Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-25T15:19:29.933Z Has data issue: false hasContentIssue false

Crystal structure of antimony oxalate hydroxide, Sb(C2O4)OH

Published online by Cambridge University Press:  29 February 2012

James A. Kaduk*
Affiliation:
INEOS Technologies, P.O. Box 3011, MC F-9, Naperville, Illinois 60566
Mark A. Toft
Affiliation:
INEOS Technologies, P.O. Box 3011, MC F-9, Naperville, Illinois 60566
Joseph T. Golab
Affiliation:
INEOS Technologies, P.O. Box 3011, MC F-9, Naperville, Illinois 60566
*

Abstract

The crystal structure of Sb(C2O4)OH has been solved by charge flipping in combination with difference Fourier techniques using laboratory X-ray powder data exhibiting significant preferred orientation and refined by the Rietveld method. The compound crystallizes in Pnma with a=5.827 13(3), b=11.294 48 (10), c=6.313 77(3) Å, V=415.537(5) Å3, and Z=4. The crystal structure contains pentagonal pyramidal Sb3+ cations, which are bridged by hydroxyl groups to form zigzag chains along the a axis. Each oxalate anion chelates to two Sb in approximately the ab plane, linking the chains into a three-dimensional framework. The H of the hydroxyl group is probably disordered in order to form stronger more-linear hydrogen bonds. The highest energy occupied molecular orbitals are the Sb3+ lone pairs. The structure is chemically reasonable compared to other antimony oxalates and to Bi(C2O4)OH.

Type
Technical Articles
Copyright
Copyright © Cambridge University Press 2010

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

Ambe, S. (1975). “Chemical properties of antimony(III) oxalate hydroxide,” J. Inorg. Nucl. Chem. JINCAO 37, 2023.10.1016/0022-1902(75)80948-1CrossRefGoogle Scholar
Brese, N. E. and O’Keefe, M. (1991). “Bond-valence parameters for solids,” Acta Crystallogr., Sect. B: Struct. Sci. ASBSDK B47, 192197 .10.1107/S0108768190011041CrossRefGoogle Scholar
Brown, I. D. (2002). The Chemical Bond in Inorganic Chemistry: The Bond Valence Model (Oxford University Press, Oxford).Google Scholar
Bruno, I. J., Cole, J. C., Kessler, M., Luo, Jie, 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. JCISD8 44, 21332144 .10.1021/ci049780bCrossRefGoogle ScholarPubMed
Clark, S. J., Segall, M. D., Pickard, C. J., Hasnip, P. J., Probert, M. J., Refson, K., and Payne, M. C. (2005). “First principles methods using CASTEP,” Z. Kristallogr. ZEKRDZ 220, 567570 .10.1524/zkri.220.5.567.65075CrossRefGoogle Scholar
Coudreau-Ducourant, D., Ducourant, B., Fourcade, R., and Mascherpa, G. (1981). “New complex of antimony oxide fluoride and oxalate: Crystal structure of (NH4)4H2(C2O4)3(SbOF)2·2H2O,” Z. Anorg. Allg. Chem. ZAACAB 476, 229236 (CSD Refcode OXFLSB).10.1002/zaac.19814760527CrossRefGoogle Scholar
Davidovich, R. L., Zemnukhova, L. A., Udovenko, A. A., and Sigula, N. I. (1983). “Synthesis and structure of rubidium oxalatofluorodiantimonates(III),” Koord. Khim. KOKHDC 9, 787792 (CSD Refcodes CANRIS, CANROY, and CANRUE).Google Scholar
Donnay, J. D. H. and Harker, D. (1937). “A new law of crystal morphology extending the law of Bravais,” Am. Mineral. AMMIAY 22, 446467.Google Scholar
Escande, P., Tichit, D., Ducourant, B., Fourcade, R., and Mascherpa, G. (1978a). “Interaction between the lone electronic pair and π bond in a nonsymmetric system: Crystal structure of sodium oxalate-antimony trifluoride (Na2C2O4(SbF3)2),” Ann. Chim. (Paris) ANCPAC 3, 117124 (CSD Refcode SBFOXS).Google Scholar
Escande, P., Tichit, D., Ducourant, M. B., Fourcade, R., and Mascherpa, G. (1978b). “Interaction paire electronique libre-liaison pi dans un systeme non symetrique: Structure cristalline de Na2C2O4(SbF3)2,” Ann. Chim. (Paris) ANCPAC 3, 124124 (ICSD collection code 200225).Google Scholar
Faber, J. and Fawcett, T. (2002). “The Powder Diffraction File: Present and future,” Acta Crystallogr., Sect. B: Struct. Sci. ASBSDK 58, 325332 .10.1107/S0108768102003312CrossRefGoogle ScholarPubMed
Hellenbrandt, M. (2004). “The Inorganic Crystal Structure Database (ICSD)—Present and future,” Crystallogr. Rev. CRRVEN 10, 1722 .10.1080/08893110410001664882CrossRefGoogle Scholar
Karlov, V. P., Butuzov, G. N., and Dobrokhotova, T. F. (1983). “Preparation and some properties of antimony(III) oxalate,” Zh. Neorg. Khim. ZNOKAQ 28, 21452146 (PDF entry 00-037-0646).Google Scholar
Korzun, B. B., Schorr, S., Schmitz, W., Fadzeyeva, A. A., Kommichau, G., and Bente, K. (2005). “Preparation of oxalates and BaBi1/2Sb1/2O3 from Ba(COO)2·0.5H2O and Sb(COO)2(OH) oxalates and Bi2O3 oxide,” J. Cryst. Growth JCRGAE 277, 205209 .10.1016/j.jcrysgro.2004.12.136CrossRefGoogle Scholar
Larson, A. C. and Von Dreele, R. B. (2004). General Structure Analysis System (GSAS), Report LAUR 86-748, Los Alamos National Laboratory, Los Alamos, NM.Google Scholar
Le Bail, A. (2008). “The profile of a Bragg reflection for extracting intensities,” in Powder Diffraction: Theory and Practice, edited by Dinnebier, R. E. and Billinge, S. J. L. (Royal Society of Chemistry, Cambridge), pp. 134165 .10.1039/9781847558237-00134CrossRefGoogle Scholar
Louër, D. and Boultif, A. (2007). “Powder pattern indexing and the dichotomy algorithm,” Z. Kristallogr. ZEKRDZ 26, 191196 .10.1524/zksu.2007.2007.suppl_26.191CrossRefGoogle Scholar
Macrae, C. F., Bruno, I. J., Chisholm, J. A., Edgington, P. R., McCabe, P., Pidcock, E., Rodriguez-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. JACGAR 41, 466470 .10.1107/S0021889807067908CrossRefGoogle Scholar
Marsh, R. E. (1997). “The perils of Cc revisited,” Acta Crystallogr., Sect. B: Struct. Sci. ASBSDK 53, 317322 (CSD Refcode SBFOXS01).10.1107/S010876819601381XCrossRefGoogle Scholar
Materials Data, Inc. (2008). JADE 8.5, computer software, Livermore, CA.Google Scholar
Millington, P. L. and Sowerby, D. B. (1992). “Phenylantimony(V) oxalates: Isolation and crystal structures of [SbPh4][SbPh2(ox)2],[SbPh3(OMe)]2ox and [SbPh4]2ox,” J. Chem. Soc. Dalton Trans. JCDTBI 11991204 (CSD Refcodes PABTAN, PABTER, and PABTIV).10.1039/dt9920001199CrossRefGoogle Scholar
Oszlányi, G. and Sütó, A. (2008). “The charge flipping algorithm,” Acta Crystallogr., Sect. A: Found. Crystallogr. ACACEQ 64, 123134 .10.1107/S0108767307046028CrossRefGoogle ScholarPubMed
Petříček, V., Dušek, M., and Palatinus, L. (2006). JANA2006, the crystallographic computing system, computer software, Institute of Physics, Praha, Czech Republic.Google Scholar
Poore, M. C. and Russell, D. R. (1971). “Crystal structure of the trisoxalatoantimonate(III) ion: Sterically-active lone pair in six-coordination,” J. Chem. Soc. D CCJDAO 1819 (CSD Refcodes AMSBOX and KSBOXL).10.1039/c29710000018CrossRefGoogle Scholar
Rietveld, H. M. (1969). “Line profile refinement method for nuclear and magnetic structures,” J. Appl. Crystallogr. JACGAR 2, 6571 .10.1107/S0021889869006558CrossRefGoogle Scholar
Rivenet, M., Roussel, P., and Abraham, F. (2008). “One-dimensional inorganic arrangement in the bismuth oxalate hydroxide Bi(C2O4)OH,” J. Solid State Chem. JSSCBI 181, 25862590 (ICSD collection code 419313).10.1016/j.jssc.2008.06.031CrossRefGoogle Scholar
Schwarz, W., Schmidt, A., and Blösl, S. (1981). “Reaction of the 1:1 addition compound of antimony(V) chloride and water with oxalic or squaric acid,” Z. Anorg. Allg. Chem. ZAACAB 477, 113118 (CSD Refocde OCOXSB).10.1002/zaac.19814770614CrossRefGoogle Scholar
Sidey, V. (2009). “The alternative presentation of the Brown-Wu bond-valence parameters for some s2-cation/O2− pairs,” Acta Crystallogr., Sect. B: Struct. Sci. ASBSDK 65, 99101 .10.1107/S0108768108040147CrossRefGoogle ScholarPubMed
Southerington, I. G., Begley, M. J., and Sowerby, D. B. (1991). “A new diphenylantimony oxide containing a planar Sb6O6 ring,” J. Chem. Soc., Chem. Commun. JCCCAT 15551556 (CSD Refocde JODLET).10.1039/c39910001555CrossRefGoogle Scholar
Stephens, P. W. (1999). “Phenomenological model of anisotropic broadening in powder diffraction,” J. Appl. Crystallogr. JACGAR 32, 281289 .10.1107/S0021889898006001CrossRefGoogle Scholar
Udovenko, A. A., Sigula, N. I., and Davidovich, R. L. (1981a). “Crystal structure of cesium dioxalatotetrefluorodiantimonate(III) monohydrate,” Koord. Khim. KOKHDC 7, 17081712 (CSD Refcode BAYLUI).Google Scholar
Udovenko, A. A., Sigula, N. I., Samarets, L. V., and Davidovich, R. L. (1981b). “Crystal structure of ammonium trioxalatotetrefluorodiantimonate(III) dihydrate,” Koord. Khim. KOKHDC 7, 450454 (CSD Refcode FLOXSB).Google Scholar