Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-05-01T06:44:32.238Z Has data issue: false hasContentIssue false

Interfacial X-Ray Scattering From Small Surfaces: Adapting Mineral-Fluid Structure Methods for Microcrystalline Materials

Published online by Cambridge University Press:  01 January 2024

Joanne E. Stubbs*
Affiliation:
Center for Advanced Radiation Sources, The University of Chicago, Chicago, IL, USA
Anna K. Wanhala
Affiliation:
Center for Advanced Radiation Sources, The University of Chicago, Chicago, IL, USA
Peter J. Eng
Affiliation:
Center for Advanced Radiation Sources, The University of Chicago, Chicago, IL, USA James Franck Institute, The University of Chicago, Chicago, IL, USA
*
*E-mail address of corresponding author: stubbs@cars.uchicago.edu

Abstract

Crystal truncation rod (CTR) X-ray diffraction is an invaluable tool for measuring mineral surface and adsorbate structures, and has been applied to several environmentally and geochemically important systems. Traditionally, the method has been restricted to single crystals with lateral dimensions >3 mm. Minerals that meet this size criterion represent a minute fraction of those that are relevant to interfacial geochemistry questions, however. Crystal screening, data collection, and CTR measurement methods have been developed for crystals of <0.3 mm in lateral size using the manganese oxide mineral chalcophanite (ZnMn3O7·3H2O) as a case study. This work demonstrates the feasibility of applying the CTR technique to previously inaccessible surfaces, opening up a large suite of candidate substrates for future study.

Type
Article
Copyright
Copyright © Clay Minerals Society 2021

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.)

Footnotes

This article is a product of the 2020 CMS Workshop held in conjunction with the 57th Annual Meeting of The Clay Minerals Society, Richland, Washington, USA.

This manuscript was presented at the 57th Annual Meeting of The Clay Minerals Society as part of its 2020 Workshop on "Emerging Methods in Clay Science" held at Pacific Northwest National Laboratories, Richland, Washington, USA

References

Bargar, J. R., Fuller, C. C., Marcus, M. A., Brearley, A. J., De la Rosa, M. P., Webb, S. M., & Caldwell, W. A. (2009). Structural characterization of terrestrial microbial Mn oxides from Pinal Creek, AZ. Geochimica et Cosmochimica Acta, 73(4), 889910. https://doi.org/10.1016/j.gca.2008.10.036Google Scholar
Bellucci, F., Lee, S., Kubicki, J., Bandura, A., Zhang, Z., Wesolowski, D., & Fenter, P. (2015). Rb+ adsorption at the quartz(101)-aqueous interface: Comparison of resonant anomalous x-ray reflectivity with ab initio calculations. Journal of Physical Chemistry C, 119(9), 47784788. https://doi.org/10.1021/jp510139tGoogle Scholar
Bourg, I., Lee, S., Fenter, P., & Tournassat, C. (2017). Stern layer structure and energetics at mica-water interfaces. Journal of Physical Chemistry C, 121(17), 94029412. https://doi.org/10.1021/acs.jpcc.7b01828CrossRefGoogle Scholar
Bracco, J. N., Lee, S. S., Stubbs, J. E., Eng, P. J., Heberling, F., Fenter, P., & Stack, A. G. (2017). Hydration structure of the barite (001)-water interface: Comparison of x-ray Reflectivity with molecular dynamics simulations. Journal of Physical Chemistry C, 121(22), 1223612248. https://doi.org/10.1021/acs.jpcc.7b02943Google Scholar
Bracco, J., Lee, S., Stubbs, J., Eng, P., Jindra, S., Warren, D., Kommu, A., Fenter, P., Kubicki, J., & Stack, A. (2019). Simultaneous adsorption and incorporation of Sr2+ at the barite (001)-water interface. Journal of Physical Chemistry C, 123(2), 11941207. https://doi.org/10.1021/acs.jpcc.8b08848Google Scholar
Brugman, S., Townsend, E., Smets, M., Accordini, P., & Vlieg, E. (2018). Concentration-dependent adsorption of CsI at the muscovite-electrolyte interface. Langmuir, 34(13), 38213826. https://doi.org/10.1021/acs.langmuir.8b00038Google Scholar
Brugman, S. J. T., Werkhoven, B. L., Townsend, E. R., Accordini, P., van Roij, R., & Vlieg, E. (2020). Monovalent-divalent cation competition at the muscovite mica surface: Experiment and theory. Journal of Colloid and Interface Science, 559, 291303. https://doi.org/10.1016/j.jcis.2019.10.009Google Scholar
Callagon, E., Fenter, P., Nagy, K., & Sturchio, N. (2014). Incorporation of Pb at the calcite (104)-water interface. Environmental Science & Technology, 48(16), 92639269. https://doi.org/10.1021/es5014888Google Scholar
Callagon, E., Lee, S., Eng, P., Laanait, N., Sturchio, N., Nagy, K., & Fenter, P. (2017). Heteroepitaxial growth of cadmium carbonate at dolomite and calcite surfaces: Mechanisms and rates. Geochimica et Cosmochimica Acta, 205, 360380. https://doi.org/10.1016/j.gca.2016.12.007Google Scholar
