Hostname: page-component-77c89778f8-7drxs Total loading time: 0 Render date: 2024-07-23T23:04:54.633Z Has data issue: false hasContentIssue false

Investigation of the electronic and geometric structures of the (110) surfaces of arsenopyrite (FeAsS) and enargite (Cu3AsS4)

Published online by Cambridge University Press:  05 July 2018

C. L. Corkhill*
Affiliation:
School of Earth, Atmospheric and Environmental Sciences, and the Williamson Research Centre for Molecular Environmental Science, The University of Manchester M13 9PL, UK
M. C. Warren
Affiliation:
School of Earth, Atmospheric and Environmental Sciences, and the Williamson Research Centre for Molecular Environmental Science, The University of Manchester M13 9PL, UK FACETS, University of Oxford, Keble Road, Oxford OX1 3RH, UK
D. J. Vaughan
Affiliation:
School of Earth, Atmospheric and Environmental Sciences, and the Williamson Research Centre for Molecular Environmental Science, The University of Manchester M13 9PL, UK

Abstract

The (110) surfaces of arsenopyrite (FeAsS) and enargite (Cu3AsS4) have been modelled using a Density Functional Theory (DFT) plane-wave pseudopotential method (CASTEP) in order to better understand aspects of the geometric and electronic structures of these minerals, which have important implications for the release of arsenic in acid mine drainage environments. In this study, bulk calculations of these minerals have been conducted to give consistent geometries for surface models and to establish reference states for changes at the surfaces in these models. Surface structure experimental data for enargite were collected using low-energy electron diffraction which confirmed an unreconstructed 161 (110) lattice, with a and b values of 6.38±0.3 Å and 9.92±0.5 Å, respectively. Surface calculations demonstrate geometric and electronic relaxation of both arsenopyrite and enargite (110) surfaces. Changes in atomic positions, interatomic distances, Mulliken charges and electronic configurations are reported. Enargite has a surface energy of ∼0.02 eV/Å2 compared with arsenopyrite which has a surface energy of ∼0.11 eV/Å2, indicating that the enargite (110) surface is more energetically stable than that of arsenopyrite. The most stable surfaces are those which relax to restore the surface coordination and partial charge balance. For both minerals this is achieved by the formation of covalent bonds. Arsenic has the most positive Mulliken charge of all the surface atoms and is, therefore, predicted to be the most reactive atom at the arsenopyrite and enargite (110) surfaces. This implies that, according to these calculations, arsenic is most likely to react with oxidative species such as O2 and H2O in environments such as those associated with acid mine drainage, potentially releasing oxides and acids of arsenic into the environment.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2011

