quimica aplicada

15
Pressure oxidation of pyrite in sulfuric acid media: a kinetic study Hu Long, David G. Dixon * Department of Metals and Materials Engineering, University of British Columbia, 309-6350 Stores Road, Vancouver, BC, Canada, V6T 1Z4 Received 25 March 2002; received in revised form 15 July 2003; accepted 22 July 2003 Abstract The oxidation kinetics of a massive pyrite (FeS 2 ) sample from Zacatecas, Mexico were investigated in sulfuric acid solution under oxygen pressure. The effects of temperature (170 – 230jC), particle size (49 – 125 Am diameter), agitation speed (650 – 950 rpm), oxygen partial pressure (345 – 1035 kPa) and pulp density (1 – 20 g/L) were evaluated. The catalytic effect of Cu(II) was also observed. Fe(III) was found to be the initial product of pyrite oxidation, although the proportion of total dissolved iron as Fe(III) did reach a minimum during each test, indicating the generation and subsequent oxidation of reduced sulfur species. Pyrite oxidation kinetics are limited by the rate of reaction at the pyrite surface, with an activation energy of 33.2 kJ/mol (7.9 kcal/mol) with respect to dissolved oxygen concentration, or 41.7 kJ/mol (10.0 kcal/mol) with respect to oxygen partial pressure, over the temperature range 170 – 230jC. The reaction order with respect to oxygen partial pressure was found to be 0.5 at 210jC, indicating the first charge transfers of both the anodic and cathodic half-cell reactions to be the most likely rate-controlling steps. Conversion data conform to the shrinking sphere model initially, but deviate at higher conversions, indicating passivation of the mineral surface, most likely by elemental sulfur, which may precipitate via the disproportionation of other reduced sulfur species (e.g., thiosulfate) that form as intermediate products of pyrite oxidation. In order to account quantitatively for this passivation phenomenon, a new ‘‘passivating shrinking sphere’’ model is proposed which fits the conversion data precisely over the entire range. D 2003 Elsevier B.V. All rights reserved. Keywords: Pyrite; Pressure oxidation; Leaching; Kinetics; Passivation; Autoclave; Refractory gold; Hydrometallurgy 1. Introduction Pyrite (FeS 2 ) is the most common iron sulfide mineral. In most pyrite ores and concentrates, the pyrite itself is rarely of economic importance and is often viewed as a gangue mineral. However, it often influences the recovery of associated metal values such as gold, copper and zinc. In many sulfidic refractory gold ores and concentrates, gold is often found finely disseminated in sulfide minerals, par- ticularly in pyrite (Gasparrini, 1983), rendering the adequate liberation of gold impossible by grinding alone. One method for recovery of gold from these refractory gold ores and concentrates is pressure 0304-386X/$ - see front matter D 2003 Elsevier B.V. All rights reserved. doi:10.1016/j.hydromet.2003.07.010 * Corresponding author. Tel.: +1-604-822-3679; fax: +1-604- 822-3619. E-mail address: [email protected] (D.G. Dixon). www.elsevier.com/locate/hydromet Hydrometallurgy 73 (2004) 335 – 349

description

quimica

Transcript of quimica aplicada

  • ish Columbia, 309-6350 Stores Road, Vancouver, BC, Canada, V6T 1Z4

    species (e.g., thiosulfate) that form as intermediate products of pyrite oxidation. In order to account quantitatively for this

    (200D 2003 Elsevier B.V. All rights reserved.

    Keywords: Pyrite; Pressure oxidation; Leaching; Kinetics; Passivation; Autoclave; Refractory gold; Hydrometallurgy

    1. Introduction often viewed as a gangue mineral. However, it often

    influences the recovery of associated metal valuesthe entire range.passivation phenomenon, a new passivating shrinking sphere model is proposed which fits the conversion data precisely overReceived 25 March 2002; received in revised form 15 July 2003; accepted 22 July 2003

    Abstract

    The oxidation kinetics of a massive pyrite (FeS2) sample from Zacatecas, Mexico were investigated in sulfuric acid solution

    under oxygen pressure. The effects of temperature (170230jC), particle size (49125 Am diameter), agitation speed (650950 rpm), oxygen partial pressure (3451035 kPa) and pulp density (120 g/L) were evaluated. The catalytic effect of Cu(II)

    was also observed.