Catalano, J. (2010). Relaxations and interfacial water ordering at the corundum (110) Surface. Journal of Physical Chemistry C, 114(14), 66246630. https://doi.org/10.1021/jp100455sGoogle Scholar
Catalano, J. (2011). Weak interfacial water ordering on isostructural hematite and corundum (001) surfaces. Geochimica et Cosmochimica Acta, 75(8), 20622071. https://doi.org/10.1016/j.gca.2011.01.025Google Scholar
Catalano, J., Trainor, T., Eng, P., Waychunas, G., & Brown, G. (2005). CTR diffraction and grazing-incidence EXAFS study of U(VI) adsorption onto alpha-Al2O3 and alpha-Fe2O3 (1(1)over-bar02) surfaces. Geochimica et Cosmochimica Acta, 69(14), 35553572. https://doi.org/10.1016/j.gca.2005.03.044Google Scholar
Catalano, J., Park, C., Zhang, Z., & Fenter, P. (2006a). Termination and water adsorption at the alpha-Al2O3(012) - Aqueous solution interface. Langmuir, 22(10), 46684673. https://doi.org/10.1021/la060177sGoogle Scholar
Catalano, J., Zhang, Z., Fenter, P., & Bedzyk, M. (2006b). Inner-sphere adsorption geometry of Se(IV) at the hematite (100)-water interface. Journal of Colloid and Interface Science, 297(2), 665671. https://doi.org/10.1016/j.jcis.2005.11.026Google Scholar
Catalano, J., Zhang, Z., Park, C., Fenter, P., & Bedzyk, M. (2007a). Bridging arsenate surface complexes on the hematite (012) surface. Geochimica et Cosmochimica Acta, 71(8), 18831897. https://doi.org/10.1016/j.gca.2007.01.015Google Scholar
Catalano, J., Fenter, P., & Park, C. (2007b). Interfacial water structure on the (012) surface of hematite: Ordering and reactivity in comparison with corundum. Geochimica et Cosmochimica Acta, 71(22), 53135324. https://doi.org/10.1016/j.gca.2007.09.019Google Scholar
Catalano, J., Park, C., Fenter, P., & Zhang, Z. (2008). Simultaneous inner- and outer-sphere arsenate adsorption on corundum and hematite. Geochimica et Cosmochimica Acta, 72(8), 19862004. https://doi.org/10.1016/j.gca.2008.02.013Google Scholar
Catalano, J., Fenter, P., & Park, C. (2009). Water ordering and surface relaxations at the hematite (110)-water interface. Geochimica et Cosmochimica Acta, 73(8), 22422251. https://doi.org/10.1016/j.gca.2009.02.001Google Scholar
Catalano, J., Fenter, P., Park, C., Zhang, Z., & Rosso, K. (2010). Structure and oxidation state of hematite surfaces reacted with aqueous Fe(II) at acidic and neutral pH. Geochimica et Cosmochimica Acta, 74(5), 14981512. https://doi.org/10.1016/j.gca.2009.12.018Google Scholar
Cheng, L., Fenter, P., Nagy, K., Schlegel, M., & Sturchio, N. (2001). Molecular-scale density oscillations in water adjacent to a mica surface. Physical Review Letters, 87(15), Article ARTN 156103. https://doi.org/10.1103/PhysRevLett.87.156103Google Scholar
Chiarello, R. P., & Sturchio, N. C. (1995). The calcite (10-14) cleavage surface in water - early results of a crystal truncation rod study. Geochimica et Cosmochimica Acta, 59(21), 45574561. https://doi.org/10.1016/0016-7037(95)00363-5Google Scholar
Clark, J. N., Ihli, J., Schenk, A. S., Kim, Y. Y., Kulak, A.N., Campbell, J. M., Nisbet, G., Meldrum, F., & Robinson, I. K. (2015). Three-dimensional imaging of dislocation propagation during crystal growth and dissolution. Nature Materials, 14(8), 780-+. https://doi.org/10.1038/nmat4320Google Scholar
de Poel, W., Pintea, S., de Jong, A., Drnec, J., Carla, F., Felici, R., & 5 others. (2014a). Dibenzo crown ether layer formation on muscovite mica. Langmuir, 30(42), 1257012577. https://doi.org/10.1021/la502879zGoogle Scholar
de Poel, W., Pintea, S., Drnec, J., Carla, F., Felici, R., Mulder, P., & 4 others. (2014b). Muscovite mica: Flatter than a pancake. Surface Science, 619, 1924. https://doi.org/10.1016/j.susc.2013.10.008Google Scholar
de Poel, W., Vaessen, S., Drnec, J., Engwerda, A., Townsend, E., Pintea, S., & 8 others. (2017). Metal ion-exchange on the muscovite mica surface. Surface Science, 665, 5661. https://doi.org/10.1016/j.susc.2017.08.013Google Scholar
de Vries, S. A., Goedtkindt, P., Bennett, S. L., Huisman, W. J., Zwanenburg, M. J., Smilgies, D. M., et al. (1998). Surface atomic structure of KDP crystals in aqueous solution: An explanation of the growth shape. Physical Review Letters, 80(10), 22292232. https://doi.org/10.1103/PhysRevLett.80.2229CrossRefGoogle Scholar
de Vries, S. A., Goedtkindt, P., Huisman, W. J., Zwanenburg, M. J., Feidenhans'l, R., Bennett, S. L., ... Vlieg, E. (1999). X-ray diffraction studies of potassium dihydrogen phosphate (KDP) crystal surfaces. Journal of Crystal Growth, 205(1–2), 202214. https://doi.org/10.1016/s0022-0248(99)00249-3Google Scholar
