Biochemistry, Texas Tech University, Lubbock, TX 79409, USA
*
Author to whom correspondence should be addressed.
Compounds 2022, 2(4), 285-292; https://doi.org/10.3390/compounds2040023
Submission received: 9 September 2022 / Revised: 18 October 2022 / Accepted: 20 October 2022 / Published: 28 October 2022
(This article belongs to the Special Issue Feature Papers in Compounds (2022–2023))

Abstract

:
The formation of a pair of co-crystals based upon isosteric halogen-bond donors, namely 1,4-diiodoperchlorobenzene and iodoperchlorobenzene, along with the acceptor 4,4-bipyridine is reported. As expected, the components in each co-crystal engage in halogen bonding interactions resulting in a one-dimensional chain-like structure. In particular, the co-crystal containing 1,4-diiodoperchlorobenzene is primarily held together by I···N halogen bonds while the solid based upon iodoperchlorobenzene forms both I···N and Cl···N interactions. Structural diversity is achieved between these co-crystals based upon the type of secondary interactions involving the chlorine atoms on each halogen-bond donor even though they are isosteric in nature.

1. Introduction

Halogen bonding is an attractive interaction between an electrophilic region on a halogen atom, namely a σ-hole, and a nucleophilic region on a different atom [1,2,3]. In particular, a σ-hole is a positive electrostatic area found at the tip of a carbon-bound halogen atom that can interact with an electron rich group such as a lone pair [4,5]. The σ-hole is most pronounced on iodine and is enhanced by neighboring electronegative atoms, particularly fluorine [6,7,8,9]. Solid state chemists utilize halogen bonding interactions since they are similar in strength and directionality when compared to hydrogen bonds. A continuing challenge in crystal engineering is how small and incremental changes, to either component of the co-crystal, can lead to structural diversity between these solids [10,11,12,13]. In these systems, the variations occur due to the type of secondary interactions that ultimately influence the resulting crystal structure, along with chemical and physical properties, even in the presence of the stronger halogen bond [14].
An ongoing focus in our research groups is to develop halogen-bond donors that contain chlorine rather than fluorine as the electron-withdrawing group. In particular, we have investigated the halogen bonding propensity of 1,4-diiodoperchlorobenzene (C6I2Cl4) where it has shown to reliably form I···N halogen bonds when coupled with various acceptor molecules [15,16,17,18,19]. To understand the formation of these halogen bonds, a molecular electrostatic potential calculation was performed on C6I2Cl4 and a σ-hole was located on the iodine atoms with a value of 146 kJ/mol which is well within range for a halogen-bond donor (Scheme 1) [19]. With a similar goal, we have also reported on the ability of iodoperchlorobenzene (C6ICl5) to engage in I···N halogen bonds with many of these same acceptors [20]. The para-chlorine atom on C6ICl5, in reference to the iodine, forms an unexpected Cl···N halogen bond which generated a one-dimensional chain-like structure when combined with appropriate ditopic acceptors. In a similar calculation, the σ-hole on both the iodine and para-chlorine atoms on C6ICl5 supported halogen bond formation with values of 157 and 78 kJ/mol, respectively (Scheme 1) [20]. Curiously, these co-crystals are all isostructural to related solids containing C6I2Cl4 as the halogen-bond donor.
Using this as inspiration, we report here the formation and structure of a pair of halogen-bonded co-crystals based upon 4,4-bipyridine (4,4-BP) along with two isosteric donors, namely C6I2Cl4 and C6ICl5 (Scheme 1). As expected, the resulting co-crystals (C6I2Cl4)·(4,4-BP) and (C6ICl5)·(4,4-BP) are primarily held together by the combination of I···N and Cl···N halogen bonds which generate a one-dimensional chain structure for each solid. Even though these co-crystals contain isosteric donors, which have previously yielded isostructural solids, the resulting materials have drastically different crystal structures. This structural diversity is based upon the type of secondary non-covalent interactions observed between the components involving chlorine atoms on a particular halogen-bond donor.