    Fe(III) was found to be the initial product of pyrite oxidation, although the proportion of total dissolved iron as Fe(III) did

    reach a minimum during each test, indicating the generation and subsequent oxidation of reduced sulfur species. Pyrite

    oxidation kinetics are limited by the rate of reaction at the pyrite surface, with an activation energy of 33.2 kJ/mol (7.9 kcal/mol)

    with respect to dissolved oxygen concentration, or 41.7 kJ/mol (10.0 kcal/mol) with respect to oxygen partial pressure, over the

    temperature range 170230jC. The reaction order with respect to oxygen partial pressure was found to be 0.5 at 210jC,indicating the first charge transfers of both the anodic and cathodic half-cell reactions to be the most likely rate-controlling

    steps.

    Conversion data conform to the shrinking sphere model initially, but deviate at higher conversions, indicating passivation of

    the mineral surface, most likely by elemental sulfur, which may precipitate via the disproportionation of other reduced sulfurDepartment of Metals and Materials Engineering, University of BritPressure oxidation of pyrite in sulfuric acid media:

    a kinetic study

    Hu Long, David G. Dixon*Hydrometallurgy 73Pyrite (FeS2) is the most common iron sulfide

    mineral. In most pyrite ores and concentrates, the

    pyrite itself is rarely of economic importance and is

    0304-386X/$ - see front matter D 2003 Elsevier B.V. All rights reserved.

    doi:10.1016/j.hydromet.2003.07.010

    * Corresponding author. Tel.: +1-604-822-3679; fax: +1-604-

    822-3619.

    E-mail address: [email protected] (D.G. Dixon).www.elsevier.com/locate/hydromet

    4) 335349such as gold, copper and zinc. In many sulfidic

    refractory gold ores and concentrates, gold is often

    found finely disseminated in sulfide minerals, par-

    ticularly in pyrite (Gasparrini, 1983), rendering the

    adequate liberation of gold impossible by grinding

    alone. One method for recovery of gold from these

    refractory gold ores and concentrates is pressure

  • H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349336oxidation in sulfuric acid solution at high temper-

    atures (above 180jC) and oxygen partial pressuresas a pretreatment prior to cyanidation. This process

    has been practiced commercially since the early

    1980s (Argall, 1986), with the object to enhance

    exposure of the gold particles to the cyanide solu-

    tion by breaking down the sulfide lattice.

    Pyrite oxidation has been studied extensively

    because of its importance in sulfide mineral separa-

    tions by flotation, in the generation of acid in mine

    waters and in leaching. Aqueous oxidation of pyrite

    has been reviewed thoroughly by Lowson (1982)

    and Hiskey and Schlitt (1982). In most of the

    pressure oxidation experiments reported (McKay

    and Halpern, 1958; Gerlach et al., 1966; Bailey

    and Peters, 1976; Papangelakis and Demopoulos,

    1991), the oxidation of pyrite was found to yield

    only the following products: ferrous sulfate, ferric

    sulfate, sulfuric acid and elemental sulfur. No sulfur

    products of intermediate oxidation state, such as

    thiosulfate (S2O32) or thionates (SnO6

    2, n= 26),were detectable under any conditions, although

    significant evidence exists to suggest that sulfur

    does indeed pass through such intermediate states

    during pyrite oxidation at lower temperatures (Con-

    way et al., 1980; Goldhaber, 1983; Morse et al.,

    1987; Kelsall and Yin, 1996). The pressure oxida-

    tion of pyrite is typically represented by the follow-

    ing two competing reactions (Bailey and Peters,

    1976):

    FeS2 7=2O2 H2O ! FeSO4 H2SO4 1

    FeS2 2O2 ! FeSO4 S 2

    As temperature increases, reaction (1) becomes

    predominant. The Fe(II) produced is subsequently

    oxidized to Fe(III):

