Hostname: page-component-8448b6f56d-sxzjt Total loading time: 0 Render date: 2024-04-19T01:43:57.568Z Has data issue: false hasContentIssue false

Using in-situ techniques to probe high-temperature reactions: thermochemical cycles for the production of synthetic fuels from CO2 and water

Published online by Cambridge University Press:  15 June 2012

Eric N. Coker*
Affiliation:
Sandia National Laboratories, PO Box 5800, MS 1349, Albuquerque, NM 87185-1349, USA
Mark A. Rodriguez
Affiliation:
Sandia National Laboratories, PO Box 5800, MS 1349, Albuquerque, NM 87185-1349, USA
Andrea Ambrosini
Affiliation:
Sandia National Laboratories, PO Box 5800, MS 1349, Albuquerque, NM 87185-1349, USA
James E. Miller
Affiliation:
Sandia National Laboratories, PO Box 5800, MS 1349, Albuquerque, NM 87185-1349, USA
Ellen B. Stechel
Affiliation:
Sandia National Laboratories, PO Box 5800, MS 1349, Albuquerque, NM 87185-1349, USA
*
a)Author to whom correspondence should be addressed. Electronic mail: encoker@sandia.gov

Abstract

Ferrites are promising materials for enabling solar-thermochemical cycles. Such cycles utilize solar-thermal energy to reduce the metal oxide, which is then re-oxidized by H2O or CO2, producing H2 or CO, respectively. Mixing ferrites with zirconia or yttria-stabilized zirconia (YSZ) greatly improves their cyclabilities. In order to understand this system, we have studied the behavior of iron oxide/8YSZ (8 mol-% Y2O3 in ZrO2) using in situ X-ray diffraction and thermogravimetric analyses at temperatures up to 1500 °C and under controlled atmosphere. The solubility of iron oxide in 8YSZ measured by XRD at room temperature was 9.4 mol-% Fe. The solubility increased to at least 10.4 mol-% Fe when heated between 800 and 1000 °C under inert atmosphere. Furthermore iron was found to migrate in and out of the 8YSZ phase as the temperature and oxidation state of the iron changed. In samples containing >9.4 mol-% Fe, stepwise heating to 1400 °C under helium caused reduction of Fe2O3 to Fe3O4 to FeO. Exposure of the FeO-containing material to CO2 at 1100 °C re-oxidized FeO to Fe3O4 with evolution of CO. Thermogravimetric analysis during thermochemical cycling of materials with a range of iron contents showed that samples with mostly dissolved iron utilized a greater proportion of the iron atoms present than did samples possessing a greater fraction of un-dissolved iron oxides.

Type
Technical Articles
Copyright
Copyright © International Centre for Diffraction Data 2012

