Next Article in Journal
Chemical Composition, Crystal Structure, and Microstructure of Slags on the Korean Peninsula from the First Copper Production Remains of the 9th Century
Previous Article in Journal
Enhanced Strength–Ductility Synergy of Mg-Al-Sn-Ca Alloy via Composite Asymmetric Extrusion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preference in the Type of Halogen Bonding Interactions within Co-Crystals of Anthraquinone with a Pair of Isosteric Perhalobenzenes

1
Department of Chemistry and Biochemistry, Missouri State University, Springfield, MO 65897, USA
2
Office of the Vice President for Research, University of Iowa, Iowa City, IA 52242, USA
3
Department of Natural Sciences and Mathematics, Webster University, St. Louis, MO 63119, USA
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(4), 325; https://doi.org/10.3390/cryst14040325
Submission received: 13 March 2024 / Revised: 26 March 2024 / Accepted: 28 March 2024 / Published: 30 March 2024
(This article belongs to the Section Crystal Engineering)

Abstract

:
The preference in the type of halogen bond accepted by anthraquinone (C14H8O2) from two isosteric donors, namely 1,4-diiodoperfluorobenzene (C6I2F4) and 1,4-diiodoperchlorobenzene (C6I2Cl4), is reported. The two co-crystals, (C6I2F4)·(C14H8O2) and (C6I2Cl4)·(C14H8O2), are sustained primarily by I···O rather than π-type halogen bonds to form these multicomponent solids. The ability for each component to engage in two divergent halogen-bonding interactions generates a one-dimensional chain structure for each co-crystal. The bias in the halogen-bonding type is due to the difference in electrostatic potential between the carbonyl oxygen and the aromatic surface on the anthraquinone. To support this observed preference, the binding energies of the I···O halogen bond were quantified for both co-crystals by using density functional theory calculations and then compared to the interaction energy for related π-type halogen bond from previously reported structures.

1. Introduction

Halogen bonding is an attractive non-covalent interaction between an electrophilic region on a halogen atom and a nucleophilic region on a different atom [1,2]. This electrophilic or positive region is defined as the σ-hole, which is found along the bond axis between a halogen and carbon atom. In general, the σ-hole is enhanced, with a larger positive value, when iodine is in the presence of neighboring electronegative atoms [3,4]. As a result of electrostatic attraction, the σ-hole interacts with an electron-rich atom or group to form a halogen bond. Normally, these halogen bonds are classified as n-type, where the σ-hole is interacting with a lone pair on a heteroatom such as nitrogen or oxygen. In contrast, π-type halogen bonds are observed between an electrophilic halogen atom and a polycyclic aromatic surface and are considerably less common [5,6].
The most investigated and reported aromatic halogen-bond donor continues to be 1,4-diiodoperfluorobenzene (C6I2F4). This is due to the greater electron-withdrawing ability of fluorine when compared to other halogen atoms. As a result, the fluorine atoms within C6I2F4 generate the largest positive σ-hole on the iodine atoms, making this an ideal aromatic donor. Due to this fact, C6I2F4 has been shown to engage in both n-type [7,8,9] and π-type [10,11] halogen bonds to form numerous co-crystals or multi-component molecular solids.
A continued focus of our research groups has been the design and formation of co-crystals that utilize 1,4-diiodoperchlorobenzene (C6I2Cl4) as a halogen-bond donor [12,13,14]. In particular, we have investigated the ability of C6I2Cl4 to behave as a template to align various pyridine-based reactants in a suitable position to undergo the solid-state [2 + 2] cycloaddition reaction [15,16,17]. In these co-crystals, the constituent molecules are held together by I···N (i.e., n-type) halogen bonds to form photoreactive solids. Unlike co-crystals containing C6I2F4, an advantage of using C6I2Cl4 as a template is its tendency to engage in homogeneous and face-to-face π-π stacking interactions [18] that, along with the I···N halogen bond, position a pair of reactant molecules within the correct distance and orientation for a photoreaction [19]. Recently, we also investigated the ability of C6I2Cl4 to form a π-type halogen bond to a polyaromatic molecule, namely naphthalene [20]. The initial selection of naphthalene was based upon its polyaromatic structure, along with a lack of a heteroatom, precluding competition for the π-type halogen bond within the co-crystal. To expand this research further and to understand the tendencies of C6I2F4 and C6I2Cl4 as donors, we investigated what type of halogen bond would be preferred when co-crystallized with a polycyclic aromatic that contains heteroatoms.
Using this as inspiration, we report here the structure of a pair of co-crystals based upon anthraquinone (C14H8O2) with either C6I2F4 or C6I2Cl4, where the I···O halogen bond is preferred over π-type interactions in these molecular solids (Scheme 1). These co-crystals have similar formulas, namely (C6I2F4)·(C14H8O2) and (C6I2Cl4)·(C14H8O2), where, in both cases, the I···O halogen bonds generate a one-dimensional chain structure. Curiously, the co-crystals are not isostructural even though they contain a molecule in common and the halogen-bond donors are isosteric (i.e., similar shape) in nature. In particular, neighboring chains in (C6I2F4)·(C14H8O2), the donor engages in both C-H···F and C=O···π interactions with the acceptor. In addition, molecules of C14H8O2 are found to π-π stack in an offset and face-to-face pattern in the co-crystal. In contrast, the chains within (C6I2Cl4)·(C14H8O2) for both the donor and acceptor molecules interact with its nearest neighbors by homogeneous and face-to-face π-π stacking forces, as seen with other co-crystals containing C6I2Cl4. Lastly, these polymeric chains interact via Type I chlorine–chlorine interactions to form an extended structure.
The preference of the I···O halogen bond over the π-type interaction in these co-crystals is based upon the larger negative electrostatic potential on the oxygen atom when compared to the aromatic surface of the acceptor. While electrostatic potentials are a useful tool in the understanding and design of halogen-bonded systems, the role of charge transfer in halogen bonding has garnered greater recognition in the area of crystal engineering [21,22]. Finally, the selectivity in the type of halogen bond was investigated by performing a series of density functional theory (DFT) calculations to determine the binding energy of the I···O halogen bond in both co-crystals. These values were then compared to the binding energies for π-type halogen bonds for related co-crystals, illustrating that I···O halogen bonds form stronger non-covalent interactions and, in turn, are observed within these molecular solids.

