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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystallization, structural study and analysis of inter­molecular inter­actions of a 2-amino­benzoxazole–fumaric acid mol­ecular salt

crossmark logo

aNational University of Uzbekistan named after Mirzo Ulugbek, 4 University St, Tashkent, 100174, Uzbekistan, bInstitute of General and Inorganic Chemistry, Academy of Sciences of Uzbekistan, 100170, M. Ulugbek Str 77a, Tashkent, Uzbekistan, and cInstitute of Bioorganic Chemistry, Academy of Sciences of Uzbekistan, M. Ulugbek, Str, 83, Tashkent, 700125, Uzbekistan
*Correspondence e-mail: torambetov_b@mail.ru

Edited by J. Reibenspies, Texas A & M University, USA (Received 24 October 2022; accepted 21 November 2022; online 30 November 2022)

The new organic salt, 2-amino­benzoxazol-3-ium 3-carb­oxy­prop-2-enoate, C7H7N2O+·C4H3O4, of the two bioactive compounds 2-amino­benzoxazole and fumaric acid, crystallizes in the ortho­rhom­bic space group Pbca using classical evaporation of their solution in water. The usual topological analysis revealed four classical (N—H⋯O and O—H⋯O) and two non-classical (C—H⋯O) hydrogen bonds in the structure. Stacking was found as well for a pair of 2-amino­benzoxazolium cations. A Hirshfeld surface analysis including the two-dimensional fingerprint plots was performed to define the residual ππ inter­actions and to qu­antify the influences of different types of inter­actions by means of topological analysis. Analysis of the pairwise inter­action energies was used to prove the formation of the corrugated paired layers of cation–anion dimers parallel to the plane (001) as a basic structural motif in the topological, as well as in the energetic structure of the crystal. It showed that the layers are connected by the hydrogen bonds inside and by stacking and ππ inter­actions and general dispersion between them.

1. Chemical context

Benzoxazole derivatives are important heterocyclic compounds that exhibit a broad range of biological activities, including anti­bacterial (Paramashivappa et al., 2003[Paramashivappa, R., Phani Kumar, P., Subba Rao, P. V. & Srinivasa Rao, A. (2003). Bioorg. Med. Chem. Lett. 13, 657-660.]), anti­microbial (Erol et al., 2022[Erol, M., Celik, I., Uzunhisarcikli, E. & Kuyucuklu, G. (2022). Polycycl. Aromat. Compd. 42, 1679-1696.]), anti­tumor (Imaizumi et al., 2020[Imaizumi, T., Otsubo, S., Komai, M., Takada, H., Maemoto, M., Kobayashi, A. & Otsubo, N. (2020). Bioorg. Med. Chem. 28, 115622.]), anti-inflammatory (Parlapalli et al., 2017[Parlapalli, A. & Manda, S. (2017). J. Chem. Pharm. Res. 9(9), 57-62.]), analgesic (Ali et al., 2022[Ali, S., Omprakash, P., Tengli, A. K., Mathew, B., Basavaraj, M. V., Parkali, P., Chandan, R. S. & Kumar, A. S. (2022). Polycyclic Aromat. Compd. 30, 1-34.]; Sattar et al., 2020[Sattar, R., Mukhtar, R., Atif, M., Hasnain, M. & Irfan, A. (2020). J. Heterocycl. Chem. 57, 2079-2107.]), anti­tubercular (Šlachtová et al., 2018[Šlachtová, V. & Brulíková, L. (2018). ChemistrySelect 3, 4653-4662.]), herbicidal (Sangi et al., 2019[Sangi, D. P., Meira, Y. G., Moreira, N. M., Lopes, T. A., Leite, M. P., Pereira-Flores, M. E. & Alvarenga, E. S. (2019). Pest Manag. Sci. 75, 262-269.]) and fungicidal (Fan et al., 2022[Fan, L., Luo, Z., Yang, C., Guo, B., Miao, J., Chen, Y., Tang, L. & Li, Y. (2022). Mol. Divers. 26, 981-992.]). 2-Amino­benzoxazoles have been found to act as ligands for the inter­nal ribosome entry site (IRES) RNA of the hepatitis C virus (HCV) (Rynearson et al., 2014[Rynearson, K. D., Charrette, B., Gabriel, C., Moreno, J., Boerneke, M. A., Dibrov, S. M. & Hermann, T. (2014). Bioorg. Med. Chem. Lett. 24, 3521-3525.]).

The formation of a salt or co-crystal presents a useful tool for advantageously modifying the physicochemical properties of an active pharmaceutical ingredient (e.g. bioavailability and processing characteristics) without altering its basic chemical structure and pharmacological properties (Guillory et al., 2003[Guillory, J. K. (2003). J. Med. Chem. 46, 1277-1277.]; Callear et al., 2009[Callear, S. K., Hursthouse, M. B. & Threlfall, T. L. (2009). CrystEngComm, 11, 1609-1614.]). The pharmaceutical co-crystal can be explained as a multi-component crystal in which at least one of the mol­ecular components is an API, along with the other component called the co-crystal former (Soares et al., 2014[Soares, F. L. F. & Carneiro, R. L. (2014). J. Pharm. Biomed. Anal. 89, 166-175.]). The co-crystal former is believed to help the active drug to disintegrate into small particles and to be transported to the blood stream where the drug is intended to play its role, and still protect the product's stability so that it has the greatest benefits and effectiveness (Blagden et al., 2007[Blagden, N., de Matas, M., Gavan, P. T. & York, P. (2007). Adv. Drug Deliv. Rev. 59, 617-630.]). It has been reported that the co-crystallization process allowed the binding of two or more crystal components in a single crystalline lattice via hydrogen bonding and van der Walls inter­molecular inter­actions without breaking the bonds or making new covalent bonds (Sonawane et al., 2013[Sonawane, A. R., Rawat, S. S. & Janolkar, N. N. (2013). Asian J. Biomed. Pharm. Sci. 27, 1-8.]; Sheikh et al., 2009[Sheikh, A. Y., Rahim, S. A., Hammond, R. B. & Roberts, K. J. (2009). CrystEngComm, 11, 501-509.]).

