Hostname: page-component-848d4c4894-m9kch Total loading time: 0 Render date: 2024-06-02T09:52:17.429Z Has data issue: false hasContentIssue false

Crystal structure of encorafenib, C22H27ClFN7O4S

Published online by Cambridge University Press:  03 April 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
Anja Vieira Dosen
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 encorafenib, C22H27ClFN7O4S, has been solved and refined using synchrotron X-ray powder diffraction data, and optimized using density functional theory techniques. Encorafenib crystallizes in space group P21 (#4) with a = 16.17355(25), b = 9.52334(11), c = 17.12368(19) Å, β = 89.9928(22)°, V = 2637.50(4) Å3, and Z = 4. The crystal structure consists of alternating layers of stacked halogenated phenyl rings and the other parts of the molecules perpendicular to the a-axis. One molecule participates in two strong N–H⋯N hydrogen bonds (one intra- and the other intermolecular), which are not present for the other molecule. The intermolecular hydrogen bonds link molecule 2 into a spiral chain along the b-axis. 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

Encorafenib (C22H27ClFN7O4S, sold under the brand name BRAFTOVI) is a kinase inhibitor. Encorafenib inhibits the BRAF gene, which encodes for B-raf protein, a proto-oncogene involved in various genetic mutations. This protein plays a role in regulating the MAP kinase/ERK signaling pathway, which impacts cell division, differentiation, and secretion. Mutations in this gene, most frequently the V600E mutation, are the most commonly identified cancer-causing mutations in melanoma, and have been isolated in various other cancers as well, including non-Hodgkin lymphoma, colorectal cancer, thyroid carcinoma, non-small cell lung carcinoma, hairy cell leukemia, and adenocarcinoma of the lung (DrugBank, https://go.drugbank.com/drugs/DB11718). On June 27, 2018, the Food and Drug Administration approved encorafenib and binimetinib (BRAFTOVI and MEKTOVI, Array BioPharma Inc.) in combination for patients with unresectable or metastatic melanoma with a BRAF V600E or V600K mutation, as detected by an FDA-approved test (PubChem, Kim et al., Reference Kim, Chen, Cheng, Gindulyte, He, He, Li, Shoemaker, Thiessen, Yu, Zaslavsky, Zhang and Bolton2021). The systematic name (CAS Registry Number 1269440-17-6) is methyl N-[(2S)-1-[[4-[3-[5-chloro-2-fluoro-3-(methanesulfonamido)phenyl]-1-propan-2-ylpyrazol-4-yl]pyrimidin-2-yl]amino]propan-2-yl]carbamate. A two-dimensional molecular diagram is shown in Figure 1.

Figure 1. The 2D molecular structure of encorafenib.

Encorafenib was claimed in US Patent 8,501,758 B2 (Huang et al., Reference Huang, Jin, Liu, Poon, Tellew, Wan, Wang and Xie2013; IRM and Novartis). The development of encorafenib and other protein kinase inhibitors is summarized in Koelblinger et al. (Reference Koelblinger, Thuerigen and Dummer2018) and Imran et al. (Reference Imran, Asdaq, Khan, Unnikrishnan Meenakshi, Alamri, Alsanie, Alhomrani, Mohzari, Alrashed, AlMotairi and Alkhaldi2021). A European Medicines Agency (2018) report indicates that single-crystal and powder X-ray diffraction have been carried out on encorafenib, but we are unaware of any published X-ray diffraction data on this compound.

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

Encorafenib was a commercial reagent, purchased from TargetMol (Batch #145984), 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 0.458119(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 mixture of silicon (NIST SRM 640c) and alumina (NIST SRM 676a) standards (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 initially using JADE Pro (MDI, 2022) on a high-quality primitive monoclinic unit cell with a = 16.13558, b = 17.09344, c = 9.50639 Å, β = 90.02°, V = 2621.98 Å3, and Z = 4. The β angle close to 90° suggested that the true symmetry might be orthorhombic. N-TREOR (Altomare et al., Reference Altomare, Cuocci, Giacovazzo, Moliterni, Rizzi, Corriero and Falcicchio2013) yielded an orthorhombic cell having similar dimensions, and suggested that the space group was P22121. Attempts to solve the structure in this space group were unsuccessful, yielding molecular overlap and/or voids. A close examination of the pattern revealed that the 010 peak at 1.533° was present, though weak. Attempts to solve the structure in space group P2221 were also unsuccessful, suggesting that the true space group was monoclinic P21, with an appropriate transformation of axes to yield the final cell with a = 16.17355(25), b = 9.52334(11), c = 17.12368(19) Å, β = 89.9928(22)°, and V = 2637.50(4) Å3. This means that there are two independent encorafenib molecules in the asymmetric unit. Reduced cell searches of both the monoclinic and orthorhombic cells in the Cambridge Structural Database (Groom et al., Reference Groom, Bruno, Lightfoot and Ward2016) with the chemistry C, H, Cl, F, N, O, and S only, yielded no hits.