2. Materials and Methods

2.1. Materials

The donor, 1,4-diiodoperfluorobenzene (C6I2F4), and the acceptor, anthraquinone (C14H8O2), were both purchased from Sigma-Aldrich Chemical (St. Louis, MO, USA) and were used without purification. The two solvents, namely toluene and chloroform, were also purchased from Sigma-Aldrich Chemical and were used as received. The donor, 1,4-diiodoperchlorobenzene (C6I2Cl4), was synthesized by a previously reported method [23]. All of the crystallization studies were performed in 20 mL scintillation vials.

2.2. Formation of (C6I2F4)·(C14H8O2)

The formation of the co-crystal (C6I2F4)·(C14H8O2) was achieved by dissolving 25.0 mg of C6I2F4 in 2.0 mL of chloroform, which was then added to a separate solution of 12.9 mg of C14H8O2 in 2.0 mL of chloroform (1:1 molar equivalent). The combined solution was then allowed to evaporate slowly. Within two days, after some loss of solvent, single crystals suitable for X-ray diffraction were formed.

2.3. Formation of (C6I2Cl4)·(C14H8O2)

The formation of the co-crystal (C6I2Cl4)·(C14H8O2) was achieved by dissolving 25.0 mg of C6I2Cl4 in 2.0 mL of toluene, which was then combined with 11.2 mg of C14H8O2 in a separate 2.0 mL toluene solution (1:1 molar equivalent). The resulting solution was then allowed to slowly evaporate. Within a day, and along with loss of solvent, single crystals suitable for X-ray diffraction were realized.

