Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-27T17:06:30.130Z Has data issue: false hasContentIssue false

Crystal structure of ractopamine hydrochloride, C18H24NO3Cl

Published online by Cambridge University Press:  29 February 2024

Colin W. Scherry
Affiliation:
North Central College, 131 S. Loomis St., Naperville, IL 60540, USA
Nicholas C. Boaz
Affiliation:
North Central College, 131 S. Loomis St., Naperville, IL 60540, USA
James A. Kaduk*
Affiliation:
North Central College, 131 S. Loomis St., Naperville, IL 60540, USA Illinois Institute of Technology, 3101 S. Dearborn St., Chicago IL 60616, USA
Anja 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 ractopamine hydrochloride has been solved and refined using synchrotron X-ray powder diffraction data, and optimized using density functional theory techniques. Ractopamine hydrochloride crystallizes in space group Pbca (#61) with a = 38.5871(49), b = 10.7691(3), c = 8.4003(2) Å, V = 3490.75(41) Å3, and Z = 8. The ractopamine cation contains two chiral centers, and the sample consists of a mixture of the S,S/R,R/S,R and R,S forms. Models for the two diastereomers S,S and S,R were refined, and yielded equivalent residuals, but the S,R form is significantly lower in energy. The crystal structure consists of layers of molecules parallel to the bc-plane. In each structure one of the H atoms on the protonated N atom acts as a donor in a strong discrete N–H⋯Cl hydrogen bond. Hydroxyl groups act as donors in O–H⋯Cl and O–H⋯O hydrogen bonds. Both the classical and C–H⋯Cl and C–H⋯O hydrogen bonds differ between the forms, helping to explain the large microstrain observed for the sample. 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), 2024. Published by Cambridge University Press on behalf of International Centre for Diffraction Data

I. INTRODUCTION

Ractopamine is an animal feed additive used to promote leanness and increase food conversion efficiency in farmed animals. It is the active ingredient in products marketed in the US as Paylean for swine, Optaflexx for cattle, and Topmax for turkeys. It was developed by Elanco Animal Health, a division of Eli Lilly and Company. As of 2014, the use of ractopamine was banned in 160 countries (Pacelle, Reference Pacelle2014), including the European Union, China, and Russia, while 27 other countries, such as Japan, the United States, South Korea, and New Zealand have deemed meat from livestock fed with ractopamine to be safe for human consumption. The systematic name (CAS Registry Number 90274-24-1) is 4-(1-hydroxy-2-((4-(4-hydroxyphenyl)butan-2-yl)amino)ethyl)phenol hydrochloride. A two-dimensional molecular diagram of ractopamine as downloaded from PubChem (Kim et al., Reference Kim, Chen, Cheng, Gindulyte, He, He, Li, Shoemaker, Thiessen, Yu, Zaslavsky, Zhang and Bolton2023) is shown in Figure 1.

Figure 1. The structure of the neutral ractopamine molecule, as downloaded from PubChem (Kim et al., Reference Kim, Chen, Cheng, Gindulyte, He, He, Li, Shoemaker, Thiessen, Yu, Zaslavsky, Zhang and Bolton2023), showing the S,S configuration. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Ractopamine has two chiral centers, one the so-called “OH-site” and the other the “Me-site” (Gunnar et al., Reference Gunnar, Reuter, Meier and Gogritchiani2009). Commercial ractopamine hydrochloride is a mixture of all four isomers. The R,R (OH,Me) isomer is known to be the most potent (Ricke et al., Reference Ricke, Smith, Feil, Larsen and Caton1999; Mills et al., Reference Mills, Kissel, Bidwell and Smith2003a, Reference Mills, Spurlock and Smith2003b). The hydrochloride salt of the R,R isomer is known as butopamine hydrochloride (Gunnar et al., Reference Gunnar, Reuter, Meier and Gogritchiani2009). R,R-ractopamine is not chirally stable, but epimerizes over a period of days. In contrast, R,R/S,R-ractopamine is an isomerically stable mixture (Gunnar et al., Reference Gunnar, Reuter, Meier and Gogritchiani2009). We are unaware of any published X-ray powder diffraction data for ractopamine hydrochloride.