The organic acids containing the donor and acceptor groups capable of classical hydrogen bonding are used for multi-component assembly, so they are frequently chosen as building blocks in supra­molecular crystal engineering (Xu et al., 2019[Xu, W., Lu, Y., Xia, Y., Liu, B., Jin, S., Zhong, B., Wang, D. & Guo, M. (2019). J. Mol. Struct. 1189, 81-93.]). Fumaric acid is the E-isomer of butenedioic acid and is one of the organic compounds found widely in nature. Fumaric acid is also a key inter­mediate in the biosynthesis of organic acids, and forms inter­esting one-, two- and three-dimensional supra­molecular architectures as adducts with various amines (Franklin et al., 2009[Franklin, S. & Balasubramanian, T. (2009). Acta Cryst. C65, o58-o61.]; Batchelor et al., 2000[Batchelor, E., Klinowski, J. & Jones, W. (2000). J. Mater. Chem. 10, 839-848.]).

[Scheme 1]

Herein, we report on the crystal structure and Hirshfeld surface analysis of a new co-crystal of the 2-amino­benzoxazole–fumaric acid mol­ecular salt (2ABHF).

2. Structural commentary

The co-crystal salt of 2-amino­benzoxazole and fumaric acid crystallizes in the primitive centrosymmetric ortho­rhom­bic space group Pbca. As seen in Fig. 1[link], the asymmetric unit of 2ABHF consists of a 2-amino­benzoxazolium cation and a semifumarate anion.

[Figure 1]
Figure 1
Mol­ecular structure of the title compound. Displacement ellipsoids are shown at the 50% probability level.

Atom N1 in the 2-amino­benzoxazolium cation is proton­ated. The 2-amino­benzoxazole ring of 2ABHF is essentially planar, with a maximum deviations from the general planarity of 0.019 (1) Å for the atom C1. The amino group in the 2-amino­benzoxazolium cation is planar, the sum of bond angles at the N atom being 359.99°.

The semifumarate anion is slightly twisted, showing a deviation from planarity of 0.175 (1) Å for atom O2 and a dihedral angle between the carboxyl­ate (O2/O3/C8) and carb­oxy­lic acid (O4/O5/C10) mean planes of 15.6 (2)°.

3. Supra­molecular features

Regarding the van der Waals radii proposed in Bondi (1964[Bondi, A. (1964). J. Phys. Chem. 68, 441-451.]) for all the atoms except hydrogen (Rowland & Taylor, 1996[Rowland, R. S. & Taylor, R. (1996). J. Phys. Chem. 100, 7384-7391.]), four classical hydrogen bonds of two types are found in the crystal structure (Table 1[link]). The first type is N—H⋯O. It is the most obvious inter­action type, forming a dimer comprising an amino­benzoxazolium cation and a semifumarate anion in their original positions (x, y, z) with the bonds N1—H1⋯O3 and N2—H2B⋯O2. The other N—H⋯O hydrogen bond, N2—H2A⋯O3, connects the amino­benzoxazolium cation with its neighbouring semifumarate anion (the symmetry operation is [{1\over 2}] − x, −[{1\over 2}] + y, z). As seen in Fig. 2[link], this, together with a hydrogen bond of the second type (O—H⋯O), leads to the formation of layers of dimers along the (001) plane. This O4—H4A⋯O2 hydrogen bond connects two semifumarate anions (original and its symmetry equivalent 1 − x, [{1\over 2}] + y, [{1\over 2}] − z). The components within a layer are connected by hydrogen bonds, while stacking/ππ inter­actions or general dispersion connect the layers with each other (Fig. 3[link]). Seemingly, the stacking is formed between the original amino­benzoxazolium cation and its symmetry equivalent at −x, 1 − y, 1 − z, because the distance between the mean planes of the benzoxazolium fragments is 3.469 (2) Å (Fig. 4[link]). In addition, there are two weak non-classical C—H⋯O hydrogen bonds in the structure. One of them stabilizes the layers by binding the dimers along the [010] direction together with the classical hydrogen bond N2—H2A⋯O3. The other one is its counterpart stabilizing the inter­layer inter­actions.

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
C6—H6⋯O1i 0.93 2.58 3.4929 (17) 167
C3—H3⋯O5ii 0.93 2.51 3.2563 (17) 137
N1—H1⋯O3 0.99 (2) 1.69 (2) 2.6720 (14) 175.8 (17)
O4—H4A⋯O2iii 1.00 (2) 1.60 (2) 2.5913 (14) 174 (2)
N2—H2A⋯O3iv 0.90 (2) 1.94 (2) 2.8355 (16) 176.4 (18)
N2—H2B⋯O2 0.94 (2) 1.89 (2) 2.7873 (16) 157.8 (18)
Symmetry codes: (i) [-x+{\script{1\over 2}}, y+{\script{1\over 2}}, z]; (ii) [-x+{\script{1\over 2}}, -y+1, z+{\script{1\over 2}}]; (iii) [-x+1, y+{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iv) [-x+{\script{1\over 2}}, y-{\script{1\over 2}}, z].
[Figure 2]
Figure 2
Crystal packing of the corrugated layers parallel to the (001) plane with the hydrogen bonds in cyan. Projection in the [010] direction.
[Figure 3]
Figure 3
Crystal packing of the mol­ecules from the same (in blue) and adjacent (in red) layers connected to the initial dimer (in green) with specific inter­actions. The colour codes of the atoms participating in inter­actions and corresponding symmetry operations of neighbours are similar.
[Figure 4]
Figure 4
The mol­ecular stacking (green projections on the 2-amino­benzoxazolium mean plane) and ππ inter­actions (blue projections) formed between adjacent dimers.