2.4. Electrostatic Potential Calculations

The molecular electrostatic potential energy surface for C14H8O2 was calculated by using Spartan’20 molecular modeling program, using density functional theory (DFT) at the B3LYP/6- 311++G** level [24]. Previously, we reported the calculated values of the σ-hole on iodine within both C6I2F4 and C6I2Cl4 to be 168.9 and 145.7 kJ/mol, respectively [15]. Both of these values are well within range for a halogen-bond donor. With respect to C14H8O2, the electrostatic potential energies were determined and returned a minimum value of −154.3 kJ/mol on the carbonyl oxygen, atom along with a minimum of −40 kJ/mol for the aromatic ring (Scheme 2).
With regard to the perhalobenzenes, namely C6I2F4 and C6I2Cl4, the aromatic rings are electron poor in nature and may potentially accept C=O···π interactions [25,26,27]. To this end, the maximum electrostatic potential on the surface of the benzene ring within C6I2F4 has a value of 81 kJ/mol at a point between two fluorine-bearing carbon atoms. In contrast, the largest electrostatic potential on C6I2Cl4 is only 54 kJ/mol and is located in the center of the benzene ring.

2.5. Computational Methods

To determine the binding energies for the I···O and π-type halogen bonds, a series of density functional theory (DFT) calculations were performed using the M06-2X density functional in the Gaussian 16 program [28]. The aug-cc-pVTZ basis set, stored in the Gaussian program, was used on all atoms except iodine. In the case of iodine, the basis set included a core potential that replaces the inner 28 electrons and was obtained from the EMSL Basis Set Exchange Library [29]. This DFT approach determines the binding energy as the difference between the acceptor and donor in the co-crystal and the two separated molecules. These binding energies were all computed using the counterpoise method, which normally decreases the value by around five-to-ten percent. The overall I···O and π-type halogen-bond energies for a given co-crystal were calculated by using atomic positions determined from the single-crystal X-ray diffraction data.

2.6. Single-Crystal X-ray Diffraction

X-ray data were collected on a Bruker D8 VENTURE DUO diffractometer (Bruker, Germany), using Mo Kα radiation (λ = 0.71073 Å) and a PHOTON III detector. Crystals were transferred from the vial and placed on a glass slide in Paratone-N oil. A crystal, along with a small amount of oil, was collected on a MiTeGen 100-micron MicroLoop. Then, the selected crystal was transferred to the instrument, where it was placed under a cold nitrogen stream at 100 K. After data collection, the unit cell was determined using a subset of all the collected data. All intensity data were corrected for Lorentz, polarization, and background effects, using the program APEX5 (version 2023.9-2). A semi-empirical correction was applied for the adsorption, using SADABS. The program SHELXT [30] was used for the initial structure solution, while SHELXL [31] was used for the refinement of the structure. Both programs were utilized within the OLEX2 software [32]. Selected crystallographic and refinement parameters for both (C6I2F4)·(C14H8O2) and (C6I2Cl4)·(C14H8O2) are listed in Table 1.

3. Results

3.1. X-ray Crystal Structure of (C6I2F4)·(C14H8O2)

Single-crystal X-ray diffraction data determined that (C6I2F4)·(C14H8O2) crystalizes in the centrosymmetric monoclinic space group P21/c. Within the asymmetric unit is one half of both C6I2F4 and C14H8O2 where inversion symmetry generates the remainder of each molecule. The co-crystal is held together by I···O halogen bonds (I···O 2.983(1) Å; C-I···O 169.02(5)°) to generate a one-dimensional chain structure (Figure 1). The donor and acceptor aromatic rings are rotated from co-planar by 72.64° within the chain (Figure 1). In addition to the halogen bond, the molecular components are also engaged in C=O···π interactions, with the closest contact being between an oxygen atom and a pair of fluorinated carbon atoms (O···C 3.080(2) and 3.086(2) Å; C=O···C 112.77(12) and 126.42(12)°) [25,26,27] (Figure 1). The location of this non-covalent interaction correlates with the position of the maximum electrostatic potential on the surface of C6I2F4. In addition, molecules of C14H8O2 are involved in an offset and face-to-face π-π stacking arrangement that generates a homogeneous column at a planar distance of 3.467 Å (Figure 2). Lastly, these halogen-bonded chains interact with their nearest neighbors by C-H···F contacts [33] (C···F 3.275(2) and 3.512(2) Å; C-H···F 141.39(12) and 155.12(12)°) (Figure 3). The combination of these non-covalent interactions results in a three-dimensional extended co-crystal.