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

Ractopamine hydrochloride was a commercial reagent, purchased from USP (Batch R066T0), 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 a beam line 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° to50° 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 synchrotron diffraction peaks are observed to be broad, and do not extend to high angles (Figure 2). After a number of unsuccessful attempts using several programs, the pattern was indexed using DICVOL06 (Louër and Boultif, Reference Louër and Boultif2007) as incorporated into EXPO2014 (Altomare et al., Reference Altomare, Cuocci, Giacovazzo, Moliterni, Rizzi, Corriero and Falcicchio2013) on a high-quality primitive orthorhombic unit cell with a = 38.49578, b = 10.75221, c = 8.39423 Å, V = 3474.49 Å3, and Z = 8. A reduced cell search in the Cambridge Structural Database (Groom et al., Reference Groom, Bruno, Lightfoot and Ward2016) with the chemistry C, H, Cl, N, and O only yielded six hits, but no ractopamine derivatives. Several programs indicated different space groups, but both JANA2006 (Petricek et al., Reference Petricek, Dusek and Palatinus2014) and DASH (David et al., Reference David, Shankland, van de Streek, Pidcock, Motherwell and Cole2006) suggested that Pbca was most probable. This suggestion was confirmed by a successful solution and refinement of the structure.

Figure 2. The synchrotron powder pattern of ractopamine hydrochloride, measured at 11-BM at APS using a wavelength of 0.458208 Å. Image generated using JADE Pro (MDI, 2023).

The ractopamine molecule (S,S) was downloaded from PubChem (Kim et al., Reference Kim, Chen, Cheng, Gindulyte, He, He, Li, Shoemaker, Thiessen, Yu, Zaslavsky, Zhang and Bolton2023) as Conformer3D_CID_56052.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), and to a Fenske-Hall Z-matrix using OpenBabel (O'Boyle et al., Reference O'Boyle, Banck, James, Morley, Vandermeersch and Hutchison2011). The crystal structure was solved using Monte Carlo simulated annealing techniques as implemented in DASH (David et al., Reference David, Shankland, van de Streek, Pidcock, Motherwell and Cole2006), with the S,S cation and a Cl atom as fragments.

The neutral molecule was protonated at the N atom, and converted to the S,R configuration using Materials Studio (Dassault Systèmes, 2022). The structure was solved independently using this cation and a Cl atom with EXPO2014 (Altomare et al., Reference Altomare, Cuocci, Giacovazzo, Moliterni, Rizzi, Corriero and Falcicchio2013).

NMR analysis of ractopamine HCl was performed using a 400-MHz Bruker Avance spectrometer equipped with a multinuclear probe with the pharmaceutical in d6DMSO (d6DMSO was stored over flame dried 3 Å molecular sieves). As shown in Figure 3, the proton NMR spectra indicated that the molecule was protonated at the secondary amine of ractopamine. Interestingly, one of the resonances assigned to the N–H proton is the result of two closely overlapping broad singlets, which are likely the result of the two sets of diastereomers present in the sample.

Figure 3. The 1H NMR spectrum of ractopamine HCl in d6DMSO.

Rietveld refinements of both structures (S,S and S,R) were carried out using GSAS-II (Toby and Von Dreele, Reference Toby and Von Dreele2013). A third refinement, with a 50/50 mixture of S,S and R,S molecules, yielded much higher residuals (Rwp ~0.14) and will not be discussed further. Only the 1.0°–25.0° (S,S) and 1.0°–22.0° (S,R) portions of the diffraction patterns were included in the refinements (d min = 1.058 and 1.201 Å, respectively). 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. Planar restraints were also applied to the phenyl rings. The restraints contributed 6.9 and 3.6% to the final χ 2. The hydrogen atoms were included in calculated positions, which were recalculated during the refinement using Materials Studio (Dassault Systèmes, 2022). 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. A fourth-order spherical harmonics model for preferred orientation was included in the refinements; the refined texture indices were 1.289 and 1.297. The peak profiles were described using the generalized microstrain model (Stephens, Reference Stephens1999). The background was modeled using a six-term shifted Chebyshev polynomial, and a peak at 6.50° 2θ to model the scattering from the Kapton capillary and any amorphous component.