Eng, P., Trainor, T., Brown, G., Waychunas, G., Newville, M., Sutton, S., & Rivers, M. (2000). Structure of the hydrated alpha-Al2O3(0001) surface. Science, 288(5468), 10291033. https://doi.org/10.1126/science.288.5468.1029Google Scholar
Fenter, P. (2002). X-ray reflectivity as a probe of mineral-fluid interfaces: A user guide [Article|Proceedings Paper]\. Applications of Synchrotron Radiation in Low-Temperature Geochemistry and Environmental Sciences, 49, 149220. https://doi.org/10.2138/gsrmg.49.1.149Google Scholar
Fenter, P., & Sturchio, N. (1999). Structure and growth of stearate monolayers on calcite: First results of an in situ X-ray reflectivity study. Geochimica et Cosmochimica Acta, 63(19–20), 31453152. https://doi.org/10.1016/S0016-7037(99)00241-0Google Scholar
Fenter, P., & Sturchio, N. (2012). Calcite (104)-water interface structure, revisited. Geochimica et Cosmochimica Acta, 97, 5869. https://doi.org/10.1016/j.gca.2012.08.021Google Scholar
Fenter, P., Teng, H., Geissbuhler, P., Hanchar, J., Nagy, K., & Sturchio, N. (2000a). Atomic-scale structure of the orthoclase (001)-water interface measured with high-resolution X-ray reflectivity. Geochimica et Cosmochimica Acta, 64(21), 36633673. https://doi.org/10.1016/S0016-7037(00)00455-5Google Scholar
Fenter, P., Geissbuhler, P., DiMasi, E., Srajer, G., Sorensen, L., & Sturchio, N. (2000b). Surface speciation of calcite observed in situ by high-resolution X-ray reflectivity. Geochimica et Cosmochimica Acta, 64(7), 12211228. https://doi.org/10.1016/S0016-7037(99)00403-2Google Scholar
Fenter, P., McBride, M., Srajer, G., Sturchio, N., & Bosbach, D. (2001). Structure of barite (001)- and (210)-water interfaces. Journal of Physical Chemistry B, 105(34), 81128119. https://doi.org/10.1021/jp0105600Google Scholar
Fenter, P., Cheng, L., Park, C., Zhang, Z., & Sturchio, N. (2003a). Structure of the orthoclase (001)- and (010)-water interfaces by high-resolution X-ray reflectivity. Geochimica et Cosmochimica Acta, 67(22), 42674275. https://doi.org/10.1016/S0016-7037(03)00374-0Google Scholar
Fenter, P., Park, C., Cheng, L., Zhang, Z., Krekeler, M., & Sturchio, N. (2003b). Orthoclase dissolution kinetics probed by in situ X-ray reflectivity: Effects of temperature, pH, and crystal orientation. Geochimica et Cosmochimica Acta, 67(2), 197211, Article PII S0016-7037(02)01084-0. https://doi.org/10.1016/S0016-7037(02)01084-0Google Scholar
Fenter, P., Zhang, Z., Park, C., Sturchio, N., Hu, X., & Higgins, S. (2007). Structure and reactivity of the dolomite (104)-water interface: New insights into the dolomite problem. Geochimica et Cosmochimica Acta, 71(3), 566579. https://doi.org/10.1016/j.gca.2006.10.006Google Scholar
Fenter, P., Park, C., & Sturchio, N. (2008). Adsorption of Rb+ and Sr2+ at the orthoclase (001)-solution interface. Geochimica et Cosmochimica Acta, 72(7), 18481863. https://doi.org/10.1016/j.gca.2007.12.016|10.1016/j.uca.2007.12.016Google Scholar
Fenter, P., Lee, S., Park, C., Catalano, J., Zhang, Z., & Sturchio, N. (2010a). Probing interfacial reactions with X-ray reflectivity and X-ray reflection interface microscopy: Influence of NaCl on the dissolution of orthoclase at pOH 2 and 85 degrees C. Geochimica et Cosmochimica Acta, 74(12), 33963411. https://doi.org/10.1016/j.gca.2010.03.027Google Scholar
Fenter, P., Lee, S., Park, C., Soderholm, L., Wilson, R., & Schwindt, O. (2010b). Interaction of muscovite (001) with Pu3+ bearing solutions at pH 3 through ex-situ observations. Geochimica et Cosmochimica Acta, 74(24), 69846995. https://doi.org/10.1016/j.gca.2010.09.025Google Scholar
Fenter, P., Kerisit, S., Raiteri, P., & Gale, J. (2013). Is the Calcite-Water Interface Understood? Direct comparisons of molecular dynamics simulations with specular x-ray reflectivity data. Journal of Physical Chemistry C, 117(10), 50285042. https://doi.org/10.1021/jp310943sGoogle Scholar
Fenter, P., Zapol, P., He, H., & Sturchio, N. (2014). On the variation of dissolution rates at the orthoclase (001) surface with pH and temperature. Geochimica et Cosmochimica Acta, 141, 598611. https://doi.org/10.1016/j.gca.2014.06.019Google Scholar
Fuller, C. C., & Bargar, J. R. (2014). Processes of zinc attenuation by biogenic manganese oxides forming in the hyporheic zone of Pinal Creek, Arizona. Environmental Science & Technology, 48(4), 21652172. https://doi.org/10.1021/es402576fGoogle Scholar