3.2. X-ray Crystal Structure of (C6I2Cl4)·(C14H8O2)

Single-crystal diffraction data determined that the molecular components of (C6I2Cl4)·(C14H8O2) crystalize in the centrosymmetric triclinic space group Pī. The asymmetric unit contains a half of a molecule of both C6I2Cl4 and C14H8O2, where inversion symmetry generates each whole molecule. The co-crystal is primarily sustained by I···O halogen bonds (I···O 2.930(2) Å; C-I···O 174.59(7)°), which generate a one-dimensional chain structure (Figure 4). Within the halogen-bonded polymer, the aromatic rings are rotated from co-planarity by 57.65° (Figure 4). As seen in single- and multi-component solids containing C6I2Cl4, the ditopic halogen-bond donor is found to engage in an infinite homogeneous and face-to-face π-π stacking arrangement (Figure 4). Due to this type of stacking pattern, along with the I···O halogen bonds, both the donor and acceptor are found parallel and slightly offset π-stacking with a centroid-to-centroid distance of 4.0473(4) Å, which is equal to the crystallographic a-axis. Lastly, these chains also interact with their nearest neighbors by Type I chlorine–chlorine interactions [34,35] (Cl···Cl 3.4878(9) Å; C-Cl···Cl 146.87(8), and 145.93(8)°; |θ1 − θ2| = 0.94°) (Figure 5). The combination of these non-covalent interactions generates a three-dimensional extended solid.

3.3. Halogen Bond Energies Using Density Functional Theory Calculations

To enumerate the strength of the I···O halogen bond within both (C6I2F4)·(C14H8O2) and (C6I2Cl4)·(C14H8O2), a theoretical investigation using density functional theory (DFT) was performed using the M062X density functional at the aug-cc-pVTZ basis set. The I···O halogen bond binding energies were determined to be −18.2 and −17.9 kJ/mol for (C6I2F4)·(C14H8O2) and (C6I2Cl4)·(C14H8O2), respectively. In comparison, the binding energy for a π-type halogen bond for both donors was also calculated in order to compare the energetic differences. In particular, a co-crystal containing C6I2F4 and anthracene [11] was selected, along with a solid based upon C6I2Cl4 and naphthalene [20]. Again, the DFT calculations used atomic positions from the published diffraction data. The binding energies for π-type halogen bonds between C6I2F4 and anthracene returned values of −16.0 and −15.9 kJ/mol. Similarly, the co-crystal containing C6I2Cl4 and naphthalene had a binding energy of −13.6 kJ/mol for the π-type halogen bond. In both examples, the I···O halogen bond returned a higher value for the binding energy than the π-type halogen bond, which supports the observed bias in these reported co-crystals.

4. Conclusions

In this contribution, we report the preference of I···O over π-type halogen bonds in a pair of co-crystals, namely (C6I2F4)·(C14H8O2) and (C6I2Cl4)·(C14H8O2). This selectivity in the type of halogen bond is largely based upon the difference in electrostatic potential energy between the carbonyl oxygen and the aromatic surface of the acceptor. Each component is ditopic, either donating or accepting two I···O halogen bonds, thereby generating a one-dimensional chain structure in both co-crystals. We note that the observation of C=O···π interactions in the co-crystal (C6I2F4)·(C14H8O2) is consistent with the enhanced electrophilic nature of C6I2F4 relative to the chlorinated analogue C6I2Cl4. Currently, we are investigating the formation of related co-crystals with other polycyclic aromatics that contain different heteroatoms.

Author Contributions

Conceptualization, R.H.G.; methodology, E.B., H.R.K. and R.H.G.; formal analysis, E.B., D.K.U., H.R.K. and R.H.G.; investigation, E.B., D.K.U., R.K.B., H.R.K. and R.H.G.; writing—original draft preparation, R.H.G.; writing—review and editing, E.B., D.K.U., R.K.B., H.R.K. 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

Data are contained within the article.