    2FeSO4 1=2O2 H2SO4 ! Fe2SO43 H2O 3

    Ferric sulfate has been reported as an importantoxidizing agent for pyrite (Lowson, 1982; Garrels andThompson, 1960). The oxidation reactions may be

    represented as follows:

    FeS2 7Fe2SO43 8H2O ! 15FeSO4 8H2SO44

    or

    FeS24Fe2SO434H2O! 9FeSO44H2SO4S5

    At high temperatures, ferric sulfate tends to hydro-

    lyze, resulting in the precipitation of ferric oxide

    (Fe2O3) or basic ferric sulfate (Fe(OH)SO4) depend-

    ing on acidity (Tozawa and Sasaki, 1986).

    Table 1 lists the experimental conditions, the orders

    of reaction for oxygen partial pressure, and the

    reported activation energies from the most relevant

    studies on pressure oxidation of pyrite in sulfuric acid

    solution at temperatures above 120jC. A number ofconclusions may be drawn from Table 1:

    1. It is accepted by all researchers that the

    pressure oxidation of pyrite is controlled by the

    surface reaction rate. Reaction orders with respect

    to oxygen partial pressure depend on both tempera-

    ture and pressure. First-order dependence is exhibited

    predominantly at lower oxygen partial pressures

    ( < 20 atm) and at all temperatures employed, sug-

    gesting a mass transfer limitation. One-half order

    dependence is found at higher oxygen partial pres-

    sures and temperatures.

    2. McKay and Halpern (1958), Gerlach et al.

    (1966) and Cornelius and Woodcock, (1958) inter-

    preted their experimental results as evidence of an

    oxygen chemisorption mechanism followed by a slow

    chemical reaction. However, Bailey and Peters (1976)

    convincingly demonstrated the mechanism of pressure

    oxidation of pyrite to be electrochemical, involving

    coupled anodic (pyrite oxidation) and cathodic (oxy-

    gen reduction) reactions.

    3. Activation energies between 46 and 55 kJ/mol

    are reported at temperatures below 160jC. Interest-ingly, Papangelakis and Demopoulos (1991) report an

    activation energy of 110.5 kJ/mol over the range of

    160180jC, which is more than twice the value

    reported by them and others at temperatures below

  • rNatural pyrite 0.5 140160 520 1

    160180 510 1

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349 337160jC. The reason for this significant shift in activa-tion energy is unclear.

    Although extensive investigations have been per-

    formed by previous researchers, all of the reported

    work has been conducted at temperatures lower than

    180jC. No data is available on the kinetics of pyritedissolution during acid pressure oxidation over the

    temperature range from 180 to 230jC, which is therange employed by most commercial plants. From a

    process optimization standpoint, it is important to

    know the behavior of pyrite during acidic pressure

    oxidation at temperatures above 180jC; hence, thenecessity of the present study.

    The main objectives of our investigation were to

    gather data on the pressure oxidation of pyrite over the

    typical temperature range employed in the commercial

    160180 1020 0.5Table 1

    Review of pyrite pressure oxidation studies

    Material Experimental conditions PO2

    [H2SO4], M T, jC PO2, atmorde

    Upgraded pyrite

    concentrate

    0.075 100130 04 1

    Natural pyrite 0.2 60130 015.5 1

    Natural pyrite 1.0 85130 020 1

    2066.4 0.5pressure oxidation of refractory gold ores, and to

    develop a reliable rate equation which may be used

    to model the pressure leaching of refractory gold ores.

    2. Experimental

    2.1. Materials

    High-grade massive pyrite specimens originating

    from Zacatecas, Mexico were obtained from Wards

    Natural Science Establishment, of Ontario, Canada.

    The samples were crushed, ground and dry-sieved to

    different narrow size fractions, namely 149 + 105, 105 + 74, 74 + 53 and 53 + 44 Am. The sam-ples were then washed to remove any fines adheringto particle surfaces which might interfere with the

    interpretation of the results, and air-dried at ambient

    temperature.