Geissbuhler, P., Fenter, P., DiMasi, E., Srajer, G., Sorensen, L., & Sturchio, N. (2004). Three-dimensional structure of the calcite-water interface by surface X-ray scattering. Surface Science, 573(2), 191203. https://doi.org/10.1016/j.susc.2004.09.036Google Scholar
Ghose, S. K., Waychunas, G. A., Trainor, T. P., & Eng, P. J. (2010). Hydrated goethite (alpha-FeOOH) (100) interface structure: Ordered water and surface functional groups. Geochimica et Cosmochimica Acta, 74(7), 19431953. https://doi.org/10.1016/j.gca.2009.12.015Google Scholar
Heberling, F., Trainor, T., Lutzenkirchen, J., Eng, P., Denecke, M., & Bosbach, D. (2011). Structure and reactivity of the calcite-water interface. Journal of Colloid and Interface Science, 354(2), 843857. https://doi.org/10.1016/j.jcis.2010.10.047Google Scholar
Heberling, F., Eng, P., Denecke, M., Lutzenkirchen, J., & Geckeis, H. (2014). Electrolyte layering at the calcite(104)-water interface indicated by Rb+- and Se(VI) K-edge resonant interface diffraction. Physical Chemistry Chemical Physics, 16(25), 1278212792. https://doi.org/10.1039/c4cp00672kGoogle Scholar
Hellebrandt, S., Lee, S. S., Knope, K. E., Lussier, A. J., Stubbs, J. E., Eng, P. J., Soderholm, L., Fenter, P., & Schmidt, M. (2016). A comparison of adsorption, reduction, and polymerization of the plutonyl(VI) and uranyl(VI) Ions from solution onto the muscovite basal plane. Langmuir, 32(41), 1047310482. https://doi.org/10.1021/acs.langmuir.6b02513Google Scholar
Hochella, M. F., Kasama, T., Putnis, A., Putnis, C. V., & Moore, J. N. (2005). Environmentally important, poorly crystalline Fe/Mn hydrous oxides: Ferrihydrite and a possibly new vernadite-like mineral from the Clark Fork River Superfund Complex. American Mineralogist, 90(4), 718724. https://doi.org/10.2138/am.2005.1591Google Scholar
Hofmann, S., Voitchovsky, K., Spijker, P., Schmidt, M., & Stumpf, T. (2016). Visualising the molecular alteration of the calcite (104) – water interface by sodium nitrate. Scientific Reports, 6, Article 21576. https://doi.org/10.1038/srep21576Google Scholar
Ihli, J., Clark, J. N., Cote, A. S., Kim, Y. Y., Schenk, A. S., Kulak, A. N., & 6 others. (2016). Strain-relief by single dislocation loops in calcite crystals grown on self-assembled monolayers. Nature Communications, 7, Article 11878. https://doi.org/10.1038/ncomms11878Google Scholar
Ihli, J., Clark, J. N., Kanwal, N., Kim, Y. Y., Holden, M. A., Harder, R. J., Tang, C. C., Ashbrook, S. E., Robinson, I. K., & Meldrum, F. C. (2019). Visualization of the effect of additives on the nanostructures of individual bio-inspired calcite crystals. Chemical Science, 10(4), 11761185. https://doi.org/10.1039/c8sc03733gGoogle Scholar
Jun, Y., Ghose, S., Trainor, T., Eng, P., & Martin, S. (2007). Structure of the hydrated (10-14) surface of rhodochrosite (MnCO3). Environmental Science & Technology, 41(11), 39183925. https://doi.org/10.1021/es062171vGoogle Scholar
Kaminski, D., Radenovic, N., Deij, M. A., van Enckevort, W. J. P., & Vlieg, E. (2005). pH-dependent liquid order at the solid-solution interface of KH2PO4 crystals. Physical Review B, 72(24), Article 245404. https://doi.org/10.1103/PhysRevB.72.245404Google Scholar
Kaminski, D., Radenovic, N., Deij, M. A., van Enckevort, W. J. P., & Vlieg, E. (2006). Liquid ordering at the KDP {100}-solution interface. Crystal Growth & Design, 6(2), 588591. https://doi.org/10.1021/cg0502338Google Scholar
Kim, S., & Baik, S. (1994). X-ray study of step morphology of MgO(100) surfaces. Applied Surface Science, 78(3), 285292. https://doi.org/10.1016/0169-4332(94)90016-7Google Scholar
Kim, S., Baik, S., Kim, H., & Kim, C. (1993). X-ray study of step morphology on a cleaved MgO(100) surface. Surface Science, 294(1–2), L935–L938.Google Scholar
Kimball, B. E., Foster, A. L., Seal, R. R., Piatak, N. M., Webb, S. M., & Hammarstrom, J. M. (2016). Copper speciation in variably toxic sediments at the Ely Copper Mine, Vermont, United States. Environmental Science & Technology, 50(3), 11261136. https://doi.org/10.1021/acs.est.5b04081Google Scholar
Kohli, V., Zhang, Z., Park, C., & Fenter, P. (2010). Rb+ and Sr2+ Adsorption at the TiO2 (110)-electrolyte interface observed with resonant anomalous x-ray reflectivity. Langmuir, 26(2), 950958. https://doi.org/10.1021/la902419zGoogle Scholar
La Plante, E. C., Eng, P. J., Lee, S. S., Sturchio, N. C., Nagy, K. L., & Fenter, P. (2018). Evolution of strain in heteroepitaxial cadmium carbonate overgrowths on dolomite. Crystal Growth & Design, 18(5), 28712882. https://doi.org/10.1021/acs.cgd.7b01716Google Scholar
