Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-25T17:12:35.785Z Has data issue: false hasContentIssue false

Crystal structure of oxfendazole, C15H13N3O3S

Published online by Cambridge University Press:  27 January 2023

James A. Kaduk*
Affiliation:
Illinois Institute of Technology, 3101 S. Dearborn St., Chicago, IL 60616, USA North Central College, 131 S. Loomis St., Naperville, IL 60540, USA
Stacy Gates-Rector
Affiliation:
ICDD, 12 Campus Blvd., Newtown Square, PA 19073-3273, USA
Thomas N. Blanton
Affiliation:
ICDD, 12 Campus Blvd., Newtown Square, PA 19073-3273, USA
*
a)Author to whom correspondence should be addressed. Electronic mail: kaduk@polycrystallography.com
Rights & Permissions [Opens in a new window]

Abstract

The crystal structure of oxfendazole has been solved and refined using synchrotron X-ray powder diffraction data, and optimized using density functional theory techniques. Oxfendazole crystallizes in space group P21/c (#14) with a = 18.87326(26), b = 10.40333(5), c = 7.25089(5) Å, β = 91.4688(10)° V = 1423.206(10) Å3, and Z = 4. The crystal structure consists of stacks of the planar portions of the L-shaped molecules, resulting in layers parallel to the bc-plane. Only weak hydrogen bonds are present. The powder pattern has been submitted to ICDD for inclusion in the Powder Diffraction File™ (PDF®).

Type
New Diffraction Data
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of International Centre for Diffraction Data

I. INTRODUCTION

Oxfendazole (sold under the brand name Synanthic®) is a broad spectrum benzimidazole anthelmintic. Its main use is for protecting livestock against roundworm, strongyles (bloodworms), and pinworms. The systematic name (CAS Registry Number 53716-50-0) is methyl N-[6-(benzenesulfinyl)-1H-benzimidazol-2-yl]carbamate. A two-dimensional (2D) molecular diagram is shown in Figure 1.

Figure 1. The 2D molecular structure of oxfendazole.

Suspensions of oxfendazole nanocrystals have been studied for use as an oral formulation, and an X-ray powder diffraction pattern of the API has been provided (Sun et al., Reference Sun, Chen, Zhao, Zhou, Zhang, Wang and Xie2022).

This work was carried out as part of a project (Kaduk et al., Reference Kaduk, Crowder, Zhong, Fawcett and Suchomel2014) to determine the crystal structures of large-volume commercial pharmaceuticals, and include high-quality powder diffraction data for them in the Powder Diffraction File (Gates-Rector and Blanton, Reference Gates-Rector and Blanton2019).

II. EXPERIMENTAL

Oxfendazole was a commercial reagent, purchased from TargetMol (Lot #113808), and was used as-received. The white powder was packed into a 1.5-mm diameter Kapton capillary, and rotated during the measurement at ~50 Hz. The powder pattern was measured at 295 K at beamline 11-BM (Antao et al., Reference Antao, Hassan, Wang, Lee and Toby2008; Lee et al., Reference Lee, Shu, Ramanathan, Preissner, Wang, Beno, Von Dreele, Ribaud, Kurtz, Antao, Jiao and Toby2008; Wang et al., Reference Wang, Toby, Lee, Ribaud, Antao, Kurtz, Ramanathan, Von Dreele and Beno2008) of the Advanced Photon Source at Argonne National Laboratory using a wavelength of 0.458208(2) Å from 0.5° to 50° 2θ with a step size of 0.001° and a counting time of 0.1 s/step. The high-resolution powder diffraction data were collected using twelve silicon crystal analyzers that allow for high angular resolution, high precision, and accurate peak positions. A silicon (NIST SRM 640c) and alumina (SRM 676a) standard (ratio Al2O3:Si = 2:1 by weight) was used to calibrate the instrument and refine the monochromatic wavelength used in the experiment.

