Efecto del Cl y Fe+3 en FeS2

Embed Size (px)

Citation preview

  • 8/9/2019 Efecto del Cl y Fe+3 en FeS2

    1/4

    Pergamon

    Geochimica et Cosmochimica Acta. Vol. 59, No. 15, pp. 3155-3158, 1995

    Copyright 0 1995 Elsevier cience Ltd

    Printed n the USA.

    All rights reserved

    00 I6-7037/95 9.50 + . I0

    0016-7037 95)00203-O

    Confirmation of a sulfur-rich layer on pyrite after oxidative dissolution

    by Fe II1) ions around pH 2

    KEIKO SASAKI,

    MASAMI TSUNEKAWA,

    TOSHIAKI

    OHTSUK A

    and

    HIDETAKA

    KONNO

    Faculty of Engineering, Hokkaido University, Sapporo 060, Japan

    *Department of Applied Chemistry, Nagoya Institute of Technology, Nagoya 466, Japan

    (Received December 29, 1994; accepted in revised form Apri l 24, 1995)

    Abstract-The

    stoichiometry of pyrite dissolution by Fe( III) ions was studied in chloride media aroun d

    pH 2. Pyrite wa s found to dissolve nonstoichiometrically during th e initial tens of hours and a S-rich layer

    was formed on the pyrite due to preferential dissolution of iron. The major constituent of the layer was

    elemental S, identified by X-ray photoelectron spectroscopy (XPS) and Raman spectroscopy.

    1. INTRODUCTION

    Pyrite is abundant in nature and the most common sulfide

    mineral. The oxidation of pyrite in aqueou s systems is closely

    related to environmental and geological problem s, like the

    formation of acid mine drainage, th e supergen e alteration o f

    ore deposits, and others. The important oxidants in the natural

    environment are Fe(II1) ions and oxygen, the forme r being

    more reactive than the latter (e.g., McKibben and Barnes,

    1986 ). Th e overall reaction of pyrite w ith the Fe(II1) ion is

    conventionally expre ssed as:

    Fe & + 1 4Fe+ + 8H20 = 15 Fe*+ + 2SOi- + 16 Hf.

    (1)

    The kinetics for the above reaction around p H 2 have been

    studied by measuring: (a) the total dissolved Fe ions and

    redox potential, E (Smith and Shumate, 1970; Rimstidt and

    Newcomb , 1993); (b) Fe( II) ions (McKibben and Barnes,

    1986);

    (c)

    Fe( III) ions and E (Wiersm a and Rimstidt, 1984) ;

    (d) Fe( III) ions (Mathew s and Robbins, 1972) ; (e) E (Gar-

    rels and Thompson, 1960), and (f) dissolved S species

    (Moses et al., 1987). Only a few papers presented experi-

    mental dissolution curves fo r both Fe and S species measu red

    simultaneously. The above dissolution schem e, Eqn. 1, is gen-

    erally acce pted as correct, but it has not been possible to locate

    stoichiometric dissolution data. Some results suggest nonstoi-

    chiometric dissolution of pyrite on the basis of surface anal-

    ysis (Buckley and Woods, 1987) or electrochemical mea-

    surements (My croft et al., 1990), and an analysis of the dis-

    solution data reported by McKibben and Barnes ( 1986) may

    also suggest nonstoichiometric dissolution. In a previous pa -

    per, we repor ted that during nonstoichiometric pyrite disso-

    lution by oxygen, Fe species dissolved preferentially leaving

    S species on the surface (Sasaki, 1994).

    To understand better the mechanism of pyrite oxidation, it

    is important to know w hether py rite dissolves stoichiometri-

    tally or nonstoichiometrically (Luther , 1987; Mo ses et al.,

    1987 ). In the present w ork, oxidative dissolution of pyrite by

    Fe( III) ions in chloride solutions is investigated. The stoichi-

    ometry is discussed based on dissolution data for Fe and S

    species measu red simultaneously and the results of XP S and

    Raman spectroscopy.

    2.1. Pyrite

    2.

