Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-25T16:22:24.704Z Has data issue: false hasContentIssue false

Powder X-ray diffraction data for dimethylarsinic acid, (CH3)2AsO(OH)

Published online by Cambridge University Press:  29 April 2021

Joel W. Reid*
Affiliation:
Canadian Light Source, 44 Innovation Boulevard, Saskatoon, SK, CanadaS7N 2V3
*
a)Author to whom correspondence should be addressed. Electronic mail: joel.reid@lightsource.ca

Abstract

Synchrotron powder diffraction data is presented for the monoclinic polymorph of dimethylarsinic acid, (CH3)2AsO(OH) (DMAV). Rietveld refinement with GSASII yielded lattice parameters of a = 15.9264(15) Å, b = 6.53999(8) Å, c = 11.3401(9) Å, and β = 125.8546(17)° (Z = 8, space group C2/c). The Rietveld-refined structure was compared with both a density functional theory (DFT)-optimized structure and the published, low-temperature single-crystal structure, and all three structures exhibited excellent agreement. The triclinic polymorph of DMAV was also DFT optimized with CRYSTAL17 to determine the positions of the hydrogen atoms. Monoclinic DMAV forms zigzag chains parallel to the b-axis with adjacent DMAV molecules connected by an O–H⋯O bond, whereas triclinic DMAV forms dimers connected by two O–H⋯O bonds.

Type
New Diffraction Data
Copyright
Copyright © The Author(s), 2021. Published by Cambridge University Press on behalf of International Centre for Diffraction Data