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. 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]
  2. 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]
  3. Politzer, P.; Murray, J.S.; Clark, T.; Resnati, G. The σ-hole revisited. Phys. Chem. Chem. Phys. 2017, 19, 32166–32178. [Google Scholar] [CrossRef] [PubMed]
  4. 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]
  5. Vainauskas, J.; Borchers, T.H.; Arhangelskis, M.; McCormick McPherson, L.J.; Spilfogel, T.S.; Hamzehpoor, E.; Topić, F.; Coles, S.J.; Perepichka, D.F.; Barrett, C.J.; et al. Halogen bonding with carbon: Directional assembly of non-derivatised aromatic carbon systems into robust supramolecular ladder architectures. Chem. Sci. 2023, 14, 13031–13041. [Google Scholar] [CrossRef] [PubMed]
  6. d’Agostino, S.; Grepioni, F.; Braga, D.; Ventura, B. Tipping the Balance with the Aid of Stoichiometry: Room Temperature Phosphorescence versus Fluorescence in Organic Cocrystals. Cryst. Growth Des. 2015, 15, 2039–2045. [Google Scholar] [CrossRef]
  7. Forni, A.; Metrangolo, P.; Pilati, T.; Resnati, G. Halogen Bond Distance as a Function of Temperature. Cryst. Growth Des. 2004, 4, 291–295. [Google Scholar] [CrossRef]
  8. Walsh, R.B.; Padgett, C.W.; Metrangolo, P.; Resnati, G.; Hanks, T.W.; Pennington, W.T. Crystal Engineering through Halogen Bonding: Complexes of Nitrogen Heterocycles with Organic Iodides. Cryst. Growth Des. 2001, 1, 165–175. [Google Scholar] [CrossRef]
  9. Hutchins, K.M.; Unruh, D.K.; Carpenter, D.D.; Groeneman, R.H. Thermal expansion along one-dimensional chains and two-dimensional sheets within co-crystals based upon halogen or hydrogen bonds. CrystEngComm 2018, 22, 7232–7235. [Google Scholar] [CrossRef]
  10. Vainauskas, J.; Topić, F.; Bushuyer, O.S.; Barrett, C.J.; Friščić, T. Halogen bonding to the azulene π-system: Cocrystal design of pleochroism. Chem. Commun. 2020, 56, 15145–15148. [Google Scholar] [CrossRef]
  11. Azzali, A.; d’Agostino, S.; Capacci, M.; Spinelli, F.; Ventura, B.; Grepioni, F. Assembling photoactive materials from polycyclic aromatic hydrocarbons (PAHs): Room temperature phosphorescence and excimer-emission in cocrystals with 1,4-diiodotetrafluorobenzene. CrystEngComm 2022, 24, 5748–5756. [Google Scholar] [CrossRef]
  12. 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]
  13. Juneja, N.; Shapiro, N.M.; Unruh, D.K.; Bosch, E.; Groeneman, R.H.; Hutchins, K.M. Controlling Thermal Expansion in Supramoelcular Halogen-Bonded Mixed Cocrystals though Synthetic Feed and Dynamic Motion. Angew. Chem. Int. Ed. 2022, 61, e202202708. [Google Scholar] [CrossRef] [PubMed]
  14. 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]
  15. 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]
  16. 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]
  17. 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]
  18. Black, S.N. Geometry and Symmetry of Aromatic Approaches in Halobenzene Crystal Structures: Defining Two Symthons. Cryst. Growth Des. 2021, 21, 6981–6991. [Google Scholar] [CrossRef]
  19. Schmidt, G.M.J. Photodimerization in the Solid State. Pure Appl. Chem. 1971, 27, 647–678. [Google Scholar] [CrossRef]
  20. Bosch, E.; Reinheimer, E.W.; Unruh, D.K.; Groeneman, R.H. Co-crystal sustained by π-type halogen bonding interactions between 1,4-diiodoperchlorobenzene and naphthalene. Acta Crystallogr. 2023, E79, 958–961. [Google Scholar] [CrossRef]
  21. Inscoe, B.; Rathnayake, H.; Mo, Y. Role of Charge Transfer in Halogen Bonding. J. Phys. Chem. A 2021, 125, 2944–2953. [Google Scholar] [CrossRef] [PubMed]
  22. Řezáč, J.; de la Lande, A. On the role of charge transfer in halogen bonding. Phys. Chem. Chem. Phys. 2017, 19, 791–803. [Google Scholar] [CrossRef]