The pattern was indexed using both JADE Pro (MDI, 2022) and N-TREOR (Altomare et al., Reference Altomare, Cuocci, Giacovazzo, Moliterni, Rizzi, Corriero and Falcicchio2013) on a primitive monoclinic unit cell with a = 18.85890, b = 10.40423, c = 7.25538 Å, β = 91.553°, V = 1423.1 Å3, and Z = 4. A reduced cell search in the Cambridge Structural Database (Groom et al., Reference Groom, Bruno, Lightfoot and Ward2016) yielded 13 hits, but no structures of oxfendazole derivatives. The suggested space group was P21/c, which was confirmed by successful solution and refinement of the structure.

An oxfendazole molecule was downloaded from PubChem (Kim et al., Reference Kim, Chen, Cheng, Gindulyte, He, He, Li, Shoemaker, Thiessen, Yu, Zaslavsky, Zhang and Bolton2019) as Conformer3D_CID_40854.sdf. It was converted to a *.mol2 file using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020). The structure was solved by Monte Carlo simulated annealing techniques as implemented in EXPO2014 (Altomare et al., Reference Altomare, Cuocci, Giacovazzo, Moliterni, Rizzi, Corriero and Falcicchio2013).

Rietveld refinement was carried out using GSAS-II (Toby and Von Dreele, Reference Toby and Von Dreele2013). Only the 1.0–25.0° portion of the pattern was included in the refinement (d min = 1.058 Å). The region 1.53–2.07° 2θ, which contained a broad peak from the Kapton capillary, was excluded. All non-H bond distances and angles were subjected to restraints, based on a Mercury/Mogul Geometry check (Bruno et al., Reference Bruno, Cole, Kessler, Luo, Motherwell, Purkis, Smith, Taylor, Cooper, Harris and Orpen2004; Sykes et al., Reference Sykes, McCabe, Allen, Battle, Bruno and Wood2011). The Mogul average and standard deviation for each quantity were used as the restraint parameters. The restraints contributed 3.1% to the final χ 2. The hydrogen atoms were included in calculated positions, which were recalculated during the refinement using Materials Studio (Dassault, 2021). The U iso of the heavy atoms were grouped by chemical similarity. The U iso for the H atoms were fixed at 1.2× the U iso of the heavy atoms to which they are attached. The peak profiles were described using the generalized microstrain model. The background was modeled using a 6-term shifted Chebyshev polynomial, and a peak at 5.87° 2θ to model the scattering from the Kapton capillary and any amorphous component.

The final refinement of 94 variables using 23,501 observations and 57 restraints yielded the residuals R wp = 0.0785 and GOF = 1.56. The largest peak (1.49 Å from C18) and hole (1.36 Å from C19) in the difference Fourier map were 0.23(5) and −0.21(5) eÅ−3, respectively. The largest errors in the difference plot (Figure 2) are in the shape of the strong lowest-angle (1 0 0) peak.

Figure 2. The Rietveld plot for the refinement of oxfendazole. The blue crosses represent the observed data points, and the green line is the calculated pattern. The cyan curve is the normalized error plot. The red curve indicates the background. The vertical scale has been multiplied by a factor of 20× for 2θ > 10.0°. The row of blue tick marks indicates the calculated reflection positions.

The crystal structure was optimized using VASP (Kresse and Furthmüller, Reference Kresse and Furthmüller1996) (fixed experimental unit cell) through the MedeA graphical interface (Materials Design, 2016). The calculation was carried out on 16 2.4 GHz processors (each with 4 GB RAM) of a 64-processor HP Proliant DL580 Generation 7 Linux cluster at North Central College. The calculation used the GGA-PBE functional, a plane wave cutoff energy of 400.0 eV, and a k-point spacing of 0.5 Å−1 leading to a 1 × 2 × 2 mesh, and took ~15.1 h. A single-point density functional theory calculation (fixed experimental cell) and population analysis were carried out using CRYSTAL17 (Dovesi et al., Reference Dovesi, Erba, Orlando, Zicovich-Wilson, Civalleri, Maschio, Rerat, Casassa, Baima, Salustro and Kirtman2018). The basis sets for the H, C, N, and O atoms in the calculation were those of Gatti et al. (Reference Gatti, Saunders and Roetti1994), and that for S was that of Peintinger et al. (Reference Peintinger, Vilela Oliveira and Bredow2013). The calculations were run on a 3.5 GHz PC using 8 k-points and the B3LYP functional, and took ~1.9 h.