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

Arribas, A., Jr. (1995) Characteristics of high sulfidation epithermal deposits and their relation to magmatic fluid. Pp. 419454 in: Magmas, Fluids and Ore Deposits (Thompson, J.F.H., editor). Short Course, 23, Mineralogical Association of Canada, Ottawa.Google Scholar
Blanchard, M., Alfredsson, M., Brodholt, J., Wright, K. and Catlow, C.R.A. (2007) Arsenic incorporation in FeS2 pyrite and its influence on dissolution: A DFT study. Geochimica et Cosmochimica Acta, 71, 624630.CrossRefGoogle Scholar
Buckley, A.N. and Walker, G.W. (1988–89) The surface composition of arsenopyrite exposed to oxidising environments. Applied Surface Science, 35, 227240.CrossRefGoogle Scholar
Clark, D.A. and Norris, P.R. (1996) Oxidation of mineral sulfides by thermophilic organisms. Minerals Engineering, 9, 11191125.CrossRefGoogle Scholar
Cordova, R., Gomez, H., Real, S.G., Schrebler, R. and Vilche, J.R. (1997) Characterisation of natural enargite/aqueous solution systems by electrochemical techniques. Journal of the Electrochemical Society, 144, 26282636.CrossRefGoogle Scholar
Corkhill, C.L. and Vaughan, D.J. (2009) Arsenopyrite oxidation: A review. Applied Geochemistry, 24, 23422361.CrossRefGoogle Scholar
Corkhill, C.L., Wincott, P.L., Lloyd, J.R. and Vaughan, D.J. (2008) The oxidative dissolution of arsenopyrite (FeAsS) and enargite (Cu3AsS4) by Leptospirillum ferrooxidans. Geochimica et Cosmochimica Acta, 72, 56165633.CrossRefGoogle Scholar
Cruz, R., Lazaro, I., Rodriguez, J.M., Monroy, M. and Gonzalez, I. (1997) Surface characterisation of arsenopyrite in acidic medium by triangular scan voltammetry on carbon paste electrodes. Hydrometallurgy, 46, 303319.CrossRefGoogle Scholar
de Leeuw, N.H., Parker, S.C., Sithole, H.M. and Ngoepe, P.E. (2000) Modelling the surface structure and reactivity of pyrite: Introducing a potential model for FeS2. Journal of Physical Chemistry B, 104, 79697976.CrossRefGoogle Scholar
Ehrlich, H.L. (1964) Bacterial oxidation of arsenopyrite and enargite. Economic Geology, 59, 1306 – 1312.CrossRefGoogle Scholar
Eyert, V., Hock, H.-K., Fiechter, S. and Tributsch, H. (1998) Electronic structure of FeS2: The crucial role of electron-lattice interaction. Physics Review B, 57, 63506359.CrossRefGoogle Scholar
Fantauzzi, M., Atzei, D., Elsener, B., Lattanzi, P.F., and Rossi, A. (2006) XPS and XAES analysis of copper, arsenic and sulfur chemical state in enargites. Surface and Interface Analysis, 38, 922930.CrossRefGoogle Scholar
Henao, J.A., Diaz de Delgado, G., de Delgado, J.M., Castrillo, F.J. and Odreman, O. (1994) Single-crystal structure refinement of enargite (Cu3AsS4). Materials Research Bulletin, 29, 11211127.CrossRefGoogle Scholar
Lattanzi, P.F., Da Pelo, S., Musu, E., Atzei, D., Elsener, B., Fantauzzi, M. and Rossi, A. (2008) Enargite oxidation: A review. Earth-Science Reviews, 86, 6288.CrossRefGoogle Scholar
Lutz, H.D. and Zminscher, J. (1996) Lattice dynamics of pyrite (FeS2) polarizable-ion model. Physics and Chemistry of Minerals, 23, 497502.CrossRefGoogle Scholar
Monkhorst, H.J. and Pack, J.D. (1976) Special points for Brillouin-zone integrations. Physical Review B, 13, 51885192.CrossRefGoogle Scholar
Morimoto, N. and Clark, L.A. (1961) Arsenopyrite crystal-chemical relations. American Mineralogist, 46, 14481496.Google Scholar
Mosselmans, J.F.W., Pattrick, R.A.D., van der Laan, G., Charnock, J.M., Vaughan, D.J., Henderson, C.M.B. and Garner, C.D. (1995) X-ray absorption near-edge spectra of transition metal disulfides FeS2 (pyrite and marcasite), CoS2, NiS2 and CuS2, and their isomorphs FeAsS and CoAsS. Physics and Chemistry of Minerals, 22, 311317.CrossRefGoogle Scholar
Nesbitt, H.W. and Muir, I.J. (1998) Oxidation states and speciation of secondary products on pyrite and arsenopyrite reacted with mine waste waters and air. Mineralogy and Petrology, 62, 123144.CrossRefGoogle Scholar