The final refinements yielded the residuals reported in Table I. The largest errors in the final difference plots (Figures 4 and 5) are small, and reflect fitting ordered models to a disordered system.

TABLE I. Refinement residuals for ractopamine hydrochloride

Figure 4. The Rietveld plot for the refinement of S,S-ractopamine hydrochloride. 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 10× for 2θ > 9.0°.

Figure 5. The Rietveld plot for the refinement of S,R-ractopamine hydrochloride. 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 10× for 2θ > 9.0°.

The structures of S,S and S,R ractopamine hydrochloride were 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 calculations 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 ~35 (S,S) and 48 (S,R) days. Rietveld refinements using the fixed VASP-optimized structures yielded much higher residuals, with Rwp around 32%. Single-point density functional theory calculations (fixed experimental cell) and population analysis were carried out using CRYSTAL23 (Erba et al., Reference Erba, Desmaris, Casassa, Civalleri, Donà, Bush, Searle, Maschio, Daga, Cossard, Ribaldone, Ascrizzi, Marana, Flament and Kirtman2023). 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 Cl was that of Peintinger et al. (Reference Peintinger, Vilela Oliveira and Bredow2013). The calculations were run on a 3.5 GHz PC using eight k-points and the B3LYP functional, and took ~3.4 (S,S) and 3.5 (S,R) hr.

III. RESULTS AND DISCUSSION

Two different ordered models (S,S and S,R) for ractopamine hydrochloride have been refined, for a sample which consists of a mixture of S,S/R,R/S,R and R,S forms. Powder diffraction cannot distinguish the two racemic pairs, so refinement of two diastereomers was sufficient. In Pbca, each structural model contains both members of an enantiomeric pair. The asymmetric units (with atom numbering) are illustrated in Figures 6 and 7. The displacement coefficients in the centers of the molecules are large, reflecting the disorder.

Figure 6. The asymmetric unit of S,S-ractopamine hydrochloride, 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).

Figure 7. The asymmetric unit of S,R-ractopamine hydrochloride, 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).

Refinements of the two models yielded comparable residuals, but the S,R model is significantly lower in energy. The S,S model also resulted in an unreasonably short cation-anion distance. The S,R model is thus to be preferred for this disordered system.

The root-mean-square (rms) Cartesian displacement of the refined S,S and S,R molecules is 0.770 Å (Figure 8). The rms displacement for the VASP-optimized molecules is 1.079 Å (Figure 9). The overall shapes of the molecules are similar, making it non-unreasonable that they could be accommodated in the same lattice. The average microstrains for the two models are 17,479 (S,S) and 16,917 (S,R) ppm, reflecting the imperfect nature of the crystals.

Figure 8. Comparison of the as-refined S,S (green) and S,R (purple) ractopamine cations. The rms deviation is 0.770 Å native, and 0.421 Å inverted. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 9. Comparison of the VASP-optimized S,S (green) and S,R (purple) ractopamine cations. The rms deviation is 1.079 Å native, and 0.900 Å inverted. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

The rms displacement between the Rietveld-refined and VASP-optimized S,S cations is 0.637 Å (Figure 10), and the displacement between the S,R cations is 0.668 Å (Figure 11). The agreements are outside the normal range for correct structures (van de Streek and Neumann, Reference van de Streek and Neumann2014), but this is to be expected from refining ordered models of a disordered structure. Overlays of the refined and optimized structures of the S,S and S,R models are presented in Figures 12 and 13. A more-standard picture of the optimized S,R structure is included as Figure 14. The remaining discussion will concentrate on the VASP-optimized structures.

Figure 10. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of the S,S-ractopamine cation. The rms Cartesian displacement is 0.637 Å. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 11. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of the S,R-ractopamine cation. The rms Cartesian displacement is 0.668 Å. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 12. Overlap of the refined S,S, and S,R structures of ractopamine hydrochloride, viewed down the c-axis. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 13. Overlap of the VASP-optimized S,S, and S,R structures of ractopamine hydrochloride, viewed down the c-axis. Image generated using Mercury (Macrae et al., Reference Macrae, Sovago, Cottrell, Galek, McCabe, Pidcock, Platings, Shields, Stevens, Towler and Wood2020).