    The samples were analyzed by two analytical labs

    in order to obtain reliable assays. The results indicated

    that the purity of the pyrite sample was 97F 1%, thatthe mole ratio of sulfur to iron was 1.92, which is very

    close to the stoichiometric value for pyrite, and that

    iron and sulfide sulfur composition in each size

    fraction showed very little variation, with maximum

    differences of only 0.6% for iron and 1% for sulfide

    sulfur. The pyrite crystal structure was confirmed by

    X-ray diffraction.

    All solutions were prepared with reagent grade

    chemicals and deionized water. Medical grade oxygen

    gas was used in all experiments.

    Activation Assumed mechanism References

    energy,

    kJ/mol

    55.7 Chemisorption +

    chemical reaction

    McKay and Halpern, 1958

    54.8 Chemisorption +

    chemical reaction

    Gerlach et al., 1966

    51.1 Electrochemical Bailey and Peters, 1976

    reaction

    46.2 Electrochemical Papangelakis and

    110.5 reaction Demopoulos, 19912.2. Apparatus

    All pressure oxidation experiments were conducted

    in a 2-L Parr titanium autoclave. The temperature was

    monitored using a thermocouple probe to an accuracy

    of F 1jC and maintained at desired set points byexternal heating and by passing cooling water through

    the internal cooling coils. Cooling water flow was

    regulated by a solenoid valve. Agitation was provided

    by dual four-pitched-blade impellers. Temperature and

    agitation speed were maintained using a Parr control-

    ler unit. Oxygen was introduced into the autoclave

    near the bottom through a dip tube, which was also

    connected to an autoclave sampler. Heat resistant

    superalloy valves were selected to avoid any corrosion

    in contact with oxidizing acid solution.

  • 2.3. Experimental conditions

    The slurry used for most experiments had a stan-

    dard composition of 1 g/L pyrite and 0.5 M H2SO4.

    Based on the results of Tozawa and Sasaki (1986) and

    on our own preliminary tests, the acidity was fixed at

    0.5 M H2SO4 for all the experiments in order to obtain

    a high solubility of Fe2O3 and to avoid Fe(OH)SO4precipitation, since dissolved iron was taken as the

    indicator of pyrite dissolution at low pulp density. The

    proper amount of pyrite to charge into the autoclave

    was determined by a series of tests at different pulp

    densities at 230jC. The results showed that when 1 g/L FeS2 was charged, even if the pyrite was dissolved

    completely, the amount of precipitated iron (owing

    mostly to solution splashing onto the heated walls of

    the reaction vessel above the solution line) corre-

    sponded to less than 5% conversion of the pyrite. At

    lower temperatures the precipitated amount was con-

    siderably less, so 1 g of FeS2/L of solution was used

    in all low pulp density experiments. At higher pulp

    densities, total mass flow of oxygen gas to the

    autoclave was taken as the indicator of pyrite oxida-

    tion, and so the extent of iron precipitation was of no

    concern.

    2.4. Test procedure

    It is well known that titanium is corroded in high

    temperature sulfuric acid solutions in the absence of

    an oxidizing agent (i.e., Cu(II), Fe(III) or O2) (Uhlig

    and Revie, 1985). Therefore, for the experiments in

    the absence of Cu(II) at low pulp density, concen-

    trated sulfuric acid was sealed in ampoules during

    O2,

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349338Fig. 1. Effect of temperature on pyrite oxidation (800 rpm, 690 kPa

    vs. r2, (d) [Fe(III)]/[FeT] vs. time.

    74 + 53 Am). (a) Conversion vs. time, (b) r vs. time, (c) s(dr/dt)

  • autoclave heating to avoid the corrosion of titanium,

    while for the experiments in the presence of Cu(II),

    pyrite was sealed in an ampoule to avoid premature

    oxidation.

    A slurry with several borosilicate glass ampoules

    containing sulfuric acid, or a solution with one boro-

    silicate glass ampoule containing pyrite, was placed in

    the autoclave, which was then sealed and heated.

    Once the desired temperature was reached, an initial

    sample was taken, then the unit was pressurized with

    oxygen for 1 min. Finally, the ampoules were broken

    by engaging the impellers and the test was thus

    initiated.