La Plante, E., Lee, S., Eng, P., Stubbs, J., Fenter, P., Sturchio, N., & Nagy, K. (2019). Dissolution kinetics of epitaxial cadmium carbonate overgrowths on dolomite. Acs Earth and Space Chemistry, 3(2), 212220. https://doi.org/10.1021/acsearthspacechem.8b00115Google Scholar
Lee, S., Nagy, K., & Fenter, P. (2007). Distribution of barium and fulvic acid at the mica-solution interface using in-situ X-ray reflectivity. Geochimica et Cosmochimica Acta, 71(23), 57635781. https://doi.org/10.1016/j.gca.2007.05.031Google Scholar
Lee, S., Fenter, P., Park, C., & Nagy, K. (2008). Fulvic acid sorption on muscovite mica as a function of pH and time using in situ X-ray reflectivity. Langmuir, 24(15), 78177829. https://doi.org/10.1021/la703456tGoogle Scholar
Lee, S., Nagy, K., Park, C., & Fenter, P. (2009). Enhanced uptake and modified distribution of mercury(II) by fulvic acid on the muscovite (001) Surface. Environmental Science& Technology, 43(14), 52955300. https://doi.org/10.1021/es900214eGoogle Scholar
Lee, S., Fenter, P., Park, C., Sturchio, N., & Nagy, K. (2010a). Hydrated cation speciation at the muscovite (001)-water interface. Langmuir, 26(22), 1664716651. https://doi.org/10.1021/la1032866Google Scholar
Lee, S., Park, C., Fenter, P., Sturchio, N., & Nagy, K. (2010b). Competitive adsorption of strontium and fulvic acid at the muscovite-solution interface observed with resonant anomalous X-ray reflectivity. Geochimica et Cosmochimica Acta, 74(6), 17621776. https://doi.org/10.1016/j.gca.2009.12.010Google Scholar
Lee, S., Nagy, K., Park, C., & Fenter, P. (2011). Heavy metal sorption at the muscovite (001)-fulvic acid interface. Environmental Science & Technology, 45(22), 95749581. https://doi.org/10.1021/es201323aGoogle Scholar
Lee, S., Fenter, P., Nagy, K., & Sturchio, N. (2012). Monovalent ion adsorption at the muscovite (001)-solution interface: Relationships among ion coverage and speciation, interfacial water structure, and substrate relaxation. Langmuir, 28(23), 86378650. https://doi.org/10.1021/la300032hGoogle Scholar
Lee, S., Fenter, P., Nagy, K., & Sturchio, N. (2013a). Changes in adsorption free energy and speciation during competitive adsorption between monovalent cations at the muscovite (001)- water interface. Geochimica et Cosmochimica Acta, 123, 416426. https://doi.org/10.1016/j.gca.2013.07.033Google Scholar
Lee, S., Schmidt, M., Laanait, N., Sturchio, N., & Fenter, P. (2013b). Investigation of structure, adsorption free energy, and overcharging behavior of trivalent yttrium adsorbed at the muscovite (001)-water interface. Journal of Physical Chemistry C, 117(45), 2373823749. https://doi.org/10.1021/jp407693xGoogle Scholar
Lee, S., Heberling, F., Sturchio, N., Eng, P., & Fenter, P. (2016). Surface charge of the calcite (104) terrace measured by Rb+ adsorption in aqueous solutions using resonant anomalous x-ray reflectivity. Journal of Physical Chemistry C, 120(28), 1521615223. https://doi.org/10.1021/acs.jpcc.6b04364Google Scholar
Lee, S., Fenter, P., Nagy, K., & Sturchio, N. (2017). Real-time observation of cation exchange kinetics and dynamics at the muscovite-water interface. Nature Communications, 8, Article ARTN 15826. https://doi.org/10.1038/ncomms15826Google Scholar
Lee, S., Schmidt, M., Sturchio, N., Nagy, K., & Fenter, P. (2019). Effect of pH on the formation of gibbsite-layer films at the muscovite (001)-water interface. Journal of Physical Chemistry C, 123(11), 65606571. https://doi.org/10.1021/acs.jpcc.8b12122Google Scholar
Liu, X. P., Lin, W., Chen, B., Zhang, F. C., Zhao, P. Q., Parsons, A., Rau, C., & Robinson, I. (2018). Coherent diffraction study of calcite crystallization during the hydration of tricalcium silicate. Materials & Design, 157, 251257. https://doi.org/10.1016/j.matdes.2018.07.031Google Scholar
Lutzenkirchen, J., Heberling, F., Supljika, F., Preocanin, T., Kallay, N., Johann, F., Weisser, L., & Eng, P. (2015). Structure-charge relationship - the case of hematite (001). Faraday Discussions, 180, 5579. https://doi.org/10.1039/c4fd00260aGoogle Scholar
Magdans, U., Gies, H., Torrelles, X., & Rius, J. (2006). Investigation of the {104} surface of calcite under dry and humid atmospheric conditions with grazing incidence X-ray diffraction (GIXRD). European Journal of Mineralogy, 18(1), 8391. https://doi.org/10.1127/0935-1221/2006/0018-0083Google Scholar
Manceau, A., Lanson, B., Schlegel, M. L., Harge, J. C., Musso, M., Eybert-Berard, L., Hazemann, J.-L., Chateiger, D., & Lamble, G. M. (2000). Quantitative Zn speciation in smelter-contaminated soils by EXAFS spectroscopy. American Journal of Science, 300(4), 289343. https://doi.org/10.2475/ajs.300.4.289Google Scholar