2. Materials and Methods

2.1. Materials

The acceptor 4,4-bipyridine (4,4-BP) and reagent grade toluene were both purchased from Sigma-Aldrich Chemical (St. Louis, MO, USA) and used as received. Both of the donors, 1,4-diiodoperchlorobenzene (C6I2Cl4) [21] and iodoperchlorobenzene (C6ICl5) [22], were synthesized by a previously reported method. All crystallization experiments were performed in 20 mL scintillation vials.

2.2. Formation of (C6I2Cl4)·(4,4-BP)

Co-crystals of (C6I2Cl4)·(4,4-BP) were formed by dissolving 25.0 mg of C6I2Cl4 in 2.0 mL of toluene, which was combined with a separate 2.0 mL toluene solution containing 8.3 mg of 4,4-BP (1:1 molar equivalent). The resulting solution was allowed to slowly evaporate and within two days, crystals suitable for X-ray diffraction were realized.

2.3. Formation of (C6ICl5)·(4,4-BP)

In a similar manner, co-crystals of (C6ICl5)·(4,4-BP) were achieved by taking 25.0 mg of C6ICl5 and dissolving it in 2.0 mL of toluene. Then, a separate 2.0 mL toluene solution of 10.4 mg of 4,4-BP was combined with the previous solution (1:1 molar equivalent). Again, the combined solution was allowed to slowly evaporate and within three days, single crystals suitable for X-ray diffraction were formed.

2.4. Single-Crystal X-ray Diffraction

X-ray data were collected on a Rigaku XtaLAB Synergy-i Kappa diffractometer (Rigaku Americas Corporation, The Woodlands, TX, USA) equipped with a PhotonJet-i X-ray source to generate Cu Kα radiation. Suitable single crystals were transferred to a glass slide in type NVH immersion oil. Each co-crystal was then mounted on a MiTeGen 50 µm MicroLoop and placed on the diffractometer under a cold nitrogen stream at 100 K. After data collection, the unit cell was redetermined using a subset of all the collected data. The intensity data were corrected for Lorentz, polarization, and background effects using CrysAlisPro (Rigaku Americas Corporation, The Woodlands, TX, USA). A numerical absorption correction was applied based on a Gaussian integration over a multi-faceted crystal then followed by a semi-empirical correction for adsorption. The program SHELXT [23] was used for the initial structure solution while SHELXL [24] was used for the refinement of each co-crystal. These programs were utilized within the OLEX2 software (OlexSys Ltd, Durham, UK) [25]. Hydrogen atoms bound to carbons were geometrically constrained using the appropriate AFIX commands. Selected crystallographic and refinement parameters for (C6I2Cl4)·(4,4-BP) and (C6ICl5)·(4,4-BP) are listed in Table 1.

2.5. Computational Methods

To determine the different I···N and Cl···N halogen-bonding energies, a series of Density Functional Theory (DFT) calculations were performed using the M06-2X density functional in the Gaussian 16 program [26]. The aug-cc-pVTZ basis set, stored in the Gaussian program, was used on all atoms except for iodine. In the case of iodine, the basis set, which included a core potential that replaces the inner 28 electrons, was obtained from the EMSL Basis Set Exchange Library [27]. This approach computes the binding energy as the difference between the energy of the co-crystal and the energies of the separated molecules. These DFT calculations used a single point energy computation with convergence for the energy change of less than 1.00 × 10−6 Hartrees. All of these binding energies were computed using the counterpoise method. The application of the counterpoise correction always decreases the binding energies by typically five to ten percent. In addition, these calculations were also carried out using non-augmented basis functions. When removing the diffuse functions, a change of one to two percent in the dissociation energy results.