    In the tests at low pulp density, slurry samples were

    withdrawn periodically from the autoclave through

    the dip tube. The withdrawn sample was cooled

    immediately (in less than 5 s) to room temperature

    and filtered. At the end of the run, the system was

    cooled quickly (in less than 15 min) to room temper-

    ature with the cooling coil. The slurry was removed

    and filtered, the residue was washed and dried, and

    both solids and solution were analyzed.

    In order to avoid contamination from any iron

    residue precipitated onto the walls of the bomb and

    internal parts, the bomb was soaked after each exper-

    iment with 1.5 L of 15% HCl solution for 30 min at a

    temperature of 80jC, then cooled to room tempera-ture. The solution was analyzed for total iron to

    determine the extent of precipitation. Then, after being

    washed with water, the bomb was soaked overnight in

    kPa

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349 339Fig. 2. Effect of particle size on pyrite oxidation (800 rpm, 210jC, 690

    [Fe(III)]/[FeT] vs. time.O2). (a) Conversion vs. time, (b) r vs. time, (c) s(dr/dt) vs. r2, (d)

  • a solution containing 10% HNO3 and 5 g/L iron as

    ferric sulfate in order to encourage the formation of a

    TiO2 film. Blank tests run on 1 L of 0.1 M H2SO4solution under conditions of 230jC and 800 rpmwithin the cleaned bomb showed that the total iron

    concentration was less than 0.14 mg/L after 1 h.

    2.5. Analytical methods

    Solution samples from the low pulp density tests

    were analyzed for Fe(II) by redox titration with

    Ce(SO4)2 solution using ferroin as an indicator (Jeff-

    ery et al., 1989), and for total iron by atomic absorp-

    tion spectrophotometry (AAS). Fe(III) was determined

    as the difference between total iron and Fe(II). At

    higher pulp densities, total oxygen flow was recorded

    digitally from Omega in-line digital oxygen mass

    flowmeters.

    3. Results and discussion

    Pressure oxidation tests were performed under the

    following conditions: agitation speeds from 650 to

    950 rpm, temperatures from 170 to 230jC, meanparticle sizes (diameters) from 49 to 125 Am, oxygenpartial pressures from 345 to 1035 kPa (50150 psi)

    and pyrite pulp densities of 1 and 20 g/L. The mean

    particle size was taken as the square root of the

    product of two adjacent sieve sizes (i.e., the geometric

    pm, 2

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349340Fig. 3. Effect of oxygen partial pressure on pyrite oxidation (800 r s(dr/dt) vs. r2, (d) [Fe(III)]/[FeT] vs. time.

    10jC, 74 + 53 Am). (a) Conversion vs. time, (b) r vs. time, (c)

  • mean sieve opening). The standard conditions for the

    leaching experiments were as follows: agitation speed

    of 800 rpm, temperature of 210jC, O2 partial pressureof 690 kPa (100 psi), particle size of 74 + 53 Am(62.6 Am mean size) and pyrite pulp density of 1 g/Lin 0.5 M H2SO4 solution.

    All of the experimental results are presented in

    Figs. 14 and all experimental conditions and calcu-

    lated parameters are presented in Table 2. The effects

    of temperature (Tests 1 through 4) are shown in Fig. 1,

    particle size (Tests 2 and 5 through 7) in Fig. 2,

    oxygen partial pressure (Tests 2, 8 and 9) in Fig. 3,

    and pyrite pulp density and the addition of copper

    (Tests 2 and 10 through 12) in Fig. 4.

    For the tests at low pulp density, pyrite conversion

    was calculated from the concentration of iron in

    solution divided by the mass of iron in the head

    sample divided by the original solution volume. The

    ratio of Fe(III) to total dissolved iron ([Fe(III)]/[FeT])

    was calculated from the concentration of Fe(III)

    divided by the total iron concentration in solution.

    The time elapsed from sampling to Fe(II) titration was

    never more than 1 h, during which time no significant

    difference in titration results was found from the

    preliminary tests.