Figure 14. The crystal structure of ractopamine hydrochloride, viewed down the b-axis. Image generated using Diamond (Crystal Impact, 2022).

The crystal structure consists of layers of molecules parallel to the bc-plane (Figure 14). The mean planes of the phenyl rings in the S,S molecule are (5,2,7) and (7,8,7), while those in the S,R molecule are (1,1,−1) and (−5,−3,4); the orientations of the rings thus differ in the two structures. The Mercury Aromatics Analyser indicates only weak interactions (>4.9 Å) in the two structures. N–H⋯Cl hydrogen bonds link the cations and anions.

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). Only the C6–C5–C4 angle of 117.8° (average = 113.6(14)°; Z-score = 3.0) in the S,S molecule is flagged as unusual. Torsion angles in S,S involving rotation about the C4–N44 and C5–C6 bonds lie in minor populations of trimodal gauche/trans distributions, and are flagged as unusual. Torsion angles involving rotation about the C7–C10 bond are approximately 0/180°, on the tails of a broad distribution centered about 90°.

Quantum chemical geometry optimizations of the isolated cations (DFT/B3LYP/6-31G*/water) using Spartan ‘18 (Wavefunction, 2020) indicated that the S,R cation is 1.2 kcal/mol lower in energy than the S,S cation, but this difference lies within the expected error range of such calculations, so the two molecules should be considered equivalent in energy. The global minimum-energy conformations (MMFF force field) of both the S,S and S,R cations are much more compact (fold on themselves), showing that intermolecular interactions are important in determining the observed solid-state conformations.

Analysis of the contributions to the total crystal energy of the structure using the Forcite module of Materials Studio (Dassault Systèmes, 2022) suggests that bond, angle, and torsion distortion terms dominate the intramolecular deformation energy. The intermolecular energy is dominated by 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.

Hydrogen bonds are prominent in both the S,S and S,R structures (Tables II and III). As expected, in each structure one of the H atoms on the protonated N atom acts as a donor in a strong discrete N–H⋯Cl hydrogen bond. In the S,R structure this H also forms a weak intramolecular N-H⋯O hydrogen bond, and the other H atom on the N also forms an intramolecular N–H⋯O hydrogen bond. In each structure, hydroxyl groups act as donors in O–H⋯Cl and O–H⋯O hydrogen bonds. The energies of the O–H⋯O hydrogen bonds were calculated using the correlation of Rammohan and Kaduk (Reference Rammohan and Kaduk2018), the energies of the N–H⋯O hydrogen bonds were calculated using the correlation of Wheatley and Kaduk (Reference Wheatley and Kaduk2019), and the energies of the O–H⋯Cl hydrogen bonds were calculated using a correlation implicit in Kaduk (Reference Kaduk2002). In the S,S structure the classical hydrogen bonds link the molecules into chains parallel to the a-axis, while in the S,R structure, they link the molecules into corrugated chains parallel to the b-axis. There are several C–H⋯Cl and C–H⋯O hydrogen bonds, which vary between the two structures.

TABLE II. Hydrogen bonds (CRYSTAL23) in S,S-ractopamine hydrochloride

a Intramolecular charge.

TABLE III. Hydrogen bonds (CRYSTAL23) in S,R-ractopamine hydrochloride

a Intramolecular charge.

The volumes enclosed by the Hirshfeld surface of the ractopamine hydrochloride asymmetric unit (Figures 15 and 16, Hirshfeld, Reference Hirshfeld1977; Spackman et al., Reference Spackman, Turner, McKinnon, Wolff, Grimwood, Jayatilaka and Spackman2021) are 428.59 and 428.02 Å3, 98.1 and 98.2% of the unit cell volume. The packing density is thus fairly typical. The only significant close contacts (red in Figures 15 and 16) involve the hydrogen bonds. The volume/non-hydrogen atom is larger than normal, at 19.0 Å3, reflecting the presence of the large chloride anion.

Figure 15. The Hirshfeld surface of S,S-ractopamine hydrochloride. 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).