An encorafenib molecule was downloaded from PubChem (Kim et al., Reference Kim, Chen, Cheng, Gindulyte, He, He, Li, Shoemaker, Thiessen, Yu, Zaslavsky, Zhang and Bolton2021) as Conformer3D_CID_50922675.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 using Monte Carlo simulated annealing techniques as implemented in EXPO2014 (Altomare et al., Reference Altomare, Cuocci, Giacovazzo, Moliterni, Rizzi, Corriero and Falcicchio2013). After 6 days, only the best of the ten runs yielded a structure without molecular overlap.

The encorafenib unit cell does not account for the first two peaks in the pattern at 0.684 and 1.369° 2θ (d = 38.396 and 19.172 Å; Figure 2). These diffraction peaks apparently are the first- and second-order peaks from an additional unit cell having one long dimension. Initial refinements indicated the presence of two additional unindexed peaks, at d = 4.1397 and 3.7309 Å. Boolean searches of these two peaks in the Powder Diffraction File (Gates-Rector and Blanton, Reference Gates-Rector and Blanton2019) yielded hits similar to paraffin wax (or polyethylene), characteristic of the side-to-side packing of long paraffin chains. Adding the 39.4 Å peak to the searches yielded entry 02-092-6981, for heptadecanoic acid (CSD Refcode DARWAU01; Gbabode et al., Reference Gbabode, Negrier, Mondieig, Calvo, Calvet and Cuevas-Diarte2007). This phase was added to the refinement to model these extra peaks, to avoid distorting the encorafenib structure to account for them. However, there is no evidence that heptadecanoic acid is actually present; this was merely a strategy to obtain a better refinement.

Figure 2. The synchrotron X-ray powder diffraction pattern of encorafenib indicates the two unindexed low-angle peaks. Note that the y-axis is a square root scale. Image generated using JADE Pro (MDI, 2022).

Rietveld refinement (begun from the VASP-optimized structure discussed below) was carried out using GSAS-II (Toby and Von Dreele, Reference Toby and Von Dreele2013). Only the 1.5–25.0° portion of the pattern was included in the refinement (d min = 1.058 Å). 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, Sam Motherwell, Purkis, Smith, Taylor, Cooper, Harris and Guy 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 9.0% to the final χ 2. The hydrogen atoms were included in calculated positions, which were recalculated during the refinement using Materials Studio (Dassault Systèmes, 2021). The U iso of the heavy atoms were grouped by chemical similarity. The U iso for the H atoms were fixed at 1.3× the U iso of the heavy atoms to which they are attached. The U iso of a few atoms were refined to slightly negative values, so they were fixed at reasonable values. A 2nd-order spherical harmonic preferred orientation model was included in the refinement. The refined texture index was 1.023(1). The concentration of the impurity phase refined to 3.4(1) wt%. The peak profiles were described using the generalized microstrain model. The background was modeled using a 6-term shifted Chebyshev polynomial, and a broad peak at 6.14° 2θ to model the scattering from the Kapton capillary and any amorphous component.

The final refinement (begun from the result of the VASP calculation) of 249 variables using 23,538 observations and 186 restraints yielded the residuals R wp = 0.1237 and GOF = 2.31. The largest peak (1.18 Å from C20) and hole (1.18 Å from C35) in the difference Fourier map were 0.43(10) and −0.43(10) eÅ−3, respectively. The largest errors in the difference plot (Figure 3) are in the intensities of some of the low-angle encorafenib peaks. A Le Bail fit to the pattern yielded R wp = 0.0932, so the Rietveld fit is acceptable.

Figure 3. The Rietveld plot for the refinement of encorafenib. The blue crosses represent the observed data points, and the green line is the calculated pattern. The cyan curve is the normalized error plot, and the red line is the background curve. The vertical scale has been multiplied by a factor of 2× for 2θ > 2.0°, 20× for 2θ > 7.60°, and 40× for 2θ > 15.3°.