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adamescu, A., Hamilton, I. P., and Al-Abadleh, H. A. (2011). “Thermodynamics of dimethylarsinic acid and arsenate interactions with hydrated iron-(oxyhydr)oxide clusters: DFT calculations,” Environ. Sci. Technol. 45, 1043810444.CrossRefGoogle ScholarPubMed
Agency for Toxic Substances and Disease Registry, ATSDR (2020). Priority list of hazardous substances. Available at: http://www.atsdr.cdc.gov/SPL/index.html.Google Scholar
Becke, A. D. (1993). “Density-functional thermochemistry. III. The role of exact exchange,” J. Chem. Phys. 98, 56485652.CrossRefGoogle Scholar
Betz, R., McCleland, C., and Marchand, H. (2011). “The monoclinic polymorph of dimethyl-arsinic acid,” Acta Crystallogr. E E67, m1013.CrossRefGoogle Scholar
Borak, J. and Hosgood, H. D. (2007). “Seafood arsenic: implications for human risk assessment,” Regul. Toxicol. Pharmacol. 47, 204212.CrossRefGoogle ScholarPubMed
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,” J. Chem. Inf. Comput. Sci. 44, 21332144.CrossRefGoogle ScholarPubMed
Capitani, E. M. D. (2011). “Arsenic toxicology - a review,” in Arsenic: Natural and Anthropogenic, edited by Deschamps, F. and Matschullat, J. (CRC Press, London), pp. 2737.CrossRefGoogle Scholar
Dovesi, R., Erba, A., Orlando, R., Zicovich-Wilson, C. M., Civalleri, B., Maschio, L., Rerat, M., Casassa, S., Baima, J., Salustro, S., and Kirtman, B. (2018). “Quantum-mechanical condensed matter simulations with CRYSTAL,” WIREs Comput. Mol. Sci. 8, e1360.CrossRefGoogle Scholar
Fodje, M., Grochulski, P., Janzen, K., Labiuk, S., Gorin, J., and Berg, R. (2014). “08B1-1: an automated beamline for macromolecular crystallography experiments at the Canadian Light Source,” J. Synchrotron Radiat. 21, 633637.CrossRefGoogle Scholar
Gates-Rector, S. D. and Blanton, T. N. (2019). “The Powder Diffraction File: a quality materials characterization database,” Powd. Diffr. 34, 352360.CrossRefGoogle Scholar
Gatti, C., Saunders, V. R., and Roetti, C. (1994). “Crystal-field effects on the topological properties of the electron-density inmolecular crystals - the case of urea,” J. Chem. Phys 101, 1068610696.CrossRefGoogle Scholar
Grimme, S., Antony, J., Ehrlich, S., and Krieg, H. (2010). “A consistent and accurate ab initio parameterization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu,” J. Chem. Phys. 132, 154104.CrossRefGoogle Scholar
Hanwell, M. D., Curtis, D. E., Lonie, D. C., Vandermeersch, T., Zurek, E., and Hutchison, G. R. (2012). “Avogadro: an advanced semantic chemical editor, visualization, and analysis platform,” J. Cheminform. 4, 17.CrossRefGoogle ScholarPubMed
Hughes, M. F. and Kenyon, E. M. (1998). “Dose-dependent effects on the disposition of monomethylarsonic acid and dimethylarsinic acid in the mouse after intravenous administration,” J. Toxicol. Environ. Health A 53, 95112.Google ScholarPubMed
Kraus, W. and Nolze, G. (1996). “POWDER CELL – a program for the representation and manipulation of crystal structures and calculation of the resulting X-ray powder patterns,” J. Appl. Crystallogr. 29, 301303.CrossRefGoogle Scholar
Limmer, M. A., Wise, P., Dykes, G. E., and Seyfferth, A. L. (2018). “Silicon decreases dimethylarsinic acid concentration in rice grain and mitigates straighthead disorder,” Environ. Sci. Technol. 52, 48094816.CrossRefGoogle ScholarPubMed
Macrae, C. F., Bruno, I. J., Chisholm, J. A., Edgington, P. R., McCabe, P., Pidcock, E., Rodriquez-Monge, L., Taylor, R., van de Streek, J., and Wood, P. A. (2008). “Mercury CSD 2.0 – new features for the visualization and investigation of crystal structures,” J. Appl. Crystallogr. 41, 466470.CrossRefGoogle Scholar
Momma, K. and Izumi, F. (2011). “VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data,” J. Appl. Crystallogr. 44, 12721276.CrossRefGoogle Scholar
Nagar, R., Sarkar, D., Makris, K., and Datta, R. (2014). “Arsenic bioaccessibility and speciation in the soils amended with organoarsenicals and drinking-water treatment residuals based on a long-term greenhouse study,” J. Hydrol. 518, 477485.CrossRefGoogle 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,” J. Comput. Chem. 34, 451459.CrossRefGoogle ScholarPubMed
Quazi, S., Sarkar, D., and Datta, R. (2013). “Human health risk from arsenical pesticide contaminated soils: a long-term greenhouse study,” J. Hazard. Mater. 262, 10311038.CrossRefGoogle ScholarPubMed
Tchounwou, P. B., Patlolla, A. K., and Centeno, J. A. (2003). “Carcinogenic and systemic health effects associated with arsenic exposure—a critical review,” Toxicol. Pathol. 31, 575588.Google ScholarPubMed
Thompson, P., Cox, D. E., and Hastings, J. B. (1987). “Rietveld refinement of Debye-Scherrer synchrotron X-ray data from A12O3,” J. Appl. Crystallogr. 20, 7983.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,” J. Appl. Crystallogr. 46, 544549.10.1107/S0021889813003531CrossRefGoogle Scholar
Tofan-Lazar, J. and Al-Abadleh, H. A. (2012). “ATR-FTIR studies on the adsorption/desorption kinetics of dimethylarsinic acid on iron−(oxyhydr)oxides,” J. Phys. Chem. A 116, 15961604.CrossRefGoogle ScholarPubMed
Trotter, J. and Zobel, T. (1965). “Stereochemistry of arsenic. Part XVI. Cacodylic acid,” J. Chem. Soc., 44664471.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 Crystallogr. B. 70, 10201032.CrossRefGoogle Scholar