4. Hirshfeld surface analysis

To further investigate the inter­molecular inter­actions present in the title compound, a Hirshfeld surface analysis (Spackman & Byrom, 1997[Spackman, M. A. & Byrom, P. G. (1997). Chem. Phys. Lett. 267, 215-220.]) was performed, and the two-dimensional fingerprint plots were generated with CrystalExplorer17 (Spackman et al., 2021[Spackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006-1011.]). The Hirshfeld surface with the normalized contact distance (dnorm) mapped over it was built using standard `high' resolution. It is shown in Fig. 5[link] in two variants: (a) with the default scaling (−0.7348, 1.1456 Å for the amino­benzoxazolium cation and −0.7771, 1.1324 Å for semifumarate anion) and (b) showing the minimal shortening in van der Waals radii (the lower limit of dnorm renormalized to −0.0001 Å). However, some inter­actions are obviously overestimated, such as H⋯H (small red dots on the renormalized Hirshfeld surfaces). This is caused by the use of van der Waals radii by Bondi with a markedly overestimated radius for hydrogen atoms (Rowland & Taylor, 1996[Rowland, R. S. & Taylor, R. (1996). J. Phys. Chem. 100, 7384-7391.]). Indeed, the biggest shortening occurs for the classical hydrogen bonds, but it is seen that all the donor and acceptor sites are involved in non-covalent inter­actions obeying Etters' rules (Etter, 1990[Etter, M. C. (1990). Acc. Chem. Res. 23, 120-126.]; Bernstein et al., 1995[Bernstein, J., Davis, R. E., Shimoni, L. & Chang, N.-L. (1995). Angew. Chem. Int. Ed. Engl. 34, 1555-1573.]).

[Figure 5]
Figure 5
Distribution of dnorm on the Hirshfeld surfaces of the cation and anion with the default normalization (a) and renormalized (b).

The two-dimensional (2D) fingerprint plots (Spackman & McKinnon, 2002[Spackman, M. A. & McKinnon, J. J. (2002). CrystEngComm, 4, 378-392.]) are shown in Fig. 6[link]. The most significant inter­actions whose contribution into the Hirshfeld surface area exceeds 5.0% at least for one of the ions in the structure are O⋯H/H⋯O (37.3 and 51.5% for the amino­benzoxazolium and semifumarate moieties, respectively), H⋯H (28.9 and 24.4%), C⋯H/H⋯C (13.7 and 9.1%), O⋯C/C⋯O (4.6 and 10.0%) and finally C⋯C (5.5 and 2.9%). Considering such a shortening of the inter­molecular distances, hydrogen bonding can be considered the most important type of inter­action in the structure. However, ππ inter­actions are also present to a large extent, which is easy to see from the localization of O⋯C/C⋯O and C⋯C contacts around the carbon atoms of the aromatic system of the cation and carboxyl­ate moiety of the anion (Fig. 7[link]). The inter­actions between the original amino­benzoxazolium cation with its symmetry equivalent at 1 − x, 1 − y, 1 − z as well as the same symmetry equivalent of the semifumarate anion are of particular inter­est. They are difficult to recognize from the structural data without additional processing using the Hirshfeld surface analysis but can play an important role in the inter­layer binding. It should be mentioned that the sharp peaks in the fingerprint plots reaching values less than the van der Waals radii of the corresponding atom types whose appearance is usually associated with the appearance of inter­molecular inter­actions are seen only for some of all the aforementioned inter­actions. They are O⋯H/H⋯O, H⋯H and C⋯C for both moieties and O⋯C/C⋯O for the semifumarate anion.

[Figure 6]
Figure 6
Contributions of the contacts O⋯H/H⋯O (a), H⋯H (b), C⋯H/H⋯C (c), O⋯C/C⋯O (d) and C⋯C (e) to the two-dimensional fingerprint plots built using the Hirshfeld surfaces of the cation (at the top) and anion (at the bottom).
[Figure 7]
Figure 7
The localization of the short contacts C⋯C (a, c) and O⋯C/C⋯O (b, d) onto the Hirsheld surfaces between the initial hydrogen-bonded dimer and its symmetry equivalents −x, 1 − y, 1 − z (a, b) and 1 − x, 1 − y, 1 − z (c, d).

5. Analysis of the pairwise inter­action energies

The inter­actions and the structural motifs in the crystal were assumed from the previous topological analysis and the next step is to confirm the supposed model using the approach of pairwise inter­actions in crystals (Konovalova et al., 2010[Konovalova, I. S., Shishkina, S. V., Paponov, B. V. & Shishkin, O. V. (2010). CrystEngComm, 12, 909-916.]; Shishkin et al., 2012[Shishkin, O. V., Dyakonenko, V. V. & Maleev, A. V. (2012). CrystEngComm, 14, 1795-1804.]). This approach allows the energetic structure of a crystal to be defined. A two-step procedure was used for the current structure. In the first step, the individual ions were considered as the building units and the hydrogen-bonded dimer of the initial ions was found to be the most strongly bound fragment of the crystal structure. It was therefore taken as a building unit for the second step, and all the following calculations were repeated from scratch. The approach was used in the same way as proposed in Shishkin et al. (2012[Shishkin, O. V., Dyakonenko, V. V. & Maleev, A. V. (2012). CrystEngComm, 14, 1795-1804.]), so just the general options concerning the calculations are shown below. The functional used was B97 (Becke, 1997[Becke, A. D. (1997). J. Chem. Phys. 107, 8554-8560.]; Schmider & Becke, 1998[Schmider, H. L. & Becke, A. D. (1998). J. Chem. Phys. 108, 9624-9631.]) with the parameterized three-body (D3) dispersion correction (Grimme et al., 2010[Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. (2010). J. Chem. Phys. 132, 154104.]) and Becke–Johnson dumping (Grimme et al., 2011[Grimme, S., Ehrlich, S. & Goerigk, L. (2011). J. Comput. Chem. 32, 1456-1465.]). The double-zeta basis set augmented with diffuse functions (Dunning, 1989[Dunning, T. H. Jr (1989). J. Chem. Phys. 90, 1007-1023.]; Kendall & Dunning, 1992[Kendall, R. A., Dunning, T. H. Jr & Harrison, R. J. (1992). J. Chem. Phys. 96, 6796-6806.]; Woon & Dunning, 1993[Woon, D. E. & Dunning, T. H. Jr (1993). J. Chem. Phys. 98, 1358-1371.]; Peterson et al., 1994[Peterson, K. A., Woon, D. E. & Dunning, T. H. Jr (1994). J. Chem. Phys. 100, 7410-7415.]; Wilson et al., 1996[Wilson, A. K., van Mourik, T. & Dunning, T. H. Jr (1996). J. Mol. Struct. Theochem, 388, 339-349.]; Davidson, 1996[Davidson, E. R. (1996). Chem. Phys. Lett. 260, 514-518.]) was used since the structure contains atoms up to the second period. The conductor-like polarizable continuum model (CPCM) was applied to all the pairs of building units to treat the charged system correctly (Barone & Cossi, 1998[Barone, V. & Cossi, M. (1998). J. Phys. Chem. A, 102, 1995-2001.]). In addition, the Boys–Bernardi counterpoise scheme (Boys & Bernardi, 1970[Boys, S. F. & Bernardi, F. (1970). Mol. Phys. 19, 553-566.]) was also used for the basis set superposition error (BSSE) in the software ORCA 5.0.2 (Neese et al., 2020[Neese, F., Wennmohs, F., Becker, U. & Riplinger, C. (2020). J. Chem. Phys. 152, 224108.]). The calculations were finalized by building the vector energy diagrams in a standard way (Shishkin et al., 2012[Shishkin, O. V., Dyakonenko, V. V. & Maleev, A. V. (2012). CrystEngComm, 14, 1795-1804.]).