3. Results

3.1. X-ray Crystal Structure of (C6I2Cl4)·(4,4-BP)

The components of (C6I2Cl4)·(4,4-BP) crystallize in the centrosymmetric triclinic space group Pī (Table 1). Within the asymmetric unit is half a molecule of both C6I2Cl4 and 4,4-BP where inverse symmetry generates the remainder of each fragment. As expected, C6I2Cl4 forms a series of I···N halogen bonds [I···N 2.904(6) Å; C–I···N 171.3(2)°] with 4,4-BP which generates a one-dimensional chain-like structure (Figure 1). Due to symmetry, the aromatic rings within 4,4-BP are coplanar. In contrast, the aromatic rings between the donor and acceptor within (C6I2Cl4)·(4,4-BP) are twisted with an angle of 56.61° (Figure 1 and Figure 2). As seen with other symmetric ditopic halogen-bond acceptors, molecules of C6I2Cl4 engage in infinite homogeneous face-to-face π–π stacking (Figure 2). The parallel and slipped orientation of the aromatic ring on C6I2Cl4 runs along the crystallographic a axis with a centroid-to-centroid distance of 4.0843(1) Å equal to that axis. Similar π–π stacking patterns were observed for C6I2Cl4 in both single- and mutli-component crystals that have been previously reported by our groups [16,19].
These halogen-bonded chains interact with their nearest neighbors by Type I Cl···Cl interactions [28,29] [Cl···Cl 3.240(2) Å; C–Cl···Cl 165.8(2)°; |θ1 − θ2| = 0°] which results in a two-dimensional sheet (Figure 3). Type I interactions are not considered halogen bonds but instead they are symmetric close contacts between halogen atoms where the difference in bond angles will be equal to zero. In particular, half of the chlorine atoms on C6I2Cl4 are found to interact in this type of secondary interaction. Lastly, the remaining chlorine atoms are found to engage in an offset Cl···π interactions with 4,4-BP (Cl···π 4.092 Å) measured from the chlorine atom to the centroid of the 4-pyridyl ring. The closest distance for this Cl···π interactions is found between a chlorine and carbon atom with a distance of 3.212(5) Å along with an angle of 170.7(2)°.

3.2. X-ray Crystal Structure of (C6ICl5)·(4,4-BP)

The components of (C6ICl5)·(4,4-BP) crystallize in the centrosymmetric monoclinic space group P21/n where a half of both molecules are found in the asymmetric unit (Table 1). Again, inversion symmetry generates the remainder of each molecule. Unlike before, molecules of 4,4-BP are found to be disordered over two positions where after a free variable refinement returned a final value of .51/.49 for the major/minor components. In addition, the iodine and para-chlorine atoms, with respect to the iodine, are equally disordered over two positions and were modeled with a .50/.50 occupancy for each atom. The ability of C6ICl5 to form both I···N [I···N 2.956(7) Å; C–I···N 162.6(3)°] and Cl···N halogen bonds [Cl···N 3.205(20) Å; C–Cl···N 158.1(9)°] to the major site for 4,4-BP results in a one-dimensional chain (Figure 4). In drastic contrast to (C6I2Cl4)·(4,4-BP), the aromatic rings, namely the halogen-bond donor and acceptor, within (C6ICl5)·(4,4-BP) are nearly co-planar with an angle of 7.24° with regard to the major orientation (Figure 5). Again, molecules of C6ICl5 engage in a face-to-face π–π stacking arrangement that results in an infinite column of the donor that runs along the crystallographic b axis (Figure 5 and Figure 6). These donors stack in a parallel and slightly offset pattern with a centroid-to-centroid distance of 3.8878(1) Å which is equal to that axis (Table 1).
Unlike (C6I2Cl4)·(4,4-BP), the secondary interaction between neighboring C6ICl5 stacks within (C6ICl5)·(4,4-BP) is an infinite chain of Cl···Cl contacts. These are similar to the trifurcated X3 synthon recognized in the structures of perhalobenzenes and trihalomesitylenes [21,30]. In this structure, however, the infinite array of these non-covalent contacts [Cl···Cl 3.528(3) Å; C–Cl···Cl 148.1(3)° and 121.8(3)°; |θ1 − θ2| = 26.3(3)°] appears between adjacent stacks of the donor along the crystallographic b axis (Figure 6). This closest chlorine–chlorine distance cannot be classified as either Type I or Type II due to the observed C–Cl···Cl bond angles [29]. In particular, Type I interactions should have a difference in bond angles equal to 0° while for Type II, the pair of angles should be close to 180° and 90°. Only half of the chlorine atoms on C6ICl5 are found to interact in this type of Cl···Cl interaction.