McBriarty, M. E., von Rudorff, G. F., Stubbs, J. E., Eng, P. J., Blumberger, J., & Rosso, K. M. (2017). Dynamic stabilization of metal oxide-water Interfaces. Journal of the American Chemical Society, 139(7), 25812584. https://doi.org/10.1021/jacs.6b13096Google Scholar
McBriarty, M. E., Stubbs, J. E., Eng, P. J., & Rosso, K. M. (2018). Potential-specific structure at the hematite-electrolyte interface. Advanced Functional Materials, 28(8), 1705618. https://doi.org/10.1002/adfm.201705618Google Scholar
McBriarty, M., Stubbs, J., Eng, P., & Rosso, K. (2019). Reductive dissolution mechanisms at the hematite-electrolyte interface probed by in situ x-ray scattering. Journal of Physical Chemistry C, 123(13), 80778085. https://doi.org/10.1021/acs.jpcc.8b07413Google Scholar
Morin, G., Ostergren, J. D., Juillot, F., Ildefonse, P., Calas, G., & Brown, G. E. (1999). XAFS determination of the chemical form of lead in smelter-contaminated soils and mine tailings: Importance of adsorption processes. American Mineralogist, 84(3), 420434. https://doi.org/10.2138/am-1999-0327Google Scholar
Noerpel, M., Lee, S., & Lenhart, J. (2016). X-ray analyses of lead adsorption on the (001), (110), and (012) hematite surfaces. Environmental Science & Technology, 50(22), 1228312291. https://doi.org/10.1021/acs.est.6b03913Google Scholar
Pareek, A., Torrelles, X., Angermund, K., Rius, J., Magdans, U., & Gies, H. (2009). Competitive adsorption of glycine and water on the fluorapatite (100) surface. Langmuir, 25(3), 14531458. https://doi.org/10.1021/la802706yGoogle Scholar
Park, C., Fenter, P., Zhang, Z., Cheng, L., & Sturchio, N. (2004). Structure of the fluorapatite (100)-water interface by high-resolution X-ray reflectivity. American Mineralogist, 89(11–12), 16471654. https://doi.org/10.2138/am-2004-11-1209Google Scholar
Park, C., Fenter, P. A., Sturchio, N. C., & Regalbuto, J. R. (2005). Probing outer-sphere adsorption of aqueous metal complexes at the oxide-water interface with resonant anomalous X-ray reflectivity. Physical Review Letters, 94(7), Article 076104. https://doi.org/10.1103/PhysRevLett.94.076104Google Scholar
Petitto, S., Tanwar, K., Ghose, S., Eng, P., & Trainor, T. (2010). Surface structure of magnetite (111) under hydrated conditions by crystal truncation rod diffraction. Surface Science, 604(13–14), 10821093. https://doi.org/10.1016/j.susc.2010.03.014Google Scholar
Pfeifer, M. A., Williams, G. J., Vartanyants, I. A., Harder, R., & Robinson, I. K. (2006). Three-dimensional mapping of a deformation field inside a nanocrystal. Nature, 442(7098), 6366. https://doi.org/10.1038/nature04867Google Scholar
Pintea, S., de Poel, W., de Jong, A., Vonk, V., van der Asdonk, P., Drnec, J., Balmes, O., Isern, H., Dufrane, T., Felici, R., & Vlieg, E. (2016). Solid-Liquid Interface Structure of Muscovite Mica in CsCl and RbBr Solutions. Langmuir, 32(49), 1295512965. https://doi.org/10.1021/acs.langmuir.6b02121Google Scholar
Pintea, S., de Poel, W., de Jong, A., Felici, R., & Vlieg, E. (2018). Solid-liquid interface structure of muscovite mica in SrCl2 and BaCl2 solutions. Langmuir, 34(14), 42414248. https://doi.org/10.1021/acs.langmuir.8b00504CrossRefGoogle ScholarPubMed
Post, J. E. (1999). Manganese oxide minerals: Crystal structures and economic and environmental significance [Article; Proceedings Paper]. Proceedings of the National Academy of Sciences of the United States of America, 96(7), 34473454. https://doi.org/10.1073/pnas.96.7.3447Google Scholar
Post, J., & Appleman, D. (1988). Chalcophanite, ZnMn3O7.3H2O -New crystal-structure determinations. American Mineralogist, 73(11–12), 14011404.Google Scholar
Post, J., & Heaney, P. (2014). Time-resolved synchrotron X-ray diffraction study of the dehydration behavior of chalcophanite. American Mineralogist, 99(10), 19561961. https://doi.org/10.2138/am-2014-4760Google Scholar
Qiu, C. R., Eng, P. J., Hennig, C., & Schmidt, M. (2018a). Formation and aggregation of ZrO2 nanoparticles on muscovite (001). Journal of Physical Chemistry C, 122(7), 38653874. https://doi.org/10.1021/acs.jpcc.7b10101Google Scholar
Qiu, C. R., Majs, F., Eng, P. J., Stubbs, J. E., Douglas, T. A., Schmidt, M., & Trainor, T. P. (2018b). In situ structural study of the surface complexation of lead(II) on the chemically mechanically polished hematite (1-102) surface. Journal of Colloid and Interface Science, 524, 6575. https://doi.org/10.1016/j.jcis.2018.04.005Google Scholar
Reedijk, M. F., Arsic, J., Hollander, F. F. A., de Vries, S. A., & Vlieg, E. (2003). Liquid order at the interface of KDP crystals with water: Evidence for icelike layers. Physical Review Letters, 90(6), Article 066103. https://doi.org/10.1103/PhysRevLett.90.066103Google Scholar