The inter­action in the hydrogen-bonded dimer is about twice as strong as any other inter­action between the individual ions (−24.0 kcal mol−1). This fact allowed us to consider the energetic structure of the crystal using the hydrogen-bonded dimer of the amino­benzoxazolium cation and semifumarate anion as a building unit in all of the following calculations. However, the dimer shows an almost isotropic distribution of inter­action energy and just taking into account the types of inter­actions found during the previous steps allowed us to classify this structure as layered with the corrugated paired layers as a structural motif. The sum of inter­action energies between the central dimer and its sixteen neighbours from the first coordination shell is −82.0 kcal mol−1. There are six inter­actions with energies close to each other and about 3–3.5 times higher than any other in the structure (Table 2[link]). All of the strong in-layer inter­actions (Fig. 8[link] in red) correspond to the classical hydrogen bonds or a classical hydrogen bond reinforced by the non-classical hydrogen bond C6—H6⋯O1. Two residual high energies belong to the aforementioned inter­layer stacking and ππ inter­actions (Fig. 8[link] in blue). The final non-classical hydrogen bond, C3—H3⋯O5, introduces a negligible contribution to the inter­action of the layers (−1.4 kcal mol−1). The total inter­action energies are −51.7 within the layers (001) and −30.3 kcal mol−1 between the layers.

Table 2
Symmetry codes, binding types and inter­action energies (kcal mol−1) of the building units (dimers) with neighbours

Pair of building units Symmetry operation of neighbouring building unit Eint Inter­action
1 [{1\over 2}] + x, [{1\over 2}] − y, 1 − z −2.1 Non-specific
2 [{1\over 2}] + x, [{3\over 2}] − y, 1 − z −1.6 Non-specific
3 x, 1 − y, 1 − z −10.6 Stacking
4 [{1\over 2}] + x, y, [{1\over 2}] − z −3.0 Non-specific
5 [{1\over 2}] + x, [{1\over 2}] − y, 1 − z −2.1 Non-specific
6 [{1\over 2}] − x, 1 − y, −[{1\over 2}] + z −1.4 C3—H3⋯O5
7 x, [{3\over 2}] − y, −[{1\over 2}] + z −0.9 Non-specific
8 [{1\over 2}] − x, 1 − y, 1/ 2 + z −1.4 C3—H3⋯O5
9 1 − x, −[{1\over 2}] + y, [{1\over 2}] − z −11.4 O4—H4A⋯O2
10 [{1\over 2}] + x, [{3\over 2}] − y, 1 − z −1.6 Non-specific
11 1 − x, 1 − y, 1 − z −10.0 Stacking
12 [{1\over 2}] + x, y, [{1\over 2}] − z −3.0 Non-specific
13 x, [{3\over 2}] − y, [{1\over 2}] + z −0.9 Non-specific
14 [{1\over 2}] − x, −[{1\over 2}] + y, z −10.3 N2—H2A⋯O3, C6—H6⋯O1
15 1 − x, [{1\over 2}] + y, [{1\over 2}] − z −11.4 O4—H4A⋯O2
16 [{1\over 2}] − x, [{1\over 2}] + y, z −10.3 N2—H2A⋯O3, C6—H6⋯O1
[Figure 8]
Figure 8
Crystal packing of the mol­ecules with hydrogen bonds in cyan (a) and energy vector diagrams of mol­ecules (b) with a single layer in red, as well as the energy vector diagrams of hydrogen-bonded dimers (c) with the hydrogen-bonding (in layer) inter­actions in red and stacking/ππ inter­actions (between layers) in blue.

6. Database survey

A search of the Cambridge Structural Database (CSD, Version 5.43, update of November 2021; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) for the 2-amino­benzoxazole unit resulted in twelve hits. They include the following analogues: a 5-chloro derivative and its monohydrate and co-crystals with some acids (PAHBUW, Lynch, 2004[Lynch, D. E. (2004). Acta Cryst. E60, o1715-o1716.]; XUVPAG, Lynch, 2009[Lynch, D. E. (2009). CSD Communication (refcode XUVPAG, CCDC 716694). CCDC, Cambridge, England.]; XEGLAW, Lynch et al., 2000a[Lynch, D. E., Daly, D. & Parsons, S. (2000a). Acta Cryst. C56, 1478-1479.]; RADGAF, Lynch et al., 2003[Lynch, D. E., Barfield, J., Frost, J., Antrobus, R. & Simmons, J. (2003). Cryst. Eng. 6, 109-122.]; LOMTIQ, LOMTEM, LOMTAI, Lynch et al., 2000b[Lynch, D. E., Singh, M. & Parsons, S. (2000b). Cryst. Eng. 3, 71-79.]; FAFNIL, FAFNEH, Kruszynski et al., 2010[Kruszynski, R. & Trzesowska-Kruszynska, A. (2010). Acta Cryst. C66, o449-o454.]), and containing derivatives of the 2-amino­benzoxazole unit (EBAMAY, Coleman et al., 2014[Coleman, N., Brown, B. M., Oliván-Viguera, A., Singh, V., Olmstead, M. M., Valero, M. S., Köhler, R. & Wulff, H. (2014). Mol. Pharmacol. 86, 342-357.]; KAVFOC, Colegate et al., 1989[Colegate, S., Dorling, P., Huxtable, C., Shaw, T., Skelton, B., Vogel, P. & White, A. (1989). Aust. J. Chem. 42, 1249-1255.]; VAKMUT, Silva et al., 2021[Silva, L. A., Hottes, E., da Silva, A. M., Freire, L. M. S., Guedes, G. P. & Ferreira, A. B. B. (2021). J. Braz. Chem. Soc. 32, 1009-1016.]). The survey shows that in the structures FAFNIL, FAFNEH, LOMTIQ and XEGLAW, the nitro­gen atom of the oxazole ring is protonated by a hydrogen from the acid. However, the free 2-amino­benzoxazole and its co-crystals with some adducts are not listed in the database. There are numerous submitted structures for fumaric and maleic acids as an adduct for co-crystals of several compounds. However, no co-crystal complexes containing both 2-amino­benzoxazole derivatives and fumaric acid have been documented in the CSD.