3.3. Halogen Bond Energies Using Density Functional Theory Calculations

To enumerate the strength of both I···N and Cl···N halogen bonds within each co-crystal, a series of theoretical investigations using Density Functional Theory (DFT) calculations were performed. In particular, the M062X density functional was employed along with an aug-cc-pVTZ basis set. These halogen-bond strengths were calculated by using atomic positions determined from single-crystal X-ray diffraction data. The I···N halogen-bonding energy within (C6I2Cl4)·(4,4-BP) was determined to be −22.0 kJ/mol. In contrast, the I···N halogen bonding value for (C6ICl5)·(4,4-BP) yielded a lower value of −17.9 kJ/mol. In a similar approach, the strength of the Cl···N halogen bond within (C6ICl5)·(4,4-BP) was determined to be −7.4 kJ/mol. Interestingly, both of these energies are slightly lower than the published values for the co-crystal containing C6ICl5 and trans-1,2-bi(4-pyridyl)ethylene which had values of −19.0 and −8.5 kJ/mol for the I···N and Cl···N halogen bonds, respectively [20]. These calculated binding energies support the presence of halogen bonds within these co-crystals.

4. Conclusions

In this contribution, we report the structural diversity between a pair of co-crystals that differ only by the isosteric halogen-bond donor. This diversity is achieved by the type of secondary non-covalent interactions formed by the chlorine atoms on the given donor. In particular, these one-dimensional halogen-bonded chains either interact with neighbors via Type I Cl···Cl interaction or an infinite Cl···Cl contact that gives rise to the particular crystal packing. Currently, we are investigating the structures of related co-crystals with these donors and other symmetric bipyridine acceptors.

Author Contributions

Conceptualization, R.H.G.; methodology, H.R.K.J., E.B. and R.H.G.; formal analysis, H.R.K.J., D.K.U. and R.H.G.; investigation, H.R.K.J., N.M.S. and D.K.U.; writing—original draft preparation, R.H.G.; writing—review and editing, H.R.K.J., N.M.S., E.B., D.K.U. and R.H.G.; supervision, R.H.G.; funding acquisition, R.H.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