The structure of encorafenib was optimized (fixed experimental unit cell) with density functional theory techniques using VASP (Kresse and Furthmüller, Reference Kresse and Furthmüller1996) through the MedeA graphical interface (Materials Design, 2016). The calculations were 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 × 1 mesh, and took ~74 h. A single-point density functional calculation (fixed experimental cell) and population analysis were carried out using CRYSTAL17 (Dovesi et al., Reference Dovesi, Erba, Orlando, Zicovich-Wilson, Civalleri, Maschio, Rérat, 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 those for Cl, F, and S were those of Peintinger et al. (Reference Peintinger, Oliveira and Bredow2013). The calculations were run on a 3.5 GHz PC using 8 k-points and the B3LYP functional, and took ~8.1 h.

III. RESULTS AND DISCUSSION

The root-mean-square (rms) Cartesian displacement between the Rietveld-refined and DFT-optimized structures of molecule 1 (the lower atom numbers) is 0.477 Å (Figure 4), and the maximum difference is 1.232 Å, at the methyl group C30. Other large displacements occur at the periphery of the molecule. The comparable quantities for molecule 2 are 0.622 and 2.020 Å, at the methyl group C83 (Figure 5). The orientation of the isopropyl group is significantly different in the refined and optimized structures. The agreement is outside the normal range for correct structures (van de Streek and Neumann, Reference van de Streek and Neumann2014). The number of refined parameters is large (~250), so the refined structure may be less-accurate and less-precise than usual. The current structure is the most-reasonable one we have been able to obtain. The following discussion concentrates on the DFT-optimized structure. The asymmetric unit (with atom numbering) is illustrated in Figure 6. Some U iso are larger than others, suggesting the possibility of disorder, especially in the side chains. Since the DFT optimization requires an ordered model, we concentrated on refining ordered models.

Figure 4. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of encorafenib molecule 1. The rms Cartesian displacement is 0.477 Å. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 5. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of encorafenib molecule 2. The rms Cartesian displacement is 0.622 Å. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 6. The asymmetric unit of encorafenib, with the atom numbering. The atoms are represented by 50% probability spheroids. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

The best view of the crystal structure is down the short b-axis (Figure 7). The crystal structure consists of alternating layers of stacked halogenated phenyl rings and the other parts of the molecules perpendicular to the a-axis. The Miller indices of the mean planes of the phenyl rings are 1,9,−3 and 4,7,2. The mean planes of the pyrazine rings are 9,7,13 and −7,−5,11. The mean planes of the pyrimidine rings are 11,5,7 and −7,−4,9. The Mercury Aromatics Analyser indicates only one moderate interaction with a distance of 4.12 Å between the phenyl rings of molecules 1 and 2.

Figure 7. The crystal structure of encorafenib, viewed down the b-axis. Image generated using Diamond (Crystal Impact, 2022).

Almost 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 C2–N3 bond distance of 1.702 Å (average = 1.629(13) Å; Z-score = 5.6) is flagged as unusual. Occasionally bonds involving S are unusual in a VASP-optimized structure, so this characteristic may reflect the quality of the DFT calculation. The C24–N10–C28 angle of 117.6° (average = 123.9(16)°; Z-score 4.0) is flagged as unusual. Torsion angles involving rotations about the C15–N8, C16–C22, C17–C19, and C79–C85 bonds are flagged as unusual, but lie in low-frequency portions of broad distributions. Torsion angles involving rotation about the S65–N76, N73–C91, and C86–C87 bonds lie on the tails of bimodal distributions, so are slightly unusual. Torsion involving rotation about the C28–N10 and C23–C24 bonds is truly unusual. The conformation of the ester side chain of molecule 1 seems to be unusual, consistent with the large displacement coefficients in this portion of the molecule.

The two crystallographically independent molecules have similar overall shapes, but very different conformations (Figure 8). Quantum chemical geometry optimization of the isolated encorafenib molecules (DFT/B3LYP/6-31G*/water) using Spartan ‘18 (Wavefunction, 2020) indicated that the observed conformation of molecule 1 (lower atom numbers) is 0.3 kcal/mol lower in energy than that of molecule 2. The difference is well within the expected accuracy of such calculations, so the molecules should be considered equivalent in energy and the molecule is apparently flexible. A conformational analysis (MMFF force field) shows that the minimum-energy conformation is ~6 kcal/mol lower in energy, but contains a strong intramolecular N–H⋯O hydrogen bond (N12–H55⋯O5/6 or N73–H112⋯O68/69). An isolated encorafenib molecule thus folds on itself, showing that intermolecular interactions are important in determining the solid-state conformations.

Figure 8. Comparison of the molecular structures of molecules 1 (purple) and 2 (green) in the crystal structure of encorafenib.