7. Synthesis and crystallization

A 1:1 stoichiometric ratio of 2-amino­benzaxazole (0.134 g, 1.0 mmol) and fumaric acid (0.116 g, 1.0 mmol) was dissolved and mixed well in distilled water (3 ml). The mixture was held at 333 K for 10 minutes under stirring. The solution was allowed to stand at room temperature in a beaker with small holes in the cover for evaporation. After about 3 weeks, rectangular single crystals of the C7H7N2O·C4H3O4 co-crystal appeared.

8. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. The hydrogen atoms were refined isotropically with a mixed model. Those involved in classical hydrogen bonds were found in difference-Fourier maps and are free from any constraints or restraints. The other hydrogen atoms were positioned geometrically (C—H = 0.93 Å) and refined using a riding model with Uiso(H) = 1.2Ueq(C).

Table 3
Experimental details

Crystal data
Chemical formula C7H7N2O+·C4H3O4
Mr 250.21
Crystal system, space group Orthorhombic, Pbca
Temperature (K) 293
a, b, c (Å) 7.0694 (1), 12.9543 (2), 24.5079 (4)
V3) 2244.41 (6)
Z 8
Radiation type Cu Kα
μ (mm−1) 1.02
Crystal size (mm) 0.16 × 0.14 × 0.12
 
Data collection
Diffractometer XtaLAB Synergy, Single source at home/near, HyPix3000
Absorption correction Multi-scan CrysAlis PRO; Rigaku OD, 2021[Rigaku OD (2021). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.])
Tmin, Tmax 0.761, 1.000
No. of measured, independent and observed [I > 2σ(I)] reflections 7276, 2129, 1871
Rint 0.023
(sin θ/λ)max−1) 0.609
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.035, 0.096, 1.05
No. of reflections 2129
No. of parameters 180
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.20, −0.16
Computer programs: CrysAlis PRO (Rigaku OD, 2021[Rigaku OD (2021). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.]), SHELXT2018/2 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2016/6 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]) and OLEX2 (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]).

Supporting information


Computing details top

Data collection: CrysAlis PRO (Rigaku OD, 2021); cell refinement: CrysAlis PRO (Rigaku OD, 2021); data reduction: CrysAlis PRO (Rigaku OD, 2021); program(s) used to solve structure: SHELXT2018/2 (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2016/6 (Sheldrick, 2015b); molecular graphics: Olex2 (Dolomanov et al., 2009); software used to prepare material for publication: Olex2 (Dolomanov et al., 2009).