Nesbitt, H.W., Muir, I.J. and Pratt, A.R. (1995) Oxidation of arsenopyrite by air and air-saturated, distilled water and implications for mechanisms of oxidation. Geochimica et Cosmochimica Acta, 59, 17731786.CrossRefGoogle Scholar
Payne, M.C., Teter, M.P., Allan, D.C., Arias, T.A. and Joannopoulos, J.D. (1992) Iterative minimisation techniques for ab initio total-energy calculations: molecular dynamics and conjugate gradients. Reviews of Modern Physics, 64, 10451097.CrossRefGoogle Scholar
Perdew, J.P., Burke, K. and Ernzerhof, M. (1996) Generalised gradient approximation made simple. Physical Review Letters, 77, 38653868.CrossRefGoogle Scholar
Pratt, A.R. (2004) Photoelectron core levels for enargite, Cu3AsS4 . Surface and Interface Analysis, 36, 654657.CrossRefGoogle Scholar
Probert, M.J. (2003) Improved algorithm for geometry optimisation using damped molecular dynamics. Journal of Computational Physics, 191, 130146.CrossRefGoogle Scholar
Pulay, P. (1980) Convergence acceleration of iterative sequences, the case of SCF iteration. Chemical Physics Letters, 73, 393393.CrossRefGoogle Scholar
Qui, G., Xiao, Q., Hu, Y., Qin, W. and Wang, D. (2004) Theoretical study of the surface energy and electronic structure of pyrite FeS2 (100) using a total-energy pseudopotential method, CASTEP. Journal of Colloid and Interface Science, 270, 127132.Google Scholar
Reich, M. and Becker, U. (2006) First-principles calculations of the thermodynamic mixing properties of arsenic incorporation into pyrite and marcasite. Chemical Geology, 225, 278290.CrossRefGoogle Scholar
Rimstidt, J.D. and Vaughan, D.J. (2003) Pyrite oxidation: A state-of-the-art assessment of the reaction mechanism. Geochimica et Cosmochimica Acta, 67, 873880.CrossRefGoogle Scholar
Rossi, A., Atzei, D., Da Pelo, S., Frau, F., Lattanzi, P.F., England, K.E.R. and Vaughan, D.J. (2001) Quantitative X-ray photoelectron spectroscopy study of enargite (Cu3AsS4) surface. Surface and Interface Analysis, 31, 465470.CrossRefGoogle Scholar
Rosso, K.M., Becker, U. and Hochella, M.F. (1999a) Atomically resolved structure of pyrite (100) surfaces: An experimental and theoretical investigation with implications for reactivity. American Mineralogist, 84, 15351548.CrossRefGoogle Scholar
Rosso, K.M., Becker, U. and Hochella, M.F. (1999b) The interaction of pyrite (100) surfaces with O2 and H2O: Fundamental oxidation mechanisms. American Mineralogist, 84, 15491561.CrossRefGoogle Scholar
Schaufuss, A.G., Nesbitt, H.W., Sciani, M.J., Hoecsht, H., Bancroft, M.G. and Szargan, R. (2000) Reactivity of surface sites on fractured arsenopyrite (FeAsS) toward oxygen. American Mineralogist, 85, 17541766.CrossRefGoogle Scholar
Segall, M.D., Pickard, C.J., Shah, R. and Payne, M.C. (1996) Population analysis of plane-wave electronic structure calculations of bulk materials. Physical Review B, 54, 16317.CrossRefGoogle ScholarPubMed
Segall, M.D., Lindan, P.J.D., Probert, M.J., Pickard, C.J., Hasnip, P.J., Clark, S.J. and Payne, M.C. (2002) First-principles simulation: ideas, illustrations and the CASTEP code. Journal of Physics: Condensed Matter, 14, 27172744.Google Scholar
Sithole, H.M., Ngoepe, P.E. and Wright, K. (2003) Atomistic simulation of the structure and elastic properties of pyrite (FeS2) as a function of pressure. Physics and Chemistry of Minerals, 30, 615619.CrossRefGoogle Scholar
Tossell, J.A. (1977) SCF-Xa scattered wave MO studies of the electronic structure of ferrous iron in octahedral coordination with sulfur. Journal of Chemical Physics, 66, 57125719.CrossRefGoogle Scholar
Vanderbilt, D. (1990) Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Physical Review B, 41, 78927895.CrossRefGoogle Scholar
von Oertzen, G.U., Skinner, W.M. and Nesbitt, H.W. (2005) Ab initio and x-ray photoemission spectroscopy study of the bulk and surface electronic structure of pyrite (100) with implications for reactivity. Physical Review B, 72, 235427.CrossRefGoogle Scholar
von Oertzen, G.U., Skinner, W.M. and Nesbitt, H.W. (2006) Ab initio and XPS studies of pyrite (100) surface states. Radiation Physics and Chemistry, 75, 18551860.CrossRefGoogle Scholar