III. RESULTS AND DISCUSSION

The synchrotron powder pattern of this study matches the pattern for oxfendazole (in a single-column figure) reported by Sun et al. (Reference Sun, Chen, Zhao, Zhou, Zhang, Wang and Xie2022) well enough to conclude that they represent the same material (Figure 3), and thus that this material is a representative sample. Sun et al. did not start at an angle low-enough to measure the lowest-angle peak. The root-mean-square Cartesian displacement between the Rietveld-refined and DFT-optimized structures is 0.183 Å (Figure 4), and the maximum difference is 0.426 Å, at O4. The excellent agreement is strong evidence that the structure is correct (van de Streek and Neumann, Reference van de Streek and Neumann2014). This discussion concentrates on the DFT-optimized structure. The asymmetric unit (with atom numbering) is illustrated in Figure 5. The best view of the crystal structure is down the b-axis (Figure 6). The crystal structure consists of stacks of the planar portions of the L-shaped molecules. The mean plane of the benzimidazole ring system is approximately 8,1,10, and the plane of the phenyl ring is approximately 13,8,−5. The Mercury Aromatics Analyser indicates three moderate interactions: one π–π with a distance of 3.80 Å, one H–π with a distance of 5.32 Å, and one side-side with a distance of 6.42 Å.

Figure 3. Comparison of the synchrotron pattern of oxfendazole (black) to that reported by Sun et al. (Reference Sun, Chen, Zhao, Zhou, Zhang, Wang and Xie2022; green). The literature pattern, measured using Cu radiation, was digitized using UN-SCAN-IT (Silk Scientific, 2013), and converted to the synchrotron wavelength of 0.458208 Å using JADE Pro (MDI, 2022), which was also used to generate the image.

Figure 4. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of oxfendazole. The rms Cartesian displacement is 0.183 Å. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 5. The asymmetric unit of oxfendazole, with the atom numbering. The atoms are represented by 50% probability spheroids/ellipsoids. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 6. The crystal structure of oxfendazole, viewed down the b-axis. Image generated using Diamond (Crystal Impact, 2022).

All of the bond distances and bond angles fall within the normal ranges indicated by a Mercury/Mogul Geometry check (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020). The C8–C11–C10 angle of 115.8° in the benzimidazole ring system is flagged as slightly unusual (average = 120.1(14)°, Z-score = 3.1). The torsion angles involving rotation about the C17–N7 bond are flagged as unusual. These lie in a minor population of a distribution with the majority rotated by 180°. They represent the orientation of the side chain with respect to the benzimidazole ring system, and indicate that the conformation of the molecule in the solid state is unusual.

Quantum chemical geometry optimization of the oxfendazole molecule (DFT/B3LYP/6-31G*/water) using Spartan ‘18 (Wavefunction, 2020) indicated that in the solid state it is in essentially a local minimum energy conformation (rms displacement = 0.174 Å). A conformational analysis (MMFF force field) indicates that the global minimum-energy conformation has the side chain rotation approximately 180° from the observed conformation. Intermolecular interactions seem to be important in determining the solid-state conformation.

Analysis of the contributions to the total crystal energy of the structure using the Forcite module of Materials Studio (Dassault, 2021) suggests that the intramolecular deformation energy is dominated by angle distortion terms, as might be expected for a fused ring system. The intermolecular energy is dominated by electrostatic attractions, which in this force field analysis include hydrogen bonds. The hydrogen bonds are better analyzed using the results of the DFT calculation.