    For the tests at high pulp density, a digital oxygen

    mass flowmeter was used to record the consumption

    of oxygen during pressure oxidation. Thus, pyrite

    (800 r

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349 341Fig. 4. Effect of pulp density and copper addition on pyrite oxidationr vs. time, (c) s(dr/dt) vs. r2, (d) [Fe(III)]/[FeT] vs. time (low pulp de

    pm, 210jC, 690 kPa O2, 74 + 53 Am). (a) Conversion vs. time, (b)

    nsity only).

  • 074 + 5

    0

    5

    .9

    21

    438

    0

    74 + 5

    5

    5

    .9

    73

    466

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349342conversion was calculated from the oxygen consumed

    Table 2

    Summary of experimental conditions and calculated parameters

    Test. no. 1 2a 3

    T, jC 230 210 19d0, Am 74 + 53 74 + 53 PO2, kPa 690 690 69

    [FeS2], g/L 1 1 1

    [CuSO4], g/L 0 0 0

    [H2SO4], M 0.5 0.5 0.

    s, min 25.6 38.7 59m 1.09 2.83 3.

    rp2 0.342 0.398 0.

    Test. no. 7 8 9

    T, jC 210 210 21d0, Am 53 + 44 74 + 53 PO2, kPa 690 1035 34

    [FeS2], g/L 1 1 1

    [CuSO4], g/L 0 0 0

    [H2SO4], M 0.5 0.5 0.

    s, min 28.7 31.5 52m 3.22 2.74 2.

    rp2 0.272 0.336 0.

    a Standard conditions.according to reactions (1) and (3). Given that the bulk

    of dissolved iron reported as Fe(III) in every test, this

    method should have incurred, at most, 12% relative

    error in pyrite conversion.

    3.1. Effect of agitation speed

    Agitation speed is often important in liquidsolid

    reactions, especially those under diffusion control.

    The initial rate of pyrite extraction was recorded at

    650, 800 and 950 rpm under the standard conditions.

    It was observed that agitation speed had no significant

    effect on the initial rate of pyrite oxidation when

    maintained at 800 rpm or higher. Based on these

    results, a stirring speed of 800 rpm was chosen for

    all subsequent experiments to minimize iron precipi-

    tation by solution splashing.

    3.2. Effect of temperature

    The effect of temperature (Tests 1 through 4) on

    pyrite conversion is shown in Fig. 1(a). At 230jC,nearly all of the pyrite dissolved within 20 min. Data

    for the first 15 min of leaching at each temperature arewell described by the shrinking sphere model denot-

    4 5 6

    170 210 210

    3 74 + 53 149 + 105 105 + 74690 690 690

    1 1 1

    0 0 0

    0.5 0.5 0.5

    99.1 77.3 53.3

    1.73 2.43 1.70

    0.734 0.415 0.473

    10 11 12

    210 210 210

    3 74 + 53 74 + 53 74 + 53690 690 690

    20 1 20

    0 5 5

    0.5 0.5 0.5

    25.4 32.0 16.9

    2.34

    0.353 ing surface chemical reaction control:

    dX

    dt 31 X

    2=3

    sor

    drds

    1s

    6

    or, in integrated form:

    X 1 1 ts

    3or r 1 t

    s7

    where r=(1X)1/3 = d/d0 (particle diameter/initial par-ticle diameter), s = time for complete oxidation inminutes.

    This fit is demonstrated by the linear plots of r vs.time, shown in the inset of Fig. 1(b). (Similar plots to

    test the shrinking core model denoting product layer

    diffusion control failed to yield straight lines, and

    examination under a scanning electron microscope

    (SEM) found no product layers on partially leached

    pyrite surfaces.) The inverse slopes (or the x-inter-

    cepts) of these r vs. time lines represent s, the time forcomplete reaction predicted by the shrinking sphere

    model. (All parameters calculated from experimental

    results are tabulated in Table 2.)

  • Plotting the natural logarithm of 1/s vs. the inverseabsolute temperature, 1/s, gives an Arrhenius plot, theslope of which represents E/R, where E is theapparent Arrhenius activation energy of pyrite oxida-

    tion and R is the gas constant. As shown in Fig. 5, the

    apparent Arrhenius activation energy by this method

    is 41.7 kJ/mol (10.0 kcal/mol). That this is not the true

    activation energy will be demonstrated below, in the

    section on the effect of oxygen partial pressure.