  23. 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]
  24. Spartan ’20, Version 1.0.1; Wavefunction, Inc.: Irvine, CA, USA, 2020.
  25. Korenaga, T.; Tanaka, H.; Ema, T.; Sakai, T. Intermolecular oxygen atom⋯π interaction in the crystal packing of chiral amino alcohol bearing a pentafluorophenyl group. J. Fluor. Chem. 2003, 122, 201–205. [Google Scholar] [CrossRef]
  26. Berger, G.; Soubhye, J.; van der Lee, A.; Velde, C.V.; Wintjens, R.; Dubois, P.; Clément, S.; Meyer, F. Interplay between Halogen Bonding and Lone Pair-π Interactions: A Computational and Crystal Packing Study. ChemPlusChem 2014, 79, 552–558. [Google Scholar] [CrossRef]
  27. Novotný, J.; Bazzi, S.; Marek, R.; Kozelka, J. Lone-pair-π interactions: Analysis of the physical origin and biological implications. Phys. Chem. Chem. Phys. 2016, 18, 19472–19481. [Google Scholar] [CrossRef] [PubMed]
  28. 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. (Eds.) Gaussian 16, Revision, C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  29. 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]
  30. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. 2015, A71, 3–8. [Google Scholar] [CrossRef]
  31. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. 2015, C71, 3–8. [Google Scholar]
  32. 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]
  33. Hobza, P.; Havlas, Z. Blue-Shifting Hydrogen Bonds. Chem. Rev. 2000, 100, 4253–4264. [Google Scholar] [CrossRef] [PubMed]
  34. Scheiner, S. Characterization of Type I and Type II Interactions between Halogen Atoms. Cryst. Growth Des. 2022, 22, 2692–2702. [Google Scholar] [CrossRef]
  35. 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] [PubMed]
Scheme 1. Rendering of the molecular components within the pair of halogen-bonded co-crystals, namely (C6I2F4)·(C14H8O2) and (C6I2Cl4)·(C14H8O2).
Scheme 1. Rendering of the molecular components within the pair of halogen-bonded co-crystals, namely (C6I2F4)·(C14H8O2) and (C6I2Cl4)·(C14H8O2).
Crystals 14 00325 sch001
Scheme 2. The molecular electrostatic potential energy surface for C14H8O2. The unit for the scale is in kJ/mol.
Scheme 2. The molecular electrostatic potential energy surface for C14H8O2. The unit for the scale is in kJ/mol.
Crystals 14 00325 sch002
Figure 1. X-ray structure of (C6I2F4)·(C14H8O2) illustrating the extended structure sustained by I···O halogen bonds and O···π interactions. The I···O halogen bonds and O···π interactions are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, fluorine is gold, iodine is purple, and oxygen is red.
Figure 1. X-ray structure of (C6I2F4)·(C14H8O2) illustrating the extended structure sustained by I···O halogen bonds and O···π interactions. The I···O halogen bonds and O···π interactions are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, fluorine is gold, iodine is purple, and oxygen is red.
Crystals 14 00325 g001
Figure 2. X-ray structure of (C6I2F4)·(C14H8O2) illustrating both the I···O halogen bonds, along with the offset and face-to-face π-π stacking of the acceptor. Halogen bonds are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, fluorine is gold, iodine is purple, and oxygen is red.
Figure 2. X-ray structure of (C6I2F4)·(C14H8O2) illustrating both the I···O halogen bonds, along with the offset and face-to-face π-π stacking of the acceptor. Halogen bonds are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, fluorine is gold, iodine is purple, and oxygen is red.
Crystals 14 00325 g002
Figure 3. X-ray structure of (C6I2F4)·(C14H8O2) illustrating both the I···O halogen bonds, along with the C-H···F contacts. The I···O halogen bonds and C-H···F contacts are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, fluorine is gold, iodine is purple, and oxygen is red.