Only weak hydrogen bonds are present in the structure (Table I). There are two N–H⋯O hydrogen bonds to the carbonyl group O4. The energies of these bonds were calculated using the correlation of Wheatley and Kaduk (Reference Wheatley and Kaduk2019). N7–H27 makes an additional intermolecular N–H⋯N hydrogen bond to N5. Several C–H⋯O hydrogen bonds (one intramolecular) also contribute to the lattice energy.

TABLE I. Hydrogen bonds (CRYSTAL17) in oxfendazole.

a Intramolecular.

The volume enclosed by the Hirshfeld surface of oxfendazole (Figure 7, Hirshfeld, Reference Hirshfeld1977; Turner et al., Reference Turner, McKinnon, Wolff, Grimwood, Spackman, Jayatilaka and Spackman2017) is 348.53 Å3, 97.96% of 1/4 the unit cell volume. The packing density is thus fairly typical. The only significant close contacts (red in Figure 7) involve the hydrogen bonds. The volume/non-hydrogen atom is smaller than normal at 16.2 Å3.

Figure 7. The Hirshfeld surface of oxfendazole. Intermolecular contacts longer than the sums of the van der Waals radii are colored blue, and contacts shorter than the sums of the radii are colored red. Contacts equal to the sums of radii are white. Image generated using CrystalExplorer (Turner et al., Reference Turner, McKinnon, Wolff, Grimwood, Spackman, Jayatilaka and Spackman2017).

The Bravais–Friedel–Donnay–Harker (Bravais, Reference Bravais1866; Friedel, Reference Friedel1907; Donnay and Harker, Reference Donnay and Harker1937) morphology suggests that we might expect platy morphology for oxfendazole, with {100} as principal faces. A second-order spherical harmonic preferred orientation model was included in the refinement. The texture index was 1.000(0), indicating that preferred orientation was not significant for this rotated capillary specimen. The powder pattern of oxfendazole from this synchrotron data set has been submitted to ICDD for inclusion in the Powder Diffraction File.

IV. DEPOSITED DATA

The Crystallographic Information Framework (CIF) files containing the results of the Rietveld refinement (including the raw data) and the DFT geometry optimization were deposited with the ICDD. The data can be requested at .

ACKNOWLEDGMENTS

The use of the Advanced Photon Source at Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-06CH11357. This work was partially supported by the International Centre for Diffraction Data. We thank Lynn Ribaud and Saul Lapidus for their assistance in the data collection.

CONFLICTS OF INTEREST

The authors have no conflicts of interest to declare.