    Returning to Fig. 1(a) and (b), it may be seen that

    the shrinking sphere model (represented by dotted

    curves) deviates from the conversion and j data, mostmarkedly at the lowest temperature. From Eq. (7), the

    quantity s(dr/dt) should always have a value ofone. However, as shown in Fig. 1(c), plots of s(dr/

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349 343dt) vs. r2 deviate negatively and (more or less)linearly from one below a certain value of r2. Weattribute this deviation to passivation of the mineral

    surface, most likely by elemental sulfur. Indeed, traces

    of elemental sulfur were detected in the residue of the

    test run at 170jC, which also displayed the mostsevere passivation behavior. (It bears noting that the

    values of dr/dt were obtained directly from the datausing weighted central differences of the form:

    drdt

    icti1 titi1 ti1

    ri ri1ti ti1

    ti ti1

    ti1 ti1

    ri1 riti1 ti

    8

    which ensures second-order level accuracy.)

    Fig. 5. Arrhenius plot for determining the effect of temperature onpyrite oxidation.3.3. The passivating shrinking sphere model

    If we define the point on the x-axis of Fig. 1(c)

    representing the onset of passivation (i.e., the point of

    deviation from s(dr/dt) = 1) as rp2, then assuming alinear deviation, to the left of this point the rate is

    given thus:

    s drdt

    1 mr2p r2 9

    where m is the slope of the line and represents the rate

    of passivation. The integrated form of this passivat-

    ing shrinking sphere model is

    r 1 tsat rzrp 10

    r k rp k tankuk rp tanku

    at rVrp and mr2p < 1 11

    r rp1 rpu at rVrp and mr

    2p 1 12

    r k rp k tanhkuk rp tanhku

    at rVrp and mr2p > 1 13

    where k=(j1mrp2j/m)1/2 and u =m[rp (1 t/s)].When mrp

    2V 1, then the line intersects the y-axisand the reaction goes to completion, since the oxida-

    tion rate is still non-zero at 100% conversion; hence,

    incomplete passivation. However, when mrp2>1, then

    the line intersects the x-axis and the reaction fails to

    go to completion, since the oxidation rate falls to zero

    before complete conversion is reached; hence, com-

    plete passivation. Instead, a limiting particle size is

    approached asymptotically:

    limt!lr k at mr

    2pz1 14

    One can see both situations in Fig. 1(c). Complete

    passivation is achieved at the three lowest temper-

    atures, and this observation is borne out in the

  • tests, suggesting that the rate of passivation is a strong

    function of temperature only. Also, the relative parti-

    cle surface area at the onset of passivation, rp2, appears

    to be only a weak function of particle size.

    The relationship between the ratio of [Fe(III)]/[FeT]

    and particle size is shown in Fig. 2(d). It was again

    found that the ratio of [Fe(III)]/[FeT] decreased with

    time at the beginning, and then increased asymptoti-

    cally to about 80%. However, in this case, roughly the

    same minimum [Fe(III)]/[FeT] ratio was achieved at

    each particle size, at times roughly proportional to s.

    3.5. Effect of oxygen partial pressure

    The effect of oxygen partial pressure (Tests 2, 8 and

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349344conversion data shown in Fig. 1(a), and most clearly

    in the r vs. time plots shown in Fig. 1(b), where thebold curves represent the passivating shrinking sphere

    model. (It will be noted that the lines shown in Fig.

    1(c) were calculated based on the passivation param-

    eters which resulted in the best fit of the integrated

    rate laws, given by Eqs. (11)(13), to the r vs. timedata given in Fig. 1(b). These parameters are also

    tabulated in Table 2.)

    This model is similar in concept to a model put

    forward by Crundwell and Godorr (1997) to explain

    the passivation of gold during cyanidation. However,

    whereas the mathematical form of our model has been

    derived from empirical considerations, their model

    was derived based on the