    EXPERIMENTAL

    Pyrite was supplied from the Yanahara mines (Japan) and its com-

    position was listed in Table 1. The major impurity is SiOz and its

    3155

    mass percent was estimated to be around 0.25 in the sample. The

    pyrite was ground using a tungsten carbide planetary-type ball mill

    (Fritsch Co. Ltd., P-5) and sieved. The 200-400 mesh size fraction

    was used in the experiments. The X-ray powder diffraction lines were

    measured with NaCl as the standard and were found to be identical

    with that of FeS, (Sasaki et al., 1994a). There were no diffraction

    lines other than FeS, except a very weak single line of SiO* (d

    = 0.338 nm), which took no part in the dissolution experiments (Sa-

    saki, 1994).

    The samples were pretreated to establish a well-defined surface of

    stoichiometric composition. Oxidized surface species were removed

    according to a previous method (McKibben and Barnes, 1986; Sasaki

    et al., 1994b): the pyrite was ultrasonically cleaned in ethanol for 30 s

    with adhering mineral powder decanted, and then the pyrite was

    rinsed with 1 mol/L HNO, for one mitt, followed by rinsing in triply

    distilled water, before dehydrating in acetone, and finally it was vac-

    uum-dried using an aspirator. The pretreated surface was compared

    with a cleaved one and reproducibly verified to be stoichiometric by

    XPS (Sasaki et al., 1994b).

    The specific surface area of the pretreated pyrite was determined

    by the NZ gas adsorption BET method on a Quantasorb QS- 13 (Yuasa

    Ionics Co. Ltd.) with a QS-300 cell for low value measurements.

    Each result was in good agreement with the data obtained by the Nz

    gas adsorption seven points BET method on an Autosorb AS-l

    (Yuasa Ionics Co. Ltd.). The average of twelve experiments was 0.40

    ? 0.060 (~frlcr,,, n = 12) mlg.

    2.2.

    Dissolution Experiments

    Dissolution experiments were carried out using a batch reactor

    with an impeller and a G2 glass ball filter for Nz bubbling. The so-

    lution was sufficiently stirred to suspend pyrite particles. All assem-

    blies were made of glass and rinsed with a diluted HNO? solution

    before use to avoid Fe contamination. As the oxidant, 15 mmol/L

    FeCl

    6Hz0 was used. The initial pH was adjusted to 2.0 with super

    special grade (SSG) HCl. To avoid comnlications bv a two oxidant

    .

    system,dissolved oxygen was removed to the extent that the oxida-

    tion of pyrite and dissolved Fe(B) ions by dissolved oxygen was

    insignificant. Prior to each dissolution experiment, 200 cm of each

    solution was de-aerated by bubbling Nz gas (299.999 ) for 1 h: the

    dissolved oxygen in the de-aerated solution was under 1 ppm. Then,

    2.00 g of the just pretreated pyrite powder was added to the solution.

    The dissolution experiment was carried out for 72 h under Nz bub-

    bling at room temperature (not controlled). Solution pH was also not

    controlled, since pH does not affect the dissolution rate in the range

    between 0 and 2 (Wiersma and Rimstidt, 1984). At appropriate in-

    tervals, 1 cm of solution was sampled and filtrated with a 0.20 pm

    pore membrane filter for solution analysis, and pH and E measure-

    ments. The total Fe and Fe(B) concentrations were determined by

    the 1, IO-phenanthroline method and the total S concentration by

    ICP-AES (SEIKO Co. Ltd., SPS 1200). Dilution of the sampled

    solution was with a 0.01 mol/L SSG HCl solution. After the disso-

    lution experiment, the residue was separated by filtration and stored

    in a Nz-purged desiccator for XPS and Raman spectroscopic analysis.

  • 8/9/2019 Efecto del Cl y Fe+3 en FeS2

    2/4

    3 56

    K. Sasaki et al.

    Table 1 Composition (mg gt) of the pyrite from the Yanahara mines (Japan)*

    Na Al

    Si P K Ca h4n Fe Cu Zn

    Unkown residues

    0.046 1.38 12.0 0 .806 1.05 1.36

    0.055 458 2.10 2.22 520.983

    * Owina to the decomposition of ovrite. an accurate value of sulfur was not obtained