Figure 16. The Hirshfeld surface of S,R-ractopamine hydrochloride. 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 platy morphology for ractopamine hydrochloride, with {100} as the major faces. Afourth-order spherical harmonics model for preferred orientation was included in the refinements; the refined texture indices were 1.289 and 1.297, indicating that preferred orientation was significant in this rotated capillary specimen.

IV. DEPOSITED DATA

The powder pattern of ractopamine hydrochloride 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

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: 1231–35.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. V. 4.6.8. Crystal Impact - Dr. H. Putz & Dr. K. Brandenburg. Windows.Google Scholar
Dassault Systèmes. 2022. Materials Studio 2023. San Diego, CA, BIOVIA.Google Scholar
David, W. I. F., Shankland, K., van de Streek, J., Pidcock, E., Motherwell, W. D. S., and Cole, J. C.. 2006. “DASH: A Program for Crystal Structure Determination from Powder Diffraction Data.” Journal of Applied Crystallography 39: 910–5.CrossRefGoogle Scholar
Donnay, J. D. H., and Harker, D.. 1937. “A New Law of Crystal Morphology Extending the Law of Bravais.” American Mineralogist 22: 446–66.Google Scholar
Erba, A., Desmaris, J. K., Casassa, S., Civalleri, B., Donà, L., Bush, I. J., Searle, B., Maschio, L., Daga, L.-E., Cossard, A., Ribaldone, C., Ascrizzi, E., Marana, N. L., Flament, J.-P., and Kirtman, B.. 2023. “CRYSTAL23: A Program for Computational Solid State Physics and Chemistry.” Journal of Chemical Theory and Computation 19: 6891–932. doi:10.1021/acs.jctc.2c00958.CrossRefGoogle ScholarPubMed
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. N.. 2019. “The Powder Diffraction File: A Quality Materials Characterization Database.” Powder Diffraction 39: 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
Gunnar, A. K., Reuter, K., Meier, V., & Gogritchiani, E.. 2009. Use of RR/SR-Ractopamine. U.S. Patent Application 2009/0143480 A1.Google Scholar
Hirshfeld, F. L. 1977. “Bonded-Atom Fragments for Describing Molecular Charge Densities.” Theoretica Chemica Acta 44: 129–38.CrossRefGoogle Scholar
Kaduk, J. A. 2002. “Use of the Inorganic Crystal Structure Database as a Problem Solving Tool.” Acta Crystallographica B 58: 370–9.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: 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.. 2023. “Pubchem 2023 Update.” Nucleic Acids Research 51 (D1): D1373–80. doi:10.1093/nar/gkac956.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: 427–32.CrossRefGoogle ScholarPubMed
Louër, D., and Boultif, A.. 2007. “Powder Pattern Indexing and the Dichotomy Algorithm.” Zeitschrift fur Kristallographie Supplements 26: 191.CrossRefGoogle Scholar
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. 2023. JADE Pro version 8.8. Livermore, CA, Materials Data.Google Scholar
Mills, S. E., Kissel, J., Bidwell, C. A., and Smith, D. J.. 2003a. “Stereoselectivity of Porcine β-Adrenergic for Ractopamine Stereoisomers.” Journal of Animal Science 81: 122–9.CrossRefGoogle ScholarPubMed
Mills, S. E., Spurlock, M. E., and Smith, D. J.. 2003b. “β-Adrenergic Receptor Subtypes that Mediate Ractopamine Stimulation of Lypolysis.” Journal of Animal Science 81: 662–8.CrossRefGoogle Scholar
O'Boyle, N. M., Banck, M., James, C. A., Morley, C., Vandermeersch, T., and Hutchison, G. R.. 2011. “Open Babel: An Open Chemical Toolbox.” Journal of Chemical Informatics 3: 33. doi:10.1186/1758-2946-3-33.Google ScholarPubMed
Pacelle, W. 2014. “Banned in 160 Nations, Why is Ractopamine in U.S. Pork? (Op-Ed).” Live Science. Expert Voices: Op-Ed & Insights. https://www.livescience.com/47032-time-for-us-to-ban-ractopamine.htmlGoogle 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
Petricek, V., Dusek, M., and Palatinus, L.. 2014. “Crystallographic Computing System JANA2006: General features.” Zeitschrift für Kristallographie-Crystalline Materials 229: 345–52.CrossRefGoogle Scholar
Rammohan, A., and Kaduk, J. A.. 2018. “Crystal Structures of Alkali Metal (Group 1) Citrate Salts.” Acta Crystallographica Section B: Crystal Engineering and Materials 74: 239–52. doi:10.1107/S2052520618002330.CrossRefGoogle ScholarPubMed
Ricke, E. A., Smith, D. J., Feil, V. J., Larsen, G. L., and Caton, J. S.. 1999. “Effects of Ractopamine HCl Stereoisomers on Groth, Nitrogen Retention, and Carcass Composition in Rats.” Journal of Animal Science 77: 701–7.CrossRefGoogle Scholar
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 54: 1006–11. doi:10.1107/S1600576721002910.CrossRefGoogle ScholarPubMed
Stephens, P. W. 1999. “Phenomenological Model of Anisotropic Peak Broadening in Powder Diffraction.” Journal of Applied Crystallography 32: 281–9.CrossRefGoogle Scholar
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
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: 1020–32.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. Wavefunction Inc., 18401 Von Karman Ave., Suite 370, Irvine CA 96212.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 structure of the neutral ractopamine molecule, as downloaded from PubChem (Kim et al., 2023), showing the S,S configuration. Image generated using Mercury (Macrae et al., 2020).