Robach, O., Renaud, G., & Barbier, A. (1998). Very-high-quality MgO(001) surfaces: roughness, rumpling and relaxation. Surface Science, 401(2), 227235. https://doi.org/10.1016/S0039-6028(97)01082-0Google Scholar
Robinson, I. (1986). Crystal truncation rods and surface roughness. Physical Review B, 33(6), 38303836. https://doi.org/10.1103/PhysRevB.33.3830Google Scholar
Robinson, I. K., Vartanyants, I. A., Williams, G. J., Pfeifer, M. A., & Pitney, J. A. (2001). Reconstruction of the shapes of gold nanocrystals using coherent x-ray diffraction. Physical Review Letters, 87(19), Article 195505. https://doi.org/10.1103/PhysRevLett.87.195505Google Scholar
Schlegel, M., Nagy, K., Fenter, P., & Sturchio, N. (2002). Structures of quartz (100)- and (101)-water interfaces determined by X-ray reflectivity and atomic force microscopy of natural growth surfaces. Geochimica et Cosmochimica Acta, 66(17), 30373054. https://doi.org/10.1016/S0016-7037(02)00912-2Google Scholar
Schlegel, M., Nagy, K., Fenter, P., Cheng, L., Sturchio, N., & Jacobsen, S. (2006). Cation sorption on the muscovite (001) surface in chloride solutions using high-resolution X-ray reflectivity. Geochimica et Cosmochimica Acta, 70(14), 35493565. https://doi.org/10.1016/j.gca.2006.04.011Google Scholar
Schmidt, M., Lee, S., Wilson, R., Soderholm, L., & Fenter, P. (2012a). Sorption of tetravalent thorium on muscovite. Geochimica et Cosmochimica Acta, 88, 6676. https://doi.org/10.1016/j.gca.2012.04.001Google Scholar
Schmidt, M., Wilson, R., Lee, S., Soderholm, L., & Fenter, P. (2012b). Adsorption of plutonium oxide nanoparticles. Langmuir, 28(5), 26202627. https://doi.org/10.1021/la2037247Google Scholar
Schmidt, M., Lee, S., Wilson, R., Knope, K., Bellucci, F., Eng, P., Stubbs, J., Soderholm, L., & Fenter, P. (2013). Surface-mediated formation of Pu(IV) nanoparticles at the muscovite-electrolyte interface. Environmental Science & Technology, 47(24), 1417814184. https://doi.org/10.1021/es4037258Google Scholar
Schmidt, M., Hellebrandt, S., Knope, K., Lee, S., Stubbs, J., Eng, P., Soderholm, L., & Fenter, P. (2015). Effects of the background electrolyte on Th(IV) sorption to muscovite mica. Geochimica et Cosmochimica Acta, 165, 280293. https://doi.org/10.1016/j.gca.2015.05.039Google Scholar
Shope, C. L., Xie, Y., & Gammons, C. H. (2006). The influence of hydrous Mn-Zn oxides on diel cycling of Zn in an alkaline stream draining abandoned mine lands. Applied Geochemistry, 21(3), 476491. https://doi.org/10.1016/j.apgeochem.2005.11.004Google Scholar
Stack, A. G., Stubbs, J. E., Srinivasan, S. G., Roy, S., Bryantsev, V. S., Eng, P. J., Custelcean, R., Gordon, A. G., & Hexel, C. R. (2018). Mineral-water interface structure of xenotime (YPO4) {100}. Journal of Physical Chemistry C, 122(35), 2023220243. https://doi.org/10.1021/acs.jpcc.8b04015Google Scholar
Stubbs, J. E., Chaka, A. M., Ilton, E. S., Biwer, C. A., Engelhard, M. H., Bargar, J. R., & Eng, P. J. (2015). UO2 oxidative corrosion by nonclassical diffusion. Physical Review Letters, 114(24), Article 246103. https://doi.org/10.1103/PhysRevLett.114.246103Google Scholar
Stubbs, J. E., Biwer, C. A., Chaka, A. M., Ilton, E. S., Du, Y., Bargar, J. R., & Eng, P. J. (2017). Oxidative corrosion of the UO2(001) surface by nonclassical diffusion. Langmuir, 33(46), 1318913196. https://doi.org/10.1021/acs.langmuir.7b02800Google Scholar
Stubbs, J., Legg, B., Lee, S., Dera, P., De Yoreo, J., Fenter, P., & Eng, P. (2019). Epitaxial growth of gibbsite sheets on the basal surface of muscovite mica. Journal of Physical Chemistry C, 123(45), 2761527627. https://doi.org/10.1021/acs.jpcc.9b08219Google Scholar
Tanwar, K., Lo, C., Eng, P., Catalano, J., Walko, D., Brown, G., Waychunas, G., Chaka, A., & Trainor, T. (2007a). Surface diffraction study of the hydrated hematite (1-102) surface. Surface Science, 601(2), 460474. https://doi.org/10.1016/j.susc.2006.10.021Google Scholar
Tanwar, K., Catalano, J., Petitto, S., Ghose, S., Eng, P., & Trainor, T. (2007b). Hydrated alpha-Fe2O3( ) surface structure: Role of surface preparation. Surface Science, 601(12), L59L64. https://doi.org/10.1016/j.susc.2007.04.115Google Scholar
Tanwar, K., Petitto, S., Ghose, S., Eng, P., & Trainor, T. (2008). Structural study of Fe(II) adsorption on hematite(1-102). Geochimica et Cosmochimica Acta, 72(14), 33113325. https://doi.org/10.1016/j.gca.2008.04.020Google Scholar
Tanwar, K., Petitto, S., Ghose, S., Eng, P., & Trainor, T. (2009). Fe(II) adsorption on hematite (0001). Geochimica et Cosmochimica Acta, 73(15), 43464365. https://doi.org/10.1016/j.gca2009.04.024Google Scholar