2-Aminobenzoxazol-3-ium 3-carboxyprop-2-enoate top
Crystal data top
C7H7N2O+·C4H3O4Dx = 1.481 Mg m3
Mr = 250.21Cu Kα radiation, λ = 1.54184 Å
Orthorhombic, PbcaCell parameters from 4322 reflections
a = 7.0694 (1) Åθ = 3.6–70.5°
b = 12.9543 (2) ŵ = 1.02 mm1
c = 24.5079 (4) ÅT = 293 K
V = 2244.41 (6) Å3Needle, clear light colourless
Z = 80.16 × 0.14 × 0.12 mm
F(000) = 1040
Data collection top
XtaLAB Synergy, Single source at home/near, HyPix3000
diffractometer
2129 independent reflections
Radiation source: micro-focus sealed X-ray tube, PhotonJet (Cu) X-ray Source1871 reflections with I > 2σ(I)
Detector resolution: 10.0000 pixels mm-1Rint = 0.023
ω scansθmax = 70.0°, θmin = 3.6°
Absorption correction: multi-scan
CrysAlisPro; Rigaku OD, 2021)
h = 87
Tmin = 0.761, Tmax = 1.000k = 915
7276 measured reflectionsl = 2925
Refinement top
Refinement on F2Hydrogen site location: mixed
Least-squares matrix: fullH atoms treated by a mixture of independent and constrained refinement
R[F2 > 2σ(F2)] = 0.035 w = 1/[σ2(Fo2) + (0.0485P)2 + 0.5023P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.096(Δ/σ)max < 0.001
S = 1.05Δρmax = 0.20 e Å3
2129 reflectionsΔρmin = 0.16 e Å3
180 parametersExtinction correction: SHELXL2016/6 (Sheldrick 2015b), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
0 restraintsExtinction coefficient: 0.00154 (19)
Primary atom site location: structure-invariant direct methods
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
O10.17549 (16)0.32002 (7)0.49841 (4)0.0404 (3)
O30.35863 (16)0.61279 (7)0.38952 (4)0.0403 (3)
O20.33909 (18)0.47799 (7)0.33433 (4)0.0483 (3)
O40.59585 (19)0.87984 (7)0.25587 (4)0.0499 (3)
O50.49429 (19)0.75736 (8)0.19951 (4)0.0533 (3)
N10.26884 (18)0.46854 (8)0.46302 (5)0.0350 (3)
N20.2529 (2)0.32025 (9)0.40757 (5)0.0473 (3)
C70.23633 (19)0.48645 (10)0.51866 (5)0.0334 (3)
C80.3664 (2)0.57206 (9)0.34298 (5)0.0345 (3)
C10.2335 (2)0.36996 (10)0.45327 (5)0.0352 (3)
C20.1779 (2)0.39363 (10)0.54042 (6)0.0358 (3)
C110.5249 (2)0.78833 (10)0.24496 (5)0.0365 (3)
C100.4884 (2)0.72758 (10)0.29520 (5)0.0395 (3)
H100.5234900.7553610.3286890.047*
C90.4091 (2)0.63688 (10)0.29419 (5)0.0398 (3)
H90.3762870.6102880.2602340.048*
C60.2553 (2)0.57237 (11)0.55113 (6)0.0427 (4)
H60.2945490.6355060.5370430.051*
C30.1326 (2)0.37945 (12)0.59423 (6)0.0465 (4)
H30.0920640.3162190.6078750.056*
C50.2129 (2)0.55992 (13)0.60616 (7)0.0509 (4)
H50.2259090.6158720.6296340.061*
C40.1516 (3)0.46617 (13)0.62683 (6)0.0524 (4)
H40.1222960.4613170.6637380.063*
H10.305 (3)0.5194 (15)0.4348 (8)0.064 (6)*
H4A0.613 (3)0.9197 (18)0.2214 (9)0.091 (7)*
H2A0.219 (3)0.2538 (16)0.4034 (8)0.059 (5)*
H2B0.294 (3)0.3593 (16)0.3773 (8)0.070 (6)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O10.0619 (6)0.0247 (5)0.0345 (5)0.0050 (4)0.0037 (4)0.0047 (4)
O30.0674 (7)0.0250 (5)0.0287 (5)0.0012 (4)0.0088 (4)0.0011 (4)
O20.0852 (8)0.0237 (5)0.0359 (5)0.0046 (5)0.0142 (5)0.0037 (4)
O40.0817 (9)0.0348 (5)0.0332 (5)0.0152 (5)0.0041 (5)0.0024 (4)
O50.0897 (9)0.0405 (6)0.0296 (5)0.0093 (6)0.0004 (5)0.0009 (4)
N10.0524 (7)0.0227 (5)0.0300 (6)0.0019 (5)0.0053 (5)0.0025 (4)
N20.0812 (10)0.0263 (6)0.0343 (7)0.0064 (6)0.0067 (6)0.0021 (5)
C70.0412 (7)0.0286 (6)0.0305 (7)0.0016 (5)0.0014 (6)0.0023 (5)
C80.0480 (8)0.0251 (6)0.0305 (7)0.0028 (5)0.0064 (6)0.0000 (5)
C10.0482 (8)0.0254 (6)0.0320 (7)0.0008 (5)0.0016 (6)0.0032 (5)
C20.0448 (7)0.0298 (7)0.0327 (7)0.0001 (6)0.0008 (6)0.0028 (5)