    (Naka&ra, 1991). - -

    2.3. XPS

    Measurement

    The chemical species on the surface of the pyrite were analyzed

    by XPS, with an ESCA LAB MkII (V. G. Scientific). The samples

    were pressed o nto Cu foil on a holder and introduced into the spec-

    trometer. After evacuating to better than 10 - Pa within 15 min, the

    sample was transferred into an analyzer chamb er of better than 5

    X lo- Pa and cooled to below 150 K, then irradiated with Al Ka

    X-ray ( 15 kV, 10 mA). The binding energies, Ear were calibrated

    with &[Au 4hi2] = 84.0 eV. Intensity by area was measured after

    drawing the background by the Shirley method ( 1972 ). Details of

    the data analysis are described elsewhere (Konno et al., 199 I ),

    2.4. Raman Spectroscopy

    For Raman spectroscopy, excitation w as accomplished by a single

    line 488. 0 nm wavelength light from an Ar ion laser. The power of

    incidence was about 50 mW at the sample point. The Raman scattered

    light was detected by a triple type monochromator (JASCO R-NO T)

    equipped with a photomultiplier (Hama matsu R-464) and a photon

    counter. Th e band path of the monochromator was about 8 cm- The

    pyrite sample was diluted to 5 wt% with KBr powder and 0.30 g of

    the mixture was compressed to form a 10 mm ~#~isk for measurement

    by a rotating a pparatus. There is a possibility that heating by laser

    irradiation causes some change in spectral quality, so that the above

    measures were taken. The laser light was polarized parallel to the

    plane of incidence and the incidence angle was 80 degrees. The Ra-

    man scattered light was collected in the incidence plane in the direc-

    tion normal to the incident laser light.

    3. RESULTS

    AND DISCUSSION

    Typical dissolution curves of pyrite in 0.01 mol/L HCl

    containing 15 mmo l/L Fe( III) ions are shown in Fig. 1 . From

    the net amount of dissolved iron in Fig. 1, the average thick-

    ness of the pyrite dissolved is calculated to be 6.2 nm. It

    corresponds to an average of more than 10 lattice layers of

    pyrite crystal, thoug h uniform dissolution actually does not

    take place in acid solutions (M cKibben and Barnes, 1986 ).

    The total amount of species

    i

    in the solution per unit initial

    surface area of pyrite at time

    t

    is expressed as xi

    (t)

    in mm01

    mm*, where

    i

    indicates S, Fe( II), or Fe( III). Assuming that

    pyrite is dissolved with Fe( III) ions accordin g to Eqn. 1, the

    _

    20 40 60 80

    Time,

    t /

    h

    FIG. 1. Dissolution curves of pyrite by Fe(II1) ions in a HCI soh-

    tion at pH 2.0.

    stoichiometry at t, n(t)

    = xs ( t) lxFe t)

    , can be exp ressed the-

    oretically using x~~;,(, (

    t)

    and the initial amount of Fe( III) ions

    per unit initial surface area of pyrite, xFe(rll( 0), as:

    nthco(f) = xS(t)kk,Ill)(f) + - c(ll,(t))

    = 2xFe~l~(~)~o/(1~xFe~ll~

    (0) + -+edt)). (2)

    In Fig. 2, the theoretical stoichiometric ratio, &w (t), is com-

    pared with the observed stoichiometric ratio, nabs(t), obtained

    from the experimental data, showing that nthco(

    )

    is larger than

    nobs( ) , and that the difference between &h._(t) and noha t)

    increases with time. This indicates that Fe dissolves more eas-

    ily than sulfur from pyrite. The values of &.o( t) can also be

    calculated by using x,(t) or xFe

    t)

    , and though the values are

    less accurate than with

    +(,,)(t),

    the results were similar to

    Fig. 2.

    The XP-spectra for (a) the cleaved pyrite, (b) the pre-

    treated pyrite, and (c) the sample after dissolution experiment

    are shown in Figs. 3 and 4. The surface com position, [Fe],/

    [S], mole ratios, was calculated according to Eqn. 3:

    [FeldlISh = (kJ~Fe)I(~s~ss),

    (3)

    whe re s is a relative sensitivity factor including an escap e

    depth of electrons, and

    I

    an intensity by area in which contri-

    bution from X-ray satellite peaks was subtracted by computer

    calculation. The relative sensitivity factors were experimen-

    tally determined and ss/sFc =

    0.14 + 0.00, ( + 1c)

    ;

    the relative

    standard deviation is 5% (Konno et al., 1991; Sasaki et al.,

    199413). The compositions, [Feh/[ S],, of both the cleaved

    surface and pretreated one were determined to be 0.50. After

    72 hour dissolution by Fe( III) ions, the Fe 2p spectrum was

    different from that of pure pyrite as shown in Fig. 4, where a

    broken line indicates th e envelope for pure pyrite. The S 2p

    spectrum (c) was separated into three components, each com-

    posed of a set of 2p,,, and 2p,,, peaks (Fig. 3). Pea k (I) is

    assigned to pure pyrite at &[ S 2p,,,] = 162.2 eV, peak (II)

    at Es[S 2~?,~] = 164.25 eV to elemental S, and peak (III) at

    &[ S 2p,,,] = 166.4 eV to oxidized S species such as sulfite.