Webster University is acknowledged for financial support in the form of various Faculty Research Grants.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Varadwaj, P.R.; Varadwaj, A.; Marques, H.M. Halogen Bonding: A Halogen-Centered Noncovalent Interaction Yet to Be Understood. Inorganics 2019, 7, 40. [Google Scholar] [CrossRef] [Green Version]
  2. Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. The Halogen Bond. Chem. Rev. 2016, 116, 2478–2601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Gilday, L.C.; Robinson, S.W.; Barendt, T.A.; Langton, M.J.; Mullaney, B.R.; Beer, P.D. Halogen Bonding in Supramolecular Chemistry. Chem. Rev. 2015, 115, 7118–7195. [Google Scholar] [CrossRef] [PubMed]
  4. Robidas, R.; Reinhard, D.L.; Legault, C.Y.; Huber, S.M. Iodine(III)-Based Halogen Bond Donors: Properties and Applications. Chem. Rec. 2021, 21, 1912–1927. [Google Scholar] [CrossRef] [PubMed]
  5. Clark, T.; Hennemann, M.; Murray, J.S.; Politzer, P. Halogen bonding: The σ-hole. J. Mol. Model 2007, 13, 291–296. [Google Scholar] [CrossRef]
  6. Politzer, P.; Murray, J.S.; Clark, T.; Resnati, G. The σ-hole revisited. Phys. Chem. Chem. Phys. 2017, 19, 32166–32178. [Google Scholar] [CrossRef]
  7. Awwadi, F.F.; Taher, D.; Kailani, M.H.; Alwahsh, M.I.; Odeh, F.; Rüffer, T.; Schaarschmidt, D.; Lang, H. Halogen Bonding Interactions in Halopyridine–Iodine Monochloride Complexes. Cryst. Growth Des. 2020, 20, 543–551. [Google Scholar] [CrossRef]
  8. Robinson, S.W.; Mustoe, C.L.; White, N.G.; Brown, A.; Thompson, A.L.; Kennepohl, P.; Beer, P.D. Evidence for Halogen Bond Covalency in Acyclic and Interlocked Halogen-Bonding Receptor Anion Recognition. J. Am. Chem. Soc. 2015, 137, 499–507. [Google Scholar] [CrossRef]
  9. Varadwaj, A.; Varadwaj, P.R.; Marques, H.M.; Yamashita, K. A DFT assessment of some physical properties of iodine-centered halogen bonding and other non-covalent interactions in some experimentally reported crystal geometries. Phys. Chem. Chem. Phys. 2018, 20, 15316–15329. [Google Scholar] [CrossRef]
  10. Mandal, S.; Pal, A.K.; Mandal, N.; Mukhopadhyay, T.K.; Datta, A. Harnessing Noncovalent Interactions for a Directed Evolution of a Six-Component Molecular Crystal. J. Phys. Chem. B 2021, 125, 12584–12591. [Google Scholar] [CrossRef]
  11. Martí-Rujas, J.; Kariuki, B.M.; Hughes, C.E.; Morte-Ródenas, A.; Guo, F.; Glavcheva-Laleva, Z.; Taştemür, K.; Ooi, L.; Yeo, L.; Harris, K.D.M. Structural diversity, but no polymorphism, in a homologous family of co-crystals of urea and α,ω-dihydroxyalkanes. New J. Chem. 2011, 35, 1515–1521. [Google Scholar] [CrossRef]
  12. Alhalaweh, A.; George, S.; Boström, D.; Velaga, S.P. 1:1 and 2:1 Urea-Succinic Acid Cocrystals: Structural Diversity, Solution Chemistry and Thermodynamic Stability. Cryst. Growth Des. 2010, 10, 4847–4855. [Google Scholar] [CrossRef]
  13. Yelgaonkar, S.P.; Kiani, D.; Baltrusaitis, J.; MacGillivray, L.R. Superstructural diversity in salt-cocrystals: Higher-order hydrogen-bonded assemblies formed using U-shaped dications and with assistance of π−π stacking. Chem. Commun. 2020, 56, 6708–6710. [Google Scholar] [CrossRef] [PubMed]
  14. Hasija, A.; Hasija, A.; Chopra, D. Utilizing Co-Crystallization as a Tool to Unravel the Structural Diversity and Electronic Features of I···S Halogen Bonded Interactions in Stoichiomorphic Co-Crystals. Cryst. Growth Des. 2020, 20, 6272–6282. [Google Scholar] [CrossRef]
  15. Bosch, E.; Ferrence, G.M.; Powell, C.J.; Unruh, D.K.; Krueger, H.R.; Groeneman, R.H. Cooperative non-covalent interactions and synthetic feed as driving forces to structural diversity within organic co-crystals containing isosteric perhalobenzenes. CrystEngComm 2022, 24, 3841–3845. [Google Scholar] [CrossRef]
  16. Kruse, S.J.; Bosch, E.; Brown, F.; Groeneman, R.H. Incorporating Ester Functionality within a Solid-State [2 + 2] Cycloaddition Reaction Based Upon Halogen Bonding Interactions. Cryst. Growth Des. 2020, 20, 1969–1974. [Google Scholar] [CrossRef]