C110.0482 (8)0.0313 (6)0.0299 (7)0.0007 (6)0.0033 (6)0.0007 (5)
C100.0562 (9)0.0355 (7)0.0268 (7)0.0066 (6)0.0029 (6)0.0002 (5)
C90.0628 (9)0.0304 (7)0.0264 (6)0.0021 (6)0.0019 (6)0.0006 (5)
C60.0558 (9)0.0304 (7)0.0419 (8)0.0017 (6)0.0014 (7)0.0037 (6)
C30.0575 (9)0.0467 (8)0.0354 (8)0.0010 (7)0.0044 (7)0.0105 (6)
C50.0630 (10)0.0488 (9)0.0408 (8)0.0083 (8)0.0001 (7)0.0130 (7)
C40.0636 (10)0.0635 (10)0.0303 (7)0.0083 (8)0.0046 (7)0.0001 (7)
Geometric parameters (Å, º) top
O1—C11.3456 (16)C7—C61.375 (2)
O1—C21.4034 (16)C8—C91.4918 (18)
O3—C81.2580 (15)C2—C31.370 (2)
O2—C81.2519 (16)C11—C101.4839 (17)
O4—C111.3148 (17)C10—H100.9300
O4—H4A1.00 (2)C10—C91.302 (2)
O5—C111.2035 (16)C9—H90.9300
N1—C71.4022 (17)C6—H60.9300
N1—C11.3231 (17)C6—C51.391 (2)
N1—H10.99 (2)C3—H30.9300
N2—C11.2991 (18)C3—C41.385 (2)
N2—H2A0.90 (2)C5—H50.9300
N2—H2B0.94 (2)C5—C41.385 (2)
C7—C21.3785 (19)C4—H40.9300
C1—O1—C2105.82 (10)O5—C11—O4123.88 (12)
C11—O4—H4A109.9 (13)O5—C11—C10124.05 (13)
C7—N1—H1127.8 (11)C11—C10—H10118.7
C1—N1—C7107.72 (11)C9—C10—C11122.53 (13)
C1—N1—H1124.4 (11)C9—C10—H10118.7
C1—N2—H2A123.0 (12)C8—C9—H9117.3
C1—N2—H2B116.4 (12)C10—C9—C8125.44 (13)
H2A—N2—H2B120.4 (17)C10—C9—H9117.3
C2—C7—N1106.32 (11)C7—C6—H6121.8
C6—C7—N1132.90 (13)C7—C6—C5116.49 (14)
C6—C7—C2120.77 (13)C5—C6—H6121.8
O3—C8—C9119.97 (11)C2—C3—H3122.5
O2—C8—O3123.71 (12)C2—C3—C4115.09 (14)
O2—C8—C9116.32 (11)C4—C3—H3122.5
N1—C1—O1111.91 (11)C6—C5—H5119.2
N2—C1—O1120.18 (12)C4—C5—C6121.59 (14)
N2—C1—N1127.90 (13)C4—C5—H5119.2
C7—C2—O1108.22 (11)C3—C4—C5122.02 (14)
C3—C2—O1127.76 (13)C3—C4—H4119.0
C3—C2—C7124.02 (14)C5—C4—H4119.0
O4—C11—C10112.07 (11)
O1—C2—C3—C4178.81 (15)C1—O1—C2—C70.54 (15)
O3—C8—C9—C1019.2 (2)C1—O1—C2—C3179.11 (15)
O2—C8—C9—C10161.08 (16)C1—N1—C7—C20.81 (16)
O4—C11—C10—C9176.32 (15)C1—N1—C7—C6177.88 (16)
O5—C11—C10—C94.1 (3)C2—O1—C1—N11.09 (16)
N1—C7—C2—O10.15 (15)C2—O1—C1—N2177.33 (14)
N1—C7—C2—C3179.82 (14)C2—C7—C6—C50.0 (2)
N1—C7—C6—C5178.51 (15)C2—C3—C4—C50.3 (3)
C7—N1—C1—O11.21 (17)C11—C10—C9—C8179.35 (14)
C7—N1—C1—N2177.07 (15)C6—C7—C2—O1178.73 (13)
C7—C2—C3—C40.8 (2)C6—C7—C2—C30.9 (2)
C7—C6—C5—C41.1 (2)C6—C5—C4—C31.2 (3)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
C6—H6···O1i0.932.583.4929 (17)167
C3—H3···O5ii0.932.513.2563 (17)137
N1—H1···O30.99 (2)1.69 (2)2.6720 (14)175.8 (17)
O4—H4A···O2iii1.00 (2)1.60 (2)2.5913 (14)174 (2)
N2—H2A···O3iv0.90 (2)1.94 (2)2.8355 (16)176.4 (18)
N2—H2B···O20.94 (2)1.89 (2)2.7873 (16)157.8 (18)
Symmetry codes: (i) x+1/2, y+1/2, z; (ii) x+1/2, y+1, z+1/2; (iii) x+1, y+1/2, z+1/2; (iv) x+1/2, y1/2, z.
Symmetry codes, binding types and interaction energies (kcal mol-1) of the building units (dimers) with neighbours top
Pair of building unitsSymmetry operation of neighbouring building unitEintInteraction
1-1/2 + x, 1/2 - y, 1 - z-2.1Non-specific
2-1/2 + x, 3/2 - y, 1 - z-1.6Non-specific
3-x, 1 - y, 1 - z-10.6Stacking
4-1/2 + x, y, 1/2 - z-3.0Non-specific
51/2 + x, 1/2 - y, 1 - z-2.1Non-specific
61/2 - x, 1 - y, -1/2 + z-1.4C3—H3···O5
7x, 3/2 - y, -1/2 + z-0.9Non-specific
81/2 - x, 1 - y, 1/ 2 + z-1.4C3—H3···O5
91 - x, -1/2 + y, 1/2 - z-11.4O4—H4A···O2
101/2 + x, 3/2 - y, 1 - z-1.6Non-specific
111 - x, 1 - y, 1 - z-10.0Stacking
121/2 + x, y, 1/2 - z-3.0Non-specific
13x, 3/2 - y, 1/2 + z-0.9Non-specific
141/2 - x, -1/2 + y, z-10.3N2—H2A···O3, C6—H6···O1
151 - x, 1/2 + y, 1/2 - z-11.4O4—H4A···O2
161/2 - x, 1/2 + y, z-10.3N2—H2A···O3, C6—H6···O1
 