    Time,

    t /

    h

    FIG.

    2. Theoretical and observed ratios, n(t) = n,(t)lx,,(t).

  • 8/9/2019 Efecto del Cl y Fe+3 en FeS2

    3/4

    Stoichiometry of pyrite dissolution 315 7

    I

    I

    s 2P

    L

    L

    a>

    160 170

    Binding energy, I?& eV

    FIG. 3. The XP-spectra of S 2p for pyrite. Vertical bars indicate

    500 cps. (a) cleaved pyrite, (b) pretreated pyrite and (c) after disso-

    lution by Fe(III) ions. Component (I) is pyrite S, (II) elemental S,

    and (III) oxidized S species.

    The mole ratio, [Fe]r/ [ S], , after dissolution was determined

    to be 0.34, excluding the oxidized components at peak (III)

    in Fig. 3 and arrows in Fig. 4. It indicates that nonstoichiomet-

    tic dissolution resulted in the sulfur-rich layer on the surface,

    in agreement with the dissolution data. The accuracy of the

    semiquantitative analysis by XPS is not very high, but the

    difference established here is much larger than the relative

    standard deviation of relative sensitivity factors, as indicated

    above. Distinguishing among sulfide, elemental sulfur, and

    polysulfide is difficult only by XP-sp ectra, as the binding en-

    ergy,

    EB

    for pyrite is 162.2 -163.0 eV (e.g., Brian, 1980 ;

    Konno et al., 1991; Peisert et al., 1994; Nesbitt and Muir,

    1994), Es for elemental S is 163.7 -164.3 eV (e.g., Brion,

    1980 ; Hyland and Bancroft, 1989 ; Peisert et al., 1994) , a nd

    EB or polysulfides is 161.7- 163.4 eV (e.g., Buckley et al.,

    1988).

    Hamilton and Woods (1980) and Buckley and Woods

    ( 198 7) reported that a metal-deficient sulfide was formed by

    electrochemical oxidation of sulfides in acidic solutions (pH

    4.6), and polysulfide was detected by XPS, mass spectros-

    copy, and voltammetry. Mycroft et al. (1990 ) reported that

    after anodic oxidation at 700 mV (SCE) in near-neutral aque-

    ous chloride solutions for more than one hour, polysulfide was

    detected on pyrite by Rama n spectroscopy. The Rama n spec-

    trum was measured to elucidate further the composition of the

    S-rich layer. As shown in Fig. 5a, the Ram an spectrum of the

    same sample with (c) in Figs. 3 and 4 has five Raman bands:

    at 150, 220, 357, 387, and 470 cm-. The three of them at

    150, 220, and 470 cm- are assigned to S-S vibration modes

    Fe 2p

    h

    oxidized species

    700

    710

    720

    730

    Binding energy, &I eV

    FIG. 4. The XP-spectra of Fe 2p for pyrite. Vertical bars indicate

    1 kcps. (a) cleaved pyrite, (b) pretreated pyrite and (c) after disso-

    lution by Fe(II1) ions. A profile by dashed line in (c) is for pure pyrite.

    of elemental S and the remaining two, at 353 and 387 cm-,

    to vibration modes of pyrite, agreeing with previous studies

    (Mycroft et al., 1990 ; Mem agh and Truda , 1993 ; Li et al.,

    1993 ). According to Buckley and Woods ( 1987 ), elemental

    S dissolves moderately in cyclohexane, but polysulfide does

    not. After the sample was immersed in cyclohexane for five

    minutes, the three bands at 150, 220, and 470 cm-, corre-

    sponding to the vibration modes of the elemental S disap-

    peared, as shown in Fig. 5b. This is an additional evidence

    600

    400

    200

    Raman shift, Av

    I

    cm-

    FIG. 5. The Raman spectra for pyrite. A vertical bar indicates a

    scattering light intensity of 1 kcps. (a) after dissolution by Fe(II1)

    ions and (b) after immersion of the oxidized pyrite in cyclohexane

    for five minutes.