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

Allendorf, M.D., Diver, R.B., Siegel, N.P., and Miller, J.E. (2008). “Two-step water splitting using mixed-metal ferrites: thermodynamic analysis and characterization of synthesized materials,” Energy & Fuels, 22, 41154124.CrossRefGoogle Scholar
Ambrosini, A., Coker, E.N., Rodriguez, M.A., Livers, S., Evans, L.R., Miller, J.E., and Stechel, E.B., (2010). “Synthesis and characterization of ferrite materials for thermochemical CO2 splitting using concentrated solar energy” ACS Symposium Series, 2010, 1056 (Advances in CO 2 Conversion and Utilization), 113; Hu, Y. (ed.); American Chemical Society: Washington, DC.Google Scholar
Bechta, S.V., Krushinov, E.V., Almjashev, V.I., Vitol, S.A., Mezentseva, L.P., Petrov, Yu.B., Lopukh, D.B., Khabensky, V.B., Barrachin, M., Hellmann, S., Froment, K., Fischer, M., Tromm, W., Bottomley, D., Defoort, F., and Gusarov, V.V. (2006). “Phase Diagram of the ZrO2–FeO System,” J. Nuclear Materials, 348, 114121.CrossRefGoogle Scholar
Beck, H.P. and Kaliba, C. (1990). “On the solubility of Fe, Cr and Nb in ZrO2 and its effect on thermal dilatation and polymorphic transition,” Mat. Res. Bull., 25, 11611168.CrossRefGoogle Scholar
Berry, F.J., Loretto, M.H., and Smith, M.R.J. (1989). “Iron-zirconium oxides: An investigation of structural transformations by X-ray diffraction, electron diffraction, and iron-57 Mössbauer spectroscopy,” Solid State Chemistry, 83, 9199.CrossRefGoogle Scholar
Coker, E.N., Ambrosini, A., Rodriguez, M.A., and Miller, J.E., (2011a). “Ferrite-YSZ composites for solar thermochemical production of synthetic fuels: In operando characterization of CO2 reduction,” J. Mater. Chem., 21, 10767–76.CrossRefGoogle Scholar
Coker, E.N., Ambrosini, A., Rodriguez, M.A., Garino, T.J., and Miller, J.E., (2011b). “Production of hydrogen and carbon monoxide from water and carbon dioxide through metal oxide thermochemical cycles,” Ceram. Trans., 224, 3749.CrossRefGoogle Scholar
Coker, E.N., Ohlhausen, J.A., Ambrosini, A., and Miller, J.E., (2012). “Oxygen transport and isotopic exchange in iron oxide/YSZ thermochemically-active materials via splitting of C(18O)2 at high temperature studied by thermogravimetric analysis and secondary ion mass spectrometry,” J. Mater. Chem., 22(14), 67266732.CrossRefGoogle Scholar
Davison, S., Kershaw, R., Dwight, K., and Wold, A.J. (1988). “Preparation and characterization of cubic ZrO2 stabilized by Fe(III) and Fe(II),” Solid State Chemistry, 73, 4751.CrossRefGoogle Scholar
Diver, R.B., Miller, J.E., Allendorf, M.D., Siegel, N.P., and Hogan, R.E. (2008). “Solar thermochemical water-splitting ferrite-cycle heat engines,” J. Solar Energy Eng., 130, 041001–1.CrossRefGoogle Scholar
Ghigna, P., Spinolo, G., Anselmi-Tamburini, U., Maglia, F., Dapiaggi, M., Spina, G., and Cianchi, L., (1999). “Fe-doped zirconium oxide produced by self-sustained high-temperature synthesis: Evidence for an Fe-Zr direct bond,” J. Am. Chem. Soc., 121, 301307.CrossRefGoogle Scholar
Hartmanova, M., Poulsen, F.W., Hanic, F., Putyera, K., Tunega, D., Urusovskaya, A.A., and Oreshnikova, T.V., (1994). “Influence of copper-doping and iron-doping on cubic yttria-stabilized zirconia,” J. Mater. Sci., 29, 21522158.CrossRefGoogle Scholar
Inwang, I.B., Chyad, F., and McColm, I.J., (1995). “Crystallization of iron (III) zirconia co-gels,” J. Mater. Chem., 5(8), 12091213.CrossRefGoogle Scholar
Jiang, J.Z., Poulsen, F.W., and Mørup, S., (1999). “Structure and thermal stability of nanostructured iron-doped zirconia prepared by high-energy ball milling,” J. Mater. Res., 14(4), 13431352.CrossRefGoogle Scholar
Kodama, T. (2003). “High-temperature solar chemistry for converting solar heat to chemical fuels,” Prog. En. Comb. Sci. 29, 567597.CrossRefGoogle Scholar
Kodama, T. and Gokon, N. (2007). “Thermochemical cycles for high temperature solar hydrogen production,” Chem. Rev., 107, 40484077.CrossRefGoogle ScholarPubMed
Kodama, T., Kondoh, Y., Kiyama, A., and Shimizu, K.-I., (2003) “Hydrogen production by solar thermochemical water-splitting/methane-reforming process” Proceedings of the ASME ISES Conference, Hawaii, p.121128.Google Scholar
Kodama, T., Nakamuro, Y., and Mizuno, T. (2006). “A two-step thermochemical water splitting by iron-oxide on stabilized zirconia,” J. Sol. Energy Eng.-Trans. ASME, 128, 37.CrossRefGoogle Scholar
Lajavardi, M., Kenney, D.J., and Lin, S.H., (2000). “Time-resolved high and low temperature phase transitions of the nanocrystalline cubic phase in the Y2O3-ZrO2 and Fe2O3-ZrO2 system,” J. Chinese Chem. Soc., 47, 10651075.CrossRefGoogle Scholar
Li, P., Chen, I.W., and Penner-Hahn, J.E., (1994). “Effect of dopants on zirconia stabilization – an X-ray-absorption study. 1. Trivalent dopants,” J. Am. Ceram. Soc., 77(1), 118128.CrossRefGoogle Scholar
Meredig, B. and Wolverton, C. (2009). “First-principles thermodynamic framework for the evaluation of thermochemical H2O- or CO2-splitting materials,” Phys. Rev. B, 80, 245119245126.CrossRefGoogle Scholar
Miller, J.E. (2007). “Initial case for splitting carbon dioxide to carbon monoxide and oxygen” Sandia Report, SAND2007-8012, (order at “”).Google Scholar
Miller, J.E., Allendorf, M.D., Diver, R.B., Evans, L.R., Siegel, N.P., and Stuecker, J.N., (2007). “Metal oxide composites and structures for ultra-high temperature solar thermochemical cycles,” J. Mater. Sci., 43, 47144728.CrossRefGoogle Scholar
Nakamura, T., (1977). “Hydrogen production from water utilizing solar heat at high temperatures,” Solar En., 19, 467475.CrossRefGoogle Scholar
Raming, T., Winnubst, L., and Verweij, H., (2002). “The synthesis and characterisation of mixed Y2O3-doped zirconia and α-Fe2O3 nanosized powders,” J. Mater. Chem., 12, 37053711.CrossRefGoogle Scholar
Shannon, R.D. and Prewitt, C.T., (1969). “Effective ionic radii in oxides and fluorides,” Acta Cryst., B25, 925.Google Scholar
Štefanić, G., Gržeta, B., and Musić, S., (2000). “Influence of oxygen on the thermal behavior of the ZrO2-Fe2O3 system,” Mater. Chem. Phys., 65, 216221.CrossRefGoogle Scholar
Štefanić, G., Gržeta, B., Nomurab, K., Trojkoa, R., and Musić, S., (2001). “The influence of thermal treatment on phase development in ZrO2-Fe2O3 and HfO2-Fe2O3 systems,” J. Alloys Compounds, 327, 151160.CrossRefGoogle Scholar
Steinfeld, A. (2005). “Solar thermochemical production of hydrogen – a review,” Solar En. 78, 603615.CrossRefGoogle Scholar
Vegard, L., (1921). “Die Konstitution der Mischkristalle und die Raumfüllung der Atome,” Z Phys., 5, 1726.CrossRefGoogle Scholar
Verkerk, M.J., Winnubst, A.J.A., and Burggraaf, A.J., (1982). “Effect of impurities on sintering and conductivity of yttria-stabilized zirconia,” J. Mater. Sci., 17, 3113–22.CrossRefGoogle Scholar
Wilhelm, R.V. and Howarth, D.S., (1979). “Iron oxide-doped yttria-stabilized zirconia ceramis: Iron solubility and electrical conductivity,” Ceram. Bull., 58, 229.Google Scholar
Wyrwalski, F., Lamonier, J.F., Siffert, S., Zhilinskaya, E.A., Gengembre, L., and Aboukaïs, A., (2005). “Bulk and surface structures of iron doped zirconium oxide systems: Influence of preparation method,” J. Mater. Sci., 40, 933942.CrossRefGoogle Scholar