Figure 1

Figure 2. The synchrotron powder pattern of ractopamine hydrochloride, measured at 11-BM at APS using a wavelength of 0.458208 Å. Image generated using JADE Pro (MDI, 2023).

Figure 2

Figure 3. The 1H NMR spectrum of ractopamine HCl in d6DMSO.

Figure 3

TABLE I. Refinement residuals for ractopamine hydrochloride

Figure 4

Figure 4. The Rietveld plot for the refinement of S,S-ractopamine hydrochloride. 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 10× for 2θ > 9.0°.

Figure 5

Figure 5. The Rietveld plot for the refinement of S,R-ractopamine hydrochloride. 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 10× for 2θ > 9.0°.

Figure 6

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

Figure 7

Figure 7. The asymmetric unit of S,R-ractopamine hydrochloride, with the atom numbering. The atoms are represented by 50% probability spheroids. Image generated using Mercury (Macrae et al., 2020).

Figure 8

Figure 8. Comparison of the as-refined S,S (green) and S,R (purple) ractopamine cations. The rms deviation is 0.770 Å native, and 0.421 Å inverted. Image generated using Mercury (Macrae et al., 2020).

Figure 9

Figure 9. Comparison of the VASP-optimized S,S (green) and S,R (purple) ractopamine cations. The rms deviation is 1.079 Å native, and 0.900 Å inverted. Image generated using Mercury (Macrae et al., 2020).

Figure 10

Figure 10. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of the S,S-ractopamine cation. The rms Cartesian displacement is 0.637 Å. Image generated using Mercury (Macrae et al., 2020).

Figure 11

Figure 11. Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of the S,R-ractopamine cation. The rms Cartesian displacement is 0.668 Å. Image generated using Mercury (Macrae et al., 2020).

Figure 12

Figure 12. Overlap of the refined S,S, and S,R structures of ractopamine hydrochloride, viewed down the c-axis. Image generated using Mercury (Macrae et al., 2020).

Figure 13

Figure 13. Overlap of the VASP-optimized S,S, and S,R structures of ractopamine hydrochloride, viewed down the c-axis. Image generated using Mercury (Macrae et al., 2020).

Figure 14

Figure 14. The crystal structure of ractopamine hydrochloride, viewed down the b-axis. Image generated using Diamond (Crystal Impact, 2022).

Figure 15

TABLE II. Hydrogen bonds (CRYSTAL23) in S,S-ractopamine hydrochloride

Figure 16

TABLE III. Hydrogen bonds (CRYSTAL23) in S,R-ractopamine hydrochloride

Figure 17

Figure 15. The Hirshfeld surface of S,S-ractopamine hydrochloride. 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).

Figure 18

Figure 16. The Hirshfeld surface of S,R-ractopamine hydrochloride. 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).