  • 8/9/2019 Efecto del Cl y Fe+3 en FeS2

    4/4

    3158

    K. Sasaki et al.

    that the S-rich layer form ed on pyrite is composed of elemen-

    tal S and not of polysulfide. There seems to be a weak Raman

    band around 460 cm- in Fig. 5a, but it disappe ared after

    immersion in cyclohexan e, as shown in Fig. 5b. It indicates

    that the band was not due to polysulfide but a noise. Thu s, it

    is concluded that the reaction of pyrite with Fe(III) ions in

    chloride solutions around pH 2 proce eds non stoichiometri-

    tally during t he initial tens of hours, resulting in the formation

    of an elemental S-rich lay er on the pyrite.

    As described above and also reported previously (Sasaki,

    1994; Sasaki et al., 1994b), experimental results sho wed that

    dissolved amounts of sulfur species were lacking O-50 %

    compared with the stoichiometric values. According to the

    oxidation mechanism of pyrite by Fe(II1) ions based on the

    molecular orbital theory (Luthe r, 1987 ), the reaction is ini-

    tiated by attack of Fe( III) ions to S s- and finally one pyrite

    molecule is released as S,O:- and Fe+ ions. This, how ever,

    is not consistent with the above results. F or the present, it is

    uncertain that the prop osed m echanism can explain the overall

    oxidation reaction of pyrite by Fe(II1) ions. Further investi-

    gation is necessary considering recent reports on the detailed

    band structure of Fe& and related sulfides (Sato, 1985; Tem-

    merman et al., 1993).

    In the natural environment, nonstoichiometric dissolution

    is not considered to continue indefinitely. The formation of

    an elemental S-rich layer may reduce the dissolution rate and

    finally stop the dissolution, or oxidation of the S-rich layer

    may occur by proce sses like those by S-oxidizing bacteria

    such as Thiobacillus thiooxidans. It has been considere d that

    pyrite weathering is enhanced mainly by Thiobacillus fer-

    rooxidans, which can oxidize Fe(I1) ions in acidic environ-

    ments. M uch attention has not been paid to T. thiooxidans,

    since it mainly oxidizes elemental sulfur to sulfate but rarely

    does sulfide to sulfate, with no ability to oxidize Fe( II) ions

    to Fe(II1) ions. The present data, howeve r, su ggests th at T.

    thiooxidans

    might also play an important role in pyrite we ath-

    ering in coexistence with T. errooxidans. Synergetic effect

    by two species o f bacteria on the pyrite w eathering is an in-

    teresting subject and under investigation in our laboratory.

    Acknowledgments-We wish to express appreciation to Fritsch Japan

    Co. Ltd. for carrying out the grinding experiments and to the Hok-

    kaido Industrial Research Institute where the ICP-AES measurements

    were carried out by courtesy of Mr. K. To&a. The authors also thank

    reviewers, Dr. Kevin Rosso, Dr. Allen Pratt, and the anonymous third

    person, for their critical and constructive review.

    Editotial handling: M. F. Hochella Jr.

    REFERENCES

    Brion D. ( 1980) Etude par spectroscopic de photoelectrons de deg-

    radation superficielle de Fe&,

    CuFe

    ZnS et PbS a pair et dam

    Ieau. Appl. .%I Sci. 5, 133-152.

    Buckley A. N. and Woods R. ( 1987) The surface oxidation of pyrite.

    Appl. SurJ Sci. 27,437-452.

    Buckley A. N., Wouterlood H. J., Cartwright P. S., and Gillard R. D.

    ( 1988) Core electron binding energies of platinum and rhodium

    polysulfides.

    Inorg. Chim. Acta

    143,77-80.

    Garrels R. M. and Thompson M. E. ( 1960) Oxidation of pyrite by

    iron sulfate solutions. Amer. J . Sci. 258-A, 57-67.

    Hamilton I. C. and Woods R. (1980) An investigation of surface

    oxidation of pyrite and pyrrhotite by linear potential sweep vol-

    tammetry. J. Electroanal. Chem.

    118 327-343.

    Hyland M. M. and Bancroft G. M. ( 1989) An XPS study of gold

    deposition at low temperatures on sulfide minerals: Reducing

    agents.

    Geochim. Cosmochim. Acta 53,367-372.