  17. Sinnwell, M.A.; Santana, C.L.; Bosch, E.; MacGillivray, L.R.; Groeneman, R.H. Application of a tetrapyrimidyl cyclobutane synthesized in the organic solid state: A halogen-bonded supramolecular ladder. CrystEngComm 2020, 22, 6780–6782. [Google Scholar] [CrossRef]
  18. Bosch, E.; Kruse, S.J.; Reinheimer, E.W.; Rath, N.P.; Groeneman, R.H. Regioselective [2 + 2] cycloaddition reaction within a pair of polymorphic co-crystals based upon halogen bonding interactions. CrystEngComm 2019, 21, 6671–6675. [Google Scholar] [CrossRef]
  19. Bosch, E.; Kruse, S.J.; Krueger, H.R.; Groeneman, R.H. Role of π−π Stacking and Halogen Bonding by 1,4-Diiodoperchlorobenzene to Organize the Solid State To Achieve a [2 + 2] Cycloaddition Reaction. Cryst. Growth Des. 2019, 19, 3092–3096. [Google Scholar] [CrossRef]
  20. Shapiro, N.M.; Bosch, E.; Unruh, D.K.; Krueger, H.R.; Groeneman, R.H. Iodoperchlorobenzene acts as a dual halogen-bond donor to template a [2 + 2] cycloaddition reaction within an organic co-crystal. CrystEngComm 2021, 23, 8265–8268. [Google Scholar] [CrossRef]
  21. Reddy, C.M.; Kirchner, M.T.; Gundakaram, R.C.; Padmannabhan, K.A.; Desiraju, G.R. Isostructurality, Polymorphism and Mechanical Properties of Some Hexahalogenated Benzenes: The Nature of Halogen···Halogen Interactions. Chem. Eur. J. 2006, 12, 2222–2234. [Google Scholar] [CrossRef] [PubMed]
  22. Do, H.-Q.; Dauglis, O. A simple base-mediated halogenation of acidic sp2 C-H bonds under noncryogenic conditions. Org. Lett. 2009, 11, 421–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  24. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. 2015, C71, 3–8. [Google Scholar]
  25. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  26. 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.; et al. Gaussian 16, Revision C.01.; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  27. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus. T.L. A New Basis Set Exchange: An Open, Up-to-date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59, 4814–4820. [Google Scholar] [CrossRef]
  28. Murray, J.S.; Resnati, G.; Politzer, P. Close contacts and noncovalent interactions in crystals. Faraday Discuss. 2017, 203, 113–130. [Google Scholar] [CrossRef]
  29. Mukherjee, A.; Tothadi, S.; Desiraju, G.R. Halogen Bonds in Crystal Engineering: Like Hydrogen Bonds yet Different. Acc. Chem. Res. 2014, 47, 2514–2524. [Google Scholar] [CrossRef]
  30. Bosch, E.; Barnes, C.L. Triangular halogen−halogen−halogen interactions as a cohesive force in the structures of trihalomesitylenes. Cryst. Growth Des. 2002, 4, 299–302. [Google Scholar] [CrossRef]
Scheme 1. Rendering of the components of these co-crystals: (a) halogen-bond donors C6I2Cl4 and C6ICl5 along with (b) the halogen-bond acceptor 4,4-BP. The electrostatic potential values for the σ-hole on selected atoms on the halogen-bond donors are shown as previously reported [19,20]. All indicated values are in kJ/mol.
Scheme 1. Rendering of the components of these co-crystals: (a) halogen-bond donors C6I2Cl4 and C6ICl5 along with (b) the halogen-bond acceptor 4,4-BP. The electrostatic potential values for the σ-hole on selected atoms on the halogen-bond donors are shown as previously reported [19,20]. All indicated values are in kJ/mol.
Compounds 02 00023 sch001
Figure 1. X-ray structure of (C6I2Cl4)·(4,4-BP) illustrating the I···N halogen bonds which generate the one-dimensional chain-like structure.
Figure 1. X-ray structure of (C6I2Cl4)·(4,4-BP) illustrating the I···N halogen bonds which generate the one-dimensional chain-like structure.
Compounds 02 00023 g001
Figure 2. X-ray structure of (C6I2Cl4)·(4,4-BP) illustrating the infinite homogeneous face-to-face π–π stacking arrangement of the aromatic rings.
Figure 2. X-ray structure of (C6I2Cl4)·(4,4-BP) illustrating the infinite homogeneous face-to-face π–π stacking arrangement of the aromatic rings.
Compounds 02 00023 g002
Figure 3. X-ray structure of (C6I2Cl4)·(4,4-BP) illustrating the Type I Cl···Cl and Cl···π interactions between nearest neighboring chains.