Analysis of the contributions to the total crystal energy of the structure using the Forcite module of Materials Studio (Dassault Systèmes, 2021) suggests that angle and torsion distortion terms dominate the intramolecular deformation energy. The intermolecular energy is dominated by van der Waals and electrostatic attractions, which in this force field analysis also include hydrogen bonds. The hydrogen bonds are better analyzed using the results of the DFT calculation.

There are only a few hydrogen bonds in the structure (Table I), and the hydrogen bonding pattern is very different for the two molecules. Molecule 2 participates in two strong N–H⋯N hydrogen bonds (one intra- and the other intermolecular), which are not present for molecule 1. The intermolecular hydrogen bonds link molecule 2 into a spiral chain along the b-axis. The remaining hydrogen bonds are non-classical C–H⋯N; N/O/Cl links.

Table I. Hydrogen bonds (CRYSTAL17) in encorafenib.

a Intramolecular.

The volume enclosed by the Hirshfeld surface of the encorafenib asymmetric unit (Figure 9; Hirshfeld, Reference Hirshfeld1977; Spackman et al., Reference Spackman, Turner, McKinnon, Wolff, Grimwood, Jayatilaka and Spackman2021) is 1307.44 Å3, 99.14% of 1/2 the unit cell volume. The packing density is thus fairly typical. The volume of the Hirshfeld surface of molecule 1 (Figure 9(a)) is 648.23 Å3, and that of molecule 2 (Figure 9(b)) is 652.81 Å3; the sum of the individual volumes is 1301 Å3. The only significant close contacts (red in Figure 9) involve the hydrogen bonds. The volume/non-hydrogen atom is typical, at 18.3 Å3.

Figure 9. (a) The Hirshfeld surface of molecule 1 of encorafenib. 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 (Spackman et al., Reference Spackman, Turner, McKinnon, Wolff, Grimwood, Jayatilaka and Spackman2021). (b) The Hirshfeld surface of molecule 2 of encorafenib. 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 (Spackman et al., Reference Spackman, Turner, McKinnon, Wolff, Grimwood, Jayatilaka and Spackman2021).

The Bravais–Friedel–Donnay–Harker (Bravais, Reference Bravais1866; Friedel, Reference Friedel1907; Donnay and Harker, Reference Donnay and Harker1937) morphology suggests that we might expect an elongated particle morphology for encorafenib, with 010 as the long axis. A 2nd-order spherical harmonic model was included in the refinement. The texture index was 1.023(1), indicating that preferred orientation was slight in this rotated capillary specimen.

IV. DEPOSITED DATA

The powder pattern of encorafenib from this synchrotron data set has been submitted to ICDD for inclusion in the Powder Diffraction File. 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 .