    Konno H., Sasaki K., Tsunekawa M., Takamori T., and Furuichi R.

    ( 1991) X-ray photoelectron spectroscopic analysis of surface

    products on pyrite formed by bacterial leaching.

    Bunseki Kagaku

    40,609-616.

    Li J., Zhu X., and Wadsworth M. E. (1993) Raman spectroscopy of

    natural and oxidized metal sulfides. EPD Gong. (ed. J. P. Hager),

    229-244.

    Luther G. W., III ( 1987) Pyrite oxidation and reduction: Molecular

    orbital theory considerations. Geochim. Cosmochim. Acra

    51

    3193-3199.

    Mathews C. T. and Robbins R. G. (1972) The oxidation of iron

    disulfide by ferric sulphate. Amt. Chem. Eng. Aug., 21-25.

    McKibben M. A. and Barnes H. L. ( 1986) Oxidation of pyrite in low

    temperature acidic solutions: Rate laws and surface textures. Geo-

    chim. Cosmochim. Acta 50, 1509-1520.

    Memagh T. P. and Trudu A. G. ( 1993) A laser Raman microprobe

    study of some geologically important sulphide minerals. Chem.

    Geol. 103, 113-127.

    Moses C. 0.. Nordstrom D. K., Herman J. S., and Mills A. L. ( 1987)

    Aqueous pyrite oxidation by dissolved oxygen and by ferric iron.

    Geochim. Cosmochim. Acta

    51

    1561- 157 1.

    Mycroft J. R., Mcintyre N. S., Lorimer J. W., and Hill I. R. ( 1990)

    Detection of sulfur and polysulphides on electrochemically oxi-

    dized pyrite surfaces by X-ray photoelectron spectroscopy and Ra-

    man spectroscopy. J. Electroanal . Chem. 292, 139- 152.

    Nakamura Y. (1991) Mineral. Bunseki 8,613-614.

    Nesbitt H. W. and Muir I. J. ( 1994) X-ray photoelectron spectro-

    scopic study of a pristine pyrite surface reacted with water vapour

    and air. Geochim. Cosmochim. Acta 58,4667-4679.

    Peisert H., Chasse T., Streubel P., Meisel A., and Szargan R. ( 1994)

    Relaxation energies in XPS and XAES of solid sulfur compounds.

    J. Electron Spectrosc. Relat. Phenom. 68, 321-328.

    Rimstidt J. D. and Newcomb W. D. ( 1993) Measurement and anal-

    ysis of rate data: The rate of ferric iron with pyrite. Geochim.

    Cosmochim. Acta 57, 1919-1934.

    Sasaki K. ( 1994) Effect of grinding on the rate of oxidation of pyrite

    by oxygen in acid solutions. Geochim. Cosmochim. Acta 58,4649-

    4655.

    Sasaki K., Konno H., and Inagaki M. (1994a) Structural strain in

    pyrite evaluated by X-ray powder diffraction, J. Mater. Sci. 29,

    1666-1669.

    Sasaki K., Tsunekawa M., and Konno H. ( 1994b) Nonstoichiometry

    in the oxidative dissolution of pyrite in acid solutions. Bunseki

    Kagaku43,911-917.

    Sato K. ( 1985) Pyrite type compounds-with particular reference to

    optical characterization.

    Proc. Crystal Growth

    and Charact. 11

    109-154.

    Shirley D. A. ( 1972) High resolution X-ray photoemission spectrum

    of the valence bands of gold.

    Phys. Rev.

    B5,4709-4714.

    Smith E. E. and Shumate K. S. ( 1970) Sulfide to Sulfate Reaction

    Mechanism; Water Poll ut. Con trol Res. Ser. Rept. 14010 FPS

    02170, pp.

    I - 115,

    US Dept. Interior.

    Temmerman W. M., Durham P. J., and Vaughan D. J. ( 1993) The

    electronic structures of the pyrite-type disulfides (MS*, where M

    = Mn, Fe, Co, Ni, Cu, Zn) and the bulk properties of pyrite from

    local density approximation (LDA) band structure calculations.

    Phys. Chem. M ineral s 20,248-254.

    Wiersma C. L. and Rimstidt J. D. ( 1984) Rates of reaction of pyrite

    and marcasite with ferric iron at pH 2. Geochim. Cosmochim. Acta

    48,85-92.