Acknowledgements

We gratefully acknowledge the help with quantum-chemical calculations from the Department of X-ray Diffraction Studies and Quantum Chemistry (SSI Institute for Single Crystals of the National Academy of Sciences of Ukraine) and would like to thank Dr Yevhenii Vaksler for the criticism and advice on the way to conduct the calculations of the pairwise inter­action energies for the charged systems.

References

First citationAli, S., Omprakash, P., Tengli, A. K., Mathew, B., Basavaraj, M. V., Parkali, P., Chandan, R. S. & Kumar, A. S. (2022). Polycyclic Aromat. Compd. 30, 1–34.  Google Scholar
First citationBarone, V. & Cossi, M. (1998). J. Phys. Chem. A, 102, 1995–2001.  Web of Science CrossRef CAS Google Scholar
First citationBatchelor, E., Klinowski, J. & Jones, W. (2000). J. Mater. Chem. 10, 839–848.  Web of Science CSD CrossRef CAS Google Scholar
First citationBecke, A. D. (1997). J. Chem. Phys. 107, 8554–8560.  CrossRef CAS Web of Science Google Scholar
First citationBernstein, J., Davis, R. E., Shimoni, L. & Chang, N.-L. (1995). Angew. Chem. Int. Ed. Engl. 34, 1555–1573.  CrossRef CAS Web of Science Google Scholar
First citationBlagden, N., de Matas, M., Gavan, P. T. & York, P. (2007). Adv. Drug Deliv. Rev. 59, 617–630.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBondi, A. (1964). J. Phys. Chem. 68, 441–451.  CrossRef CAS Web of Science Google Scholar
First citationBoys, S. F. & Bernardi, F. (1970). Mol. Phys. 19, 553–566.  CrossRef CAS Web of Science Google Scholar
First citationCallear, S. K., Hursthouse, M. B. & Threlfall, T. L. (2009). CrystEngComm, 11, 1609–1614.  Web of Science CSD CrossRef CAS Google Scholar
First citationColegate, S., Dorling, P., Huxtable, C., Shaw, T., Skelton, B., Vogel, P. & White, A. (1989). Aust. J. Chem. 42, 1249–1255.  CrossRef CAS Google Scholar
First citationColeman, N., Brown, B. M., Oliván-Viguera, A., Singh, V., Olmstead, M. M., Valero, M. S., Köhler, R. & Wulff, H. (2014). Mol. Pharmacol. 86, 342–357.  CrossRef PubMed Google Scholar
First citationDavidson, E. R. (1996). Chem. Phys. Lett. 260, 514–518.  CrossRef CAS Web of Science Google Scholar
First citationDolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339–341.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationDunning, T. H. Jr (1989). J. Chem. Phys. 90, 1007–1023.  CrossRef CAS Web of Science Google Scholar
First citationErol, M., Celik, I., Uzunhisarcikli, E. & Kuyucuklu, G. (2022). Polycycl. Aromat. Compd. 42, 1679–1696.  CrossRef CAS Google Scholar
First citationEtter, M. C. (1990). Acc. Chem. Res. 23, 120–126.  CrossRef CAS Web of Science Google Scholar
First citationFan, L., Luo, Z., Yang, C., Guo, B., Miao, J., Chen, Y., Tang, L. & Li, Y. (2022). Mol. Divers. 26, 981–992.  CrossRef CAS PubMed Google Scholar
First citationFranklin, S. & Balasubramanian, T. (2009). Acta Cryst. C65, o58–o61.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationGrimme, S., Antony, J., Ehrlich, S. & Krieg, H. (2010). J. Chem. Phys. 132, 154104.  Web of Science CrossRef PubMed Google Scholar
First citationGrimme, S., Ehrlich, S. & Goerigk, L. (2011). J. Comput. Chem. 32, 1456–1465.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGuillory, J. K. (2003). J. Med. Chem. 46, 1277–1277.  CrossRef CAS Google Scholar
First citationImaizumi, T., Otsubo, S., Komai, M., Takada, H., Maemoto, M., Kobayashi, A. & Otsubo, N. (2020). Bioorg. Med. Chem. 28, 115622.  CrossRef PubMed Google Scholar
First citationKendall, R. A., Dunning, T. H. Jr & Harrison, R. J. (1992). J. Chem. Phys. 96, 6796–6806.  CrossRef CAS Web of Science Google Scholar
First citationKonovalova, I. S., Shishkina, S. V., Paponov, B. V. & Shishkin, O. V. (2010). CrystEngComm, 12, 909–916.  Web of Science CSD CrossRef CAS Google Scholar
First citationKruszynski, R. & Trzesowska-Kruszynska, A. (2010). Acta Cryst. C66, o449–o454.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationLynch, D. E. (2004). Acta Cryst. E60, o1715–o1716.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationLynch, D. E. (2009). CSD Communication (refcode XUVPAG, CCDC 716694). CCDC, Cambridge, England.  Google Scholar
First citationLynch, D. E., Barfield, J., Frost, J., Antrobus, R. & Simmons, J. (2003). Cryst. Eng. 6, 109–122.  Web of Science CSD CrossRef CAS Google Scholar
First citationLynch, D. E., Daly, D. & Parsons, S. (2000a). Acta Cryst. C56, 1478–1479.  CrossRef CAS IUCr Journals Google Scholar
First citationLynch, D. E., Singh, M. & Parsons, S. (2000b). Cryst. Eng. 3, 71–79.  CrossRef CAS Google Scholar
First citationNeese, F., Wennmohs, F., Becker, U. & Riplinger, C. (2020). J. Chem. Phys. 152, 224108.  Web of Science CrossRef PubMed Google Scholar
First citationParamashivappa, R., Phani Kumar, P., Subba Rao, P. V. & Srinivasa Rao, A. (2003). Bioorg. Med. Chem. Lett. 13, 657–660.  CrossRef PubMed CAS Google Scholar
First citationParlapalli, A. & Manda, S. (2017). J. Chem. Pharm. Res. 9(9), 57–62.  Google Scholar
First citationPeterson, K. A., Woon, D. E. & Dunning, T. H. Jr (1994). J. Chem. Phys. 100, 7410–7415.  CrossRef CAS Google Scholar
First citationRigaku OD (2021). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.  Google Scholar
First citationRowland, R. S. & Taylor, R. (1996). J. Phys. Chem. 100, 7384–7391.  CrossRef CAS Web of Science Google Scholar
First citationRynearson, K. D., Charrette, B., Gabriel, C., Moreno, J., Boerneke, M. A., Dibrov, S. M. & Hermann, T. (2014). Bioorg. Med. Chem. Lett. 24, 3521–3525.  CrossRef CAS PubMed Google Scholar
First citationSangi, D. P., Meira, Y. G., Moreira, N. M., Lopes, T. A., Leite, M. P., Pereira-Flores, M. E. & Alvarenga, E. S. (2019). Pest Manag. Sci. 75, 262–269.  CrossRef CAS PubMed Google Scholar
First citationSattar, R., Mukhtar, R., Atif, M., Hasnain, M. & Irfan, A. (2020). J. Heterocycl. Chem. 57, 2079–2107.  CrossRef CAS Google Scholar
First citationSchmider, H. L. & Becke, A. D. (1998). J. Chem. Phys. 108, 9624–9631.  CrossRef CAS Google Scholar
First citationSheikh, A. Y., Rahim, S. A., Hammond, R. B. & Roberts, K. J. (2009). CrystEngComm, 11, 501–509.  CrossRef CAS Google Scholar
First citationSheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationShishkin, O. V., Dyakonenko, V. V. & Maleev, A. V. (2012). CrystEngComm, 14, 1795–1804.  Web of Science CrossRef CAS Google Scholar
First citationSilva, L. A., Hottes, E., da Silva, A. M., Freire, L. M. S., Guedes, G. P. & Ferreira, A. B. B. (2021). J. Braz. Chem. Soc. 32, 1009–1016.  CAS Google Scholar
First citationŠlachtová, V. & Brulíková, L. (2018). ChemistrySelect 3, 4653–4662.  Google Scholar
First citationSoares, F. L. F. & Carneiro, R. L. (2014). J. Pharm. Biomed. Anal. 89, 166–175.  CrossRef CAS PubMed Google Scholar
First citationSonawane, A. R., Rawat, S. S. & Janolkar, N. N. (2013). Asian J. Biomed. Pharm. Sci. 27, 1–8.  Google Scholar
First citationSpackman, M. A. & Byrom, P. G. (1997). Chem. Phys. Lett. 267, 215–220.  CrossRef CAS Web of Science Google Scholar
First citationSpackman, M. A. & McKinnon, J. J. (2002). CrystEngComm, 4, 378–392.  Web of Science CrossRef CAS Google Scholar
First citationSpackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006–1011.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWilson, A. K., van Mourik, T. & Dunning, T. H. Jr (1996). J. Mol. Struct. Theochem, 388, 339–349.  CrossRef CAS Google Scholar
First citationWoon, D. E. & Dunning, T. H. Jr (1993). J. Chem. Phys. 98, 1358–1371.  CrossRef CAS Web of Science Google Scholar
First citationXu, W., Lu, Y., Xia, Y., Liu, B., Jin, S., Zhong, B., Wang, D. & Guo, M. (2019). J. Mol. Struct. 1189, 81–93.  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.

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds