Hostname: page-component-848d4c4894-v5vhk Total loading time: 0 Render date: 2024-06-30T09:55:02.486Z Has data issue: false hasContentIssue false

Synchrotron powder diffraction, X-ray absorption and 1H nuclear magnetic resonance data for hypoxanthine, C5H4N4O

Published online by Cambridge University Press:  13 August 2015

Joel Reid*
Affiliation:
Canadian Light Source, 44 Innovation Boulevard, Saskatoon, SK, Canada S7N 2V3
Toby Bond
Affiliation:
Canadian Light Source, 44 Innovation Boulevard, Saskatoon, SK, Canada S7N 2V3
Shiliang Wang
Affiliation:
Defence Research & Development Canada Suffield, Medicine Hat, AB, Canada T1A 8K6
Jigang Zhou
Affiliation:
Canadian Light Source, 44 Innovation Boulevard, Saskatoon, SK, Canada S7N 2V3
Anguang Hu
Affiliation:
Defence Research & Development Canada Suffield, Medicine Hat, AB, Canada T1A 8K6
*
a)Author to whom correspondence should be addressed. Electronic mails: joel.reid@lightsource.ca, joelwreid@gmail.com

Abstract

Synchrotron powder X-ray diffraction, X-ray absorption spectroscopy (XAS), and proton nuclear magnetic resonance (1H-NMR) data have been used to examine the structure of hypoxanthine, 1,7-dihydro-6H-purin-6-one (C5H4N4O), a purine base that participates in numerous metabolic processes. XAS and 1H-NMR spectroscopy were used to determine that hypoxanthine was present in its keto form (both in solid state and dissolved in an organic solvent). Rigid body refinement was performed with the Rietveld software package GSAS yielding triclinic lattice parameters of a = 7.1179 (2) Å, b = 9.7830 (3) Å, c = 10.4009 (3) Å, α = 58.876 (1)°, β = 67.609 (1)°, and γ = 71.937 (2)° (C5H4N4O, Z = 4, space group P$\bar 1$).

Type
New Diffraction Data
Copyright
Copyright © International Centre for Diffraction Data 2015 

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

Allen, F. H. (2002). “The Cambridge structural database: a quarter of a million crystal structures and rising,” Acta Crystallogr. B 58, 380388.Google Scholar
Bartl, T., Zacharová, Z., Sečkářová, P., Kolehmainen, E. and Marek, R. (2009). “NMR quantification of tautomeric populations in biogenic purine bases,” Eur. J. Org. Chem. 9, 13771383.Google Scholar
Batey, R. T., Gilbert, S. D. and Montange, R. K. (2004). “Structure of a natural guanine-responsive riboswitch complexed with the metabolite hypoxanthine,” Nature 432, 411415.Google Scholar
Becke, A. D. (1988). “Density-functional exchange-energy approximation with correct asymptotic behavior,” Phys. Rev. A 38(6), 30983100.Google Scholar
Chenon, M. T., Pugmire, R. J., Grant, D. M., Panzica, R. P. and Townsend, L. B. (1975). “Carbon-13 magnetic resonance. XXVI. Quantitative determination of the tautomeric populations of certain purines,” J. Am. Chem. Soc. 97, 46364642.CrossRefGoogle ScholarPubMed
Deng, H., Cahill, S. M., Abad, J. L., Lewandowicz, A., Callender, R. H., Schramm, V. and Jones, R. A. (2004). “Active site contacts in the purine nucleoside phosphorylase-hypoxanthine complex by NMR and ab initio calculations,” Biochemistry 43, 1596615974.Google Scholar
Dinnebier, R. E. (1999). “Rigid bodies in powder diffraction: a practical guide,” Powder Diffr. 14, 8492.Google Scholar
Gallant, B. M., Mitchell, R. R., Kwabi, D. G., Zhou, J., Zuin, L., Thompson, C. V. and Shao-Horn, Y. (2012). “Chemical and morphological changes of Li–O2 battery electrodes upon cycling,” J. Phys. Chem. C 116, 2080020805.Google Scholar
ICDD (2013), PDF-4+ 2013 (Database) edited by Dr. Kabekkodu, Soorya (International Centre for Diffraction Data, Newtown Square, PA, USA).Google Scholar
Krylov, A. I. and Gill, P. M. W. (2013). “Q-Chem: an engine for innovation,” WIREs Comput. Mol. Sci. 3, 317326.Google Scholar
Lake, C. H. and Toby, B. H. (2011). “Rigid body refinements in GSAS/EXPGUI,” Powder Diffr. 26, S13S21.CrossRefGoogle Scholar
Larson, A. C. and Von Dreele, R. B. (2004). General Structure Analysis System (GSAS) (Report No. LAUR 86-748), Los Alamos, NM: Los Alamos National Laboratory.Google Scholar
Lee, C., Yang, W. and Parr, R. G. (1988). “Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density,” Phys. Rev. B 37(2), 785789.Google Scholar
Marat, K. (2014). Spinworks: NMR Processing Software (Version 4.0.5.0). Available from http://home.cc.umanitoba.ca/~wolowiec/spinworks/ Google Scholar
Pang, B., McFaline, J. L., Burgis, N. E., Dong, M., Taghizadeh, K., Sullivan, M. R., Elmquist, C. E., Cunningham, R. P. and Dedon, P. C. (2012). “Defects in purine nucleotide metabolism lead to substantial incorporation of xanthine and hypoxanthine into DNA and RNA,” Proc. Nat. Acad. Sci. USA 109, 23192324.Google Scholar
Parker, R., Snedden, W. and Watts, R. W. E. (1969). “The mass-spectrometric identification of hypoxanthine and xanthine (‘oxypurine’) in skeletal muscle from two patients with congenital xanthine oxidase deficiency (xanthinuria),” Biochem. J. 115, 103108.CrossRefGoogle Scholar
Puig, J. G., Mateos, F. A., Jimenez, M. L. and Ramos, T. H. (1988). “Renal excretion of hypoxanthine and xanthine in primary gout,” Am. J. Med. 85, 533537.Google Scholar
Reid, J. W., Kaduk, J. A., Garimella, S. V. and Tse, J. S. (2015). “Rietveld Refinement using synchrotron powder diffraction data for curcumin, C21H20O6, and comparison with density functional theory,” Powder Diffr. 30, 6775.CrossRefGoogle Scholar
Saladino, R., Crestini, C., Costanzo, G., Negri, R., and Di Mauro, E. (2001). “A possible prebiotic synthesis of purine, adenine, cytosine, and 4(3H)-pyrimidinone from formamide: implications for the origin of life,” Bioorg. Med. Chem. 9, 12491253.CrossRefGoogle ScholarPubMed
Samuel, N. T., Lee, C. Y., Gamble, L. J., Fischer, D. A. and Castner, D. G. (2006). “NEXAFS characterization of DNA components and molecular-orientation of surface-bound DNA oligomers,” J. Electron Spectrosc. Rel. Phenom. 152, 134142.Google Scholar
Sarkar, S. D. and Nahar, L. (2007). Chemistry for Pharmacy Students (Wiley, New York).Google Scholar
Schmalle, H. W., Hänggi, G. and Dubler, E. (1988). “Structure of hypoxanthine,” Acta Crystallogr. C 44, 732736.Google Scholar
Toby, B. H. (2001). “EXPGUI, a graphical user interface for GSAS,” J. Appl. Crystallogr. 34, 210213.Google 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.CrossRefGoogle Scholar
Yang, R. Q. and Xie, Y. R. (2007). “A monoclinic polymorph of hypoxanthine,” Acta Crystallogr. E 63, o3309.Google Scholar
Zhou, J., Hu, Y., Li, X., Wang, C. and Zuin, L. (2014). “Chemical bonding in amorphous Si-coated carbon nanotubes as anodes for Li ion batteries: a XANES study,” RSC Adv. 4, 2022620229.CrossRefGoogle Scholar
Zubavichus, Y., Shaporenko, A., Korolkov, V., Grunze, M. and Zharnikov, M. (2008). “X-ray absorption spectroscopy of the nucleotide bases at the carbon, nitrogen, and oxygen K-edges,” J. Phys. Chem. B 112, 1371113716.Google Scholar