Teng, H., Fenter, P., Cheng, L., & Sturchio, N. (2001). Resolving orthoclase dissolution processes with atomic force microscopy and X-ray reflectivity. Geochimica et Cosmochimica Acta, 65(20), 34593474. https://doi.org/10.1016/S0016-7037(01)00665-2Google Scholar
Trainor, T., Eng, P., Brown, G., Robinson, I., & De Santis, M. (2002). Crystal truncation rod diffraction study of the alpha-Al2O3(1-10 2) surface. Surface Science, 496(3), 238250. https://doi.org/10.1016/S0039-6028(01)01617-XGoogle Scholar
Trainor, T., Chaka, A., Eng, P., Newville, M., Waychunas, G., Catalano, J., & Brown, G. (2004). Structure and reactivity of the hydrated hematite (0001) surface. Surface Science, 573(2), 204224. https://doi.org/10.1016/j.susc.2004.09.040Google Scholar
Trainor, T., Templeton, A., & Eng, P. (2006). Structure and reactivity of environmental interfaces: Application of grazing angle X-ray spectroscopy and long-period X-ray standing waves. Journal of Electron Spectroscopy and Related Phenomena, 150(2–3), 66-85. https://doi.org/10.1016/j.elspec.2005.04.011Google Scholar
Vanek, A., Ettler, V., Grygar, T., Boruvka, L., Sebek, O., & Drabek, O. (2008). Combined chemical and mineralogical evidence for heavy metal binding in mining- and smelting-affected alluvial soils. Pedosphere, 18(4), 464478. https://doi.org/10.1016/s1002-0160(08)60037-5Google Scholar
Waychunas, G., Trainor, T., Eng, P., Catalano, J., Brown, G., Davis, J., Rogers, J., & Bargar, J. (2005). Surface complexation studied via combined grazing-incidence EXAFS and surface diffraction: arsenate an hematite (0001) and (10-12). Analytical and Bioanalytical Chemistry, 383(1), 1227. https://doi.org/10.1007/s00216-005-3393-zGoogle Scholar
Williams, G. J., Pfeifer, M. A., Vartanyants, I. A., & Robinson, I. K. (2006). Internal structure in small Au crystals resolved by three-dimensional inversion of coherent x-ray diffraction. Physical Review B, 73(9), Article 094112. https://doi.org/10.1103/PhysRevB.73.094112Google Scholar
Xu, T. Y., Stubbs, J. E., Eng, P. J., & Catalano, J. G. (2018). Response of interfacial water to arsenate adsorption on corundum (001) surfaces: Effects of pH and adsorbate surface coverage. Geochimica et Cosmochimica Acta, 239, 198212. https://doi.org/10.1016/j.gca.2018.07.041Google Scholar
Xu, T., Stubbs, J., Eng, P., & Catalano, J. (2019). Comparative response of interfacial water structure to pH variations and arsenate adsorption on corundum (012) and (001) surfaces. Geochimica et Cosmochimica Acta, 246, 406418. https://doi.org/10.1016/j.gca.2018.12.006Google Scholar
Yan, H., Park, C., Ahn, G., Hong, S., Keane, D., Kenney-Benson, C., Chow, P., Xiao, Y., & Shen, G. (2014). Termination and hydration of forsteritic olivine (010) surface. Geochimica et Cosmochimica Acta, 145, 268280. https://doi.org/10.1016/j.gca.2014.09.005Google Scholar
Yuan, K., Bracco, J., Schmidt, M., Soderholm, L., Fenter, P., & Lee, S. (2019a). Effect of anions on the changes in the structure and adsorption mechanism of zirconium species at the muscovite (001)-water interface. Journal of Physical Chemistry C, 123(27), 1669916710. https://doi.org/10.1021/acs.jpcc.9b02894Google Scholar
Yuan, K., Lee, S. S., Cha, W., Ulvestad, A., Kim, H., Abdilla, B., Sturchio, N., & Fenter, P. (2019b). Oxidation induced strain and defects in magnetite crystals. Nature Communications, 10, 10, Article 703. https://doi.org/10.1038/s41467-019-08470-0Google Scholar
Zhang, Z., Fenter, P., Cheng, L., Sturchio, N., Bedzyk, M., Predota, M., & 11 others (2004). Ion adsorption at the rutile-water interface: Linking molecular and macroscopic properties. Langmuir, 20(12), 49544969. https://doi.org/10.1021/la0353834Google Scholar
Zhang, Z., Fenter, P., Kelly, S., Catalano, J., Bandura, A., Kubicki, J., & 5 others (2006). Structure of hydrated Zn2+ at the rutile TiO2 (110)-aqueous solution interface: Comparison of X-ray standing wave, X-ray absorption spectroscopy, and density functional theory results. Geochimica et Cosmochimica Acta, 70(16), 40394056. https://doi.org/10.1016/j.gca.2006.06.325Google Scholar
Zhang, Z., Fenter, P., Sturchio, N., Bedzyk, M., Machesky, M., & Wesolowski, D. (2007). Structure of rutile TiO2 (110) in water and 1 molal Rb+ at pH 12: Inter-relationship among surface charge, interfacial hydration structure, and substrate structural displacements. Surface Science, 601(4), 11291143. https://doi.org/10.1016/j.susc.2006.12.007Google Scholar
Supplementary material: File

Stubbs et al. supplementary material
Download undefined(File)
File 663.6 KB