References

REFERENCES

Altomare, A., Cuocci, C., Giacovazzo, C., Moliterni, A., Rizzi, R., Corriero, N., and Falcicchio, A.. 2013. “EXPO2013: A Kit of Tools for Phasing Crystal Structures from Powder Data.” Journal of Applied Crystallography 46: 1231–5.CrossRefGoogle Scholar
Antao, S. M., Hassan, I., Wang, J., Lee, P. L., and Toby, B. H.. 2008. “State-of-the-Art High-Resolution Powder X-Ray Diffraction (HRPXRD) Illustrated with Rietveld Refinement of Quartz, Sodalite, Tremolite, and Meionite.” Canadian Mineralogist 46: 1501–9.CrossRefGoogle Scholar
Bravais, A. 1866. Etudes Cristallographiques. Paris, Gauthier Villars.Google Scholar
Bruno, I. J., Cole, J. C., Kessler, M., Luo, J., Motherwell, W. D. S., Purkis, L. H., Smith, B. R., Taylor, R., Cooper, R. I., Harris, S. E., and Orpen, A. G.. 2004. “Retrieval of Crystallographically-Derived Molecular Geometry Information.” Journal of Chemical Information and Computer Sciences 44: 2133–44.CrossRefGoogle ScholarPubMed
Crystal Impact. 2022. Diamond – Crystal and Molecular Structure Visualization. Bonn, Germany, Crystal Impact – Dr. H. Putz & Dr. K. Brandenburg. https://www.crystalimpact.de/diamond.Google Scholar
Dassault Systèmes. 2021. Materials Studio 2021. San Diego, CA, BIOVIA.Google Scholar
Donnay, J. D. H., and Harker, D.. 1937. “A New Law of Crystal Morphology Extending the Law of Bravais.” American Mineralogist 22: 446–7.Google Scholar
Dovesi, R., Erba, A., Orlando, R., Zicovich-Wilson, C. M., Civalleri, B., Maschio, L., Rerat, M., Casassa, S., Baima, B., Salustro, J., and Kirtman, B.. 2018. “Quantum-Mechanical Condensed Matter Simulations with CRYSTAL.” Wiley Interdisciplinary Reviews: Computational Molecular Science 8: e1360.Google Scholar
Friedel, G. 1907. “Etudes sur la loi de Bravais.” Bulletin de la Société Française de Minéralogie 30: 326455.CrossRefGoogle Scholar
Gates-Rector, S., and Blanton, T.. 2019. “The Powder Diffraction File: A Quality Materials Characterization Database.” Powder Diffraction 39 (4): 352–60.CrossRefGoogle Scholar
Gatti, C., Saunders, V. R., and Roetti, C.. 1994. “Crystal-Field Effects on the Topological Properties of the Electron-Density in Molecular Crystals - The Case of Urea.” Journal of Chemical Physics 101: 10686–96.CrossRefGoogle Scholar
Groom, C. R., Bruno, I. J., Lightfoot, M. P., and Ward, S. C.. 2016. “The Cambridge Structural Database.” Acta Crystallographica Section B: Structural Science, Crystal Engineering and Materials 72: 171–9.CrossRefGoogle ScholarPubMed
Hirshfeld, F. L. 1977. “Bonded-Atom Fragments for Describing Molecular Charge Densities.” Theoretica Chimica Acta 44: 129–38.CrossRefGoogle Scholar
Kaduk, J. A., Crowder, C. E., Zhong, K., Fawcett, T. G., and Suchomel, M. R.. 2014. “Crystal Structure of Atomoxetine Hydrochloride (Strattera), C17H22NOCl.” Powder Diffraction 29 (3): 269–73.CrossRefGoogle Scholar
Kim, S., Chen, J., Cheng, T., Gindulyte, A., He, J., He, S., Li, Q., Shoemaker, B. A., Thiessen, P. A., Yu, B., Zaslavsky, L., Zhang, J., and Bolton, E. E.. 2019. “PubChem 2019 Update: Improved Access to Chemical Data.” Nucleic Acids Research 47 (D1): D1102–9. doi:10.1093/nar/gky1033.CrossRefGoogle ScholarPubMed
Kresse, G., and Furthmüller, J.. 1996. “Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set.” Computational Materials Science 6: 1550.CrossRefGoogle Scholar
Lee, P. L., Shu, D., Ramanathan, M., Preissner, C., Wang, J., Beno, M. A., Von Dreele, R. B., Ribaud, L., Kurtz, C., Antao, S. M., Jiao, X., and Toby, B. H.. 2008. “A Twelve-Analyzer Detector System for High-Resolution Powder Diffraction.” Journal of Synchrotron Radiation 15 (5): 427–32.CrossRefGoogle ScholarPubMed
Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M., and Wood, P. A.. 2020. “Mercury 4.0: From Visualization to Design and Prediction.” Journal of Applied Crystallography 53: 226–35.CrossRefGoogle ScholarPubMed
Materials Design. 2016. MedeA 2.20.4. Angel Fire, NM, Materials Design Inc.Google Scholar
MDI. 2022. JADE Pro version 8.2 (Computer software). Livermore, CA, USA, Materials Data.Google Scholar
Peintinger, M. F., Vilela Oliveira, D., and Bredow, T.. 2013. “Consistent Gaussian Basis Sets of Triple-Zeta Valence with Polarization Quality for Solid-State Calculations.” Journal of Computational Chemistry 34: 451–9.CrossRefGoogle ScholarPubMed
Silk Scientific. 2013. UN-SCAN-IT 7.0. Orem, UT, Silk Scientific Corporation.Google Scholar
Sun, Y., Chen, D., Zhao, Y., Zhou, K., Zhang, B., Wang, H., and Xie, S.. 2022. “Exploitation of Nanocrystal Suspension as an Effective Oral Formulation for Oxfendazole.” Drug Delivery and Translational Research 12: 1219–29.CrossRefGoogle ScholarPubMed
Sykes, R. A., McCabe, P., Allen, F. H., Battle, G. M., Bruno, I. J., and Wood, P. A.. 2011. “New Software for Statistical Analysis of Cambridge Structural Database Data.” Journal of Applied Crystallography 44: 882–6.CrossRefGoogle Scholar
Toby, B. H., and Von Dreele, R. B.. 2013. “GSAS II: the Genesis of a Modern Open Source All Purpose Crystallography Software Package.” Journal of Applied Crystallography 46: 544–9.CrossRefGoogle Scholar
Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D., and Spackman, M. A.. 2017. CrystalExplorer17. University of Western Australia. http://hirshfeldsurface.net.Google Scholar
van de Streek, J., and Neumann, M. A.. 2014. “Validation of Molecular Crystal Structures from Powder Diffraction Data with Dispersion-Corrected Density Functional Theory (DFT-D).” Acta Crystallographica Section B: Structural Science, Crystal Engineering and Materials 70 (6): 10201032.CrossRefGoogle ScholarPubMed
Wang, J., Toby, B. H., Lee, P. L., Ribaud, L., Antao, S. M., Kurtz, C., Ramanathan, M., Von Dreele, R. B., and Beno, M. A.. 2008. “A Dedicated Powder Diffraction Beamline at the Advanced Photon Source: Commissioning and Early Operational Results.” Review of Scientific Instruments 79: 085105.CrossRefGoogle ScholarPubMed
Wavefunction, Inc. 2020. Spartan ‘18 Version 1.4.5. Irvine, CA, Wavefunction Inc.Google Scholar
Wheatley, A. M., and Kaduk, J. A.. 2019. “Crystal Structures of Ammonium Citrates.” Powder Diffraction 34: 3543.CrossRefGoogle Scholar
Figure 0

Figure 1. The 2D molecular structure of oxfendazole.

Figure 1

Figure 2. The Rietveld plot for the refinement of oxfendazole. The blue crosses represent the observed data points, and the green line is the calculated pattern. The cyan curve is the normalized error plot. The red curve indicates the background. The vertical scale has been multiplied by a factor of 20× for 2θ > 10.0°. The row of blue tick marks indicates the calculated reflection positions.

Figure 2

Figure 3. Comparison of the synchrotron pattern of oxfendazole (black) to that reported by Sun et al. (2022; green). The literature pattern, measured using Cu radiation, was digitized using UN-SCAN-IT (Silk Scientific, 2013), and converted to the synchrotron wavelength of 0.458208 Å using JADE Pro (MDI, 2022), which was also used to generate the image.

Figure 3

Figure 4. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of oxfendazole. The rms Cartesian displacement is 0.183 Å. Image generated using Mercury (Macrae et al., 2020).

Figure 4

Figure 5. The asymmetric unit of oxfendazole, with the atom numbering. The atoms are represented by 50% probability spheroids/ellipsoids. Image generated using Mercury (Macrae et al., 2020).

Figure 5

Figure 6. The crystal structure of oxfendazole, viewed down the b-axis. Image generated using Diamond (Crystal Impact, 2022).

Figure 6

TABLE I. Hydrogen bonds (CRYSTAL17) in oxfendazole.

Figure 7

Figure 7. The Hirshfeld surface of oxfendazole. Intermolecular contacts longer than the sums of the van der Waals radii are colored blue, and contacts shorter than the sums of the radii are colored red. Contacts equal to the sums of radii are white. Image generated using CrystalExplorer (Turner et al., 2017).