Figure 3. X-ray structure of (C6I2Cl4)·(4,4-BP) illustrating the Type I Cl···Cl and Cl···π interactions between nearest neighboring chains.
Compounds 02 00023 g003
Figure 4. X-ray structure of (C6ICl5)·(4,4-BP) illustrating the I···N and Cl···N halogen bonds which generate the one-dimensional chain-like structure. The crystallographic disorder observed in both C6ICl5 and 4,4-BP are shown. In particular, the observed disorder in C6ICl5 is illustrated with the two colors (purple and green) on the same molecule.
Figure 4. X-ray structure of (C6ICl5)·(4,4-BP) illustrating the I···N and Cl···N halogen bonds which generate the one-dimensional chain-like structure. The crystallographic disorder observed in both C6ICl5 and 4,4-BP are shown. In particular, the observed disorder in C6ICl5 is illustrated with the two colors (purple and green) on the same molecule.
Compounds 02 00023 g004
Figure 5. X-ray structure of (C6ICl5)·(4,4-BP) illustrating the infinite homogeneous face-to-face π–π stacking arrangement of the aromatic rings. The observed disorder in C6ICl5 is illustrated with the two colors (purple and green) on the same atom. The disorder in 4,4-BP was removed for clarity.
Figure 5. X-ray structure of (C6ICl5)·(4,4-BP) illustrating the infinite homogeneous face-to-face π–π stacking arrangement of the aromatic rings. The observed disorder in C6ICl5 is illustrated with the two colors (purple and green) on the same atom. The disorder in 4,4-BP was removed for clarity.
Compounds 02 00023 g005
Figure 6. X-ray structure of (C6ICl5)·(4,4-BP) illustrating the infinite homogeneous face-to-face π–π stacking arrangement along with the infinite chain of Cl···Cl contacts. The observed disorder in C6ICl5 is illustrated with the two colors (purple and green) on the same atom.
Figure 6. X-ray structure of (C6ICl5)·(4,4-BP) illustrating the infinite homogeneous face-to-face π–π stacking arrangement along with the infinite chain of Cl···Cl contacts. The observed disorder in C6ICl5 is illustrated with the two colors (purple and green) on the same atom.
Compounds 02 00023 g006
Table 1. Crystallographic and refinement parameters for the co-crystals (C6I2Cl4)·(4,4-BP) and (C6ICl5)·(4,4-BP).
Table 1. Crystallographic and refinement parameters for the co-crystals (C6I2Cl4)·(4,4-BP) and (C6ICl5)·(4,4-BP).
Co-Crystal(C6I2Cl4)·(4,4-BP)(C6ICl5)·(4,4-BP)
FormulaC16H8Cl4I2N2C16H8Cl5IN2
Formula Mass (g·mol−1) 623.84532.39
Crystal systemtriclinicmonoclinic
Space groupPīP21/n
a (Å)4.0843(1)14.3260(3)
b (Å)9.4089(2)3.8878(1)
c (Å)12.8307(2)17.0394(4)
α (°)68.543(2)90
β (°)84.458(2)113.747(3)
γ (°)87.780(2)90
Z12
V3)456.747(18)868.68(4)
ρcalcd (g·cm−3)2.2682.035
T (K)100100
μ (mm−1)32.43621.570
F(000)292.0512.0
Radiation sourceCu KαCu Kα
Reflections collected80626948
Independent reflections18001617
Data/restraints/parameters1800/0/1101617/204/151
Rint0.05880.0573
R1 (I ≥ 2σ(I))0.03200.0673
wR (F2) (I ≥ 2σ(I))0.08400.1877
R1 (all data)0.03260.0686
wR (F2) (all data)0.08440.1895
Goodness-of-net on F21.0801.074
CCDC deposition number2,204,8442,204,845
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Krueger, H.R., Jr.; Shapiro, N.M.; Bosch, E.; Unruh, D.K.; Groeneman, R.H. Influence of Secondary Interactions on Structural Diversity between a Pair of Halogen-Bonded Co-Crystals Containing Isosteric Donors. Compounds 2022, 2, 285-292. https://doi.org/10.3390/compounds2040023

AMA Style

Krueger HR Jr., Shapiro NM, Bosch E, Unruh DK, Groeneman RH. Influence of Secondary Interactions on Structural Diversity between a Pair of Halogen-Bonded Co-Crystals Containing Isosteric Donors. Compounds. 2022; 2(4):285-292. https://doi.org/10.3390/compounds2040023

Chicago/Turabian Style

Krueger, Herman R., Jr., Nicole M. Shapiro, Eric Bosch, Daniel K. Unruh, and Ryan H. Groeneman. 2022. "Influence of Secondary Interactions on Structural Diversity between a Pair of Halogen-Bonded Co-Crystals Containing Isosteric Donors" Compounds 2, no. 4: 285-292. https://doi.org/10.3390/compounds2040023

Article Metrics

Back to TopTop