Figure 3. X-ray structure of (C6I2F4)·(C14H8O2) illustrating both the I···O halogen bonds, along with the C-H···F contacts. The I···O halogen bonds and C-H···F contacts are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, fluorine is gold, iodine is purple, and oxygen is red.
Crystals 14 00325 g003
Figure 4. X-ray structure of (C6I2Cl4)·(C14H8O2) illustrating both the I···O halogen bonds along with the homogeneous and face-to-face π-π stacking of both the donor and acceptor molecules. The I···O halogen bonds are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, chlorine is green, iodine is purple, and oxygen is red.
Figure 4. X-ray structure of (C6I2Cl4)·(C14H8O2) illustrating both the I···O halogen bonds along with the homogeneous and face-to-face π-π stacking of both the donor and acceptor molecules. The I···O halogen bonds are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, chlorine is green, iodine is purple, and oxygen is red.
Crystals 14 00325 g004
Figure 5. X-ray structure of (C6I2Cl4)·(C14H8O2) illustrating both the I···O halogen bonds along with Type I chlorine–chlorine interactions. The I···O halogen bonds and Type I chlorine–chlorine interactions are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, chlorine is green, iodine is purple, and oxygen is red.
Figure 5. X-ray structure of (C6I2Cl4)·(C14H8O2) illustrating both the I···O halogen bonds along with Type I chlorine–chlorine interactions. The I···O halogen bonds and Type I chlorine–chlorine interactions are shown with yellow dashes. Color scheme of the atoms: carbon is grey, hydrogen is white, chlorine is green, iodine is purple, and oxygen is red.
Crystals 14 00325 g005
Table 1. Crystallographic and refinement parameters for the halogen-bonded co-crystals (C6I2F4)·(C14H10O2) and (C6I2Cl4)·(C14H10O2).
Table 1. Crystallographic and refinement parameters for the halogen-bonded co-crystals (C6I2F4)·(C14H10O2) and (C6I2Cl4)·(C14H10O2).
Co-crystal(C6I2F4)·(C14H8O2)(C6I2Cl4)·(C14H8O2)
FormulaC20H8F4I2O2C20H8Cl4I2O2
Formula mass (g·mol−1) 610.06675.86
Crystal systemmonoclinictriclinic
Space groupP21/cPī
a (Å)11.3817(5)4.0473(4)
b (Å)5.9765(3)8.0764(10)
c (Å)13.5850(6)15.2448(18)
α (°)9084.625(4)
β (°)101.652(1)86.055(4)
γ (°)9086.488(4)
Z21
V3)905.05(7)494.20(10)
ρcalcd (g·cm−3)2.2392.271
T (K)100100
μ (mm−1)3.5283.738
F(000)572.0318.0
Radiation sourceMo KαMo Kα
Reflections collected22,41429,776
Independent reflections22402455
Data/restraints/parameters2240/0/1272455/0/128
Rint0.03340.0575
R1 (I ≥ 2σ(I))0.01540.0190
wR (F2) (I ≥ 2σ(I))0.03740.0399
R1 (all data)0.01700.0208
wR (F2) (all data)0.03810.0407
Goodness-of-net on F21.1371.107
CCDC deposition number2,339,4952,332,122
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bosch, E.; Unruh, D.K.; Brooks, R.K.; Krueger, H.R.; Groeneman, R.H. Preference in the Type of Halogen Bonding Interactions within Co-Crystals of Anthraquinone with a Pair of Isosteric Perhalobenzenes. Crystals 2024, 14, 325. https://doi.org/10.3390/cryst14040325

AMA Style

Bosch E, Unruh DK, Brooks RK, Krueger HR, Groeneman RH. Preference in the Type of Halogen Bonding Interactions within Co-Crystals of Anthraquinone with a Pair of Isosteric Perhalobenzenes. Crystals. 2024; 14(4):325. https://doi.org/10.3390/cryst14040325

Chicago/Turabian Style

Bosch, Eric, Daniel K. Unruh, Richard K. Brooks, Herman R. Krueger, and Ryan H. Groeneman. 2024. "Preference in the Type of Halogen Bonding Interactions within Co-Crystals of Anthraquinone with a Pair of Isosteric Perhalobenzenes" Crystals 14, no. 4: 325. https://doi.org/10.3390/cryst14040325

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

Article Metrics

Back to TopTop