ACKNOWLEDGEMENTS

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, 12311235.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, 15011509.CrossRefGoogle Scholar
Bravais, A. 1866. Etudes Cristallographiques. Gauthier Villars, Paris.Google Scholar
Bruno, I. J., Cole, J. C., Kessler, M., Luo, J., Sam Motherwell, W. D., Purkis, L. H., Smith, B. R., Taylor, R., Cooper, R. I., Harris, S. E., and Guy Orpen, A.. 2004. “Retrieval of Crystallographically-Derived Molecular Geometry Information.” Journal of Chemical Information and Computer Sciences 44, 21332144.CrossRefGoogle ScholarPubMed
Crystal Impact. 2022. Diamond. V. 4.6.8. Crystal Impact - Dr. H. Putz & Dr. K. Brandenburg. Windows.Google Scholar
Dassault Systèmes. 2021. Materials Studio. V. 2021. BIOVIA. Windows.Google Scholar
Donnay, J. D. H., and Harker, D.. 1937. “A New Law of Crystal Morphology Extending the Law of Bravais.” American Mineralogist 22, 446447.Google Scholar
Dovesi, R., Erba, A., Orlando, R., Zicovich-Wilson, C. M., Civalleri, B., Maschio, L., Rérat, M., Casassa, S., Baima, J., Salustro, S., and Kirtman, B.. 2018. “Quantum-Mechanical Condensed Matter Simulations with CRYSTAL.” Wiley Interdisciplinary Reviews: Computational Molecular Science 8, e1360.Google Scholar
European Medicines Agency. 2018. “Braftovi.” Assessment Report EMA/CHMP/554696/2018.Google Scholar
Friedel, G. 1907. “Etudes sur la loi de Bravais.” Bulletin de Minéralogie 30, 326455.Google Scholar
Gates-Rector, S., and Blanton, T. N.. 2019. “The Powder Diffraction File: A Quality Materials Characterization Database.” Powder Diffraction 39, 352360.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, 1068610696.CrossRefGoogle Scholar
Gbabode, G., Negrier, P., Mondieig, D., Calvo, E. M., Calvet, T., and Cuevas-Diarte, M. À.. 2007. “Structures of the High-Temperature Solid Phases of the Odd-Numbered Fatty Acids from Tridecanoic Acid to Tricosanoic Acid.” Chemistry - A European Journal, 31503159. doi:10.1002/chem.200600955.CrossRefGoogle ScholarPubMed
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, 171179.CrossRefGoogle ScholarPubMed
Hirshfeld, F. L. 1977. “Bonded-Atom Fragments for Describing Molecular Charge Densities.” Theoretica Chemica Acta 44, 129138.CrossRefGoogle Scholar
Huang, S., Jin, X., Liu, Z., Poon, D., Tellew, J., Wan, Y., Wang, X., and Xie, Y.. 2013. “Compounds and Compositions as Protein Kinase Inhibitors.” US Patent 8,501,758 B2.Google Scholar
Imran, M., Asdaq, S. M., Khan, S. A., Unnikrishnan Meenakshi, D., Alamri, A. S., Alsanie, W. F., Alhomrani, M., Mohzari, Y., Alrashed, A., AlMotairi, M., and Alkhaldi, E. H. 2021. “Innovations and Patent Trends in the Development of USFDA Approved Protein Kinase Inhibitors in the Last Two Decades.” Pharmaceuticals 14, 710.CrossRefGoogle ScholarPubMed
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, 269273.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.. 2021. “PubChem in 2021: New Data Content and Improved Web Interfaces.” Nucleic Acids Research, D1388D1395. doi:10.1093/nar/gkaa971.CrossRefGoogle ScholarPubMed
Koelblinger, P., Thuerigen, O., and Dummer, R.. 2018. “Development of Encorafenib for BRAF-Mutated Advanced Melanoma.” Current Opinion in Oncology 30, 125133.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, 427432.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, 226235.CrossRefGoogle ScholarPubMed
Materials Data Inc (MDI). 2022. JADE Pro. V. 8.6. Materials Data Inc.Google Scholar
Materials Design. 2016. MedeA. V. 2.20.4. Windows.Google Scholar
Peintinger, M. F., Oliveira, D. V., and Bredow, T.. 2013. “Consistent Gaussian Basis Sets of Triple-Zeta Valence with Polarization Quality for Solid-State Calculations.” Journal of Computational Chemistry 34, 451459.CrossRefGoogle ScholarPubMed
Spackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D., and Spackman, M. A.. 2021. “Crystalexplorer: A Program for Hirshfeld Surface Analysis, Visualization and Quantitative Analysis of Molecular Crystals.” Journal of Applied Crystallography, 10061011. doi:10.1107/S1600576721002910.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, 882886.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: 544549.CrossRefGoogle 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, 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. V. 1.4.5. Wavefunction Inc, Irvine, CA.Google Scholar
Figure 0

Figure 1. The 2D molecular structure of encorafenib.

Figure 1

Figure 2. The synchrotron X-ray powder diffraction pattern of encorafenib indicates the two unindexed low-angle peaks. Note that the y-axis is a square root scale. Image generated using JADE Pro (MDI, 2022).

Figure 2

Figure 3. The Rietveld plot for the refinement of encorafenib. The blue crosses represent the observed data points, and the green line is the calculated pattern. The cyan curve is the normalized error plot, and the red line is the background curve. The vertical scale has been multiplied by a factor of 2× for 2θ > 2.0°, 20× for 2θ > 7.60°, and 40× for 2θ > 15.3°.

Figure 3

Figure 4. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of encorafenib molecule 1. The rms Cartesian displacement is 0.477 Å. Image generated using Mercury (Macrae et al., 2020).

Figure 4

Figure 5. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of encorafenib molecule 2. The rms Cartesian displacement is 0.622 Å. Image generated using Mercury (Macrae et al., 2020).

Figure 5

Figure 6. The asymmetric unit of encorafenib, with the atom numbering. The atoms are represented by 50% probability spheroids. Image generated using Mercury (Macrae et al., 2020).

Figure 6

Figure 7. The crystal structure of encorafenib, viewed down the b-axis. Image generated using Diamond (Crystal Impact, 2022).

Figure 7

Figure 8. Comparison of the molecular structures of molecules 1 (purple) and 2 (green) in the crystal structure of encorafenib.

Figure 8

Table I. Hydrogen bonds (CRYSTAL17) in encorafenib.

Figure 9

Figure 9. (a) The Hirshfeld surface of molecule 1 of encorafenib. 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 (Spackman et al., 2021). (b) The Hirshfeld surface of molecule 2 of encorafenib. 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 (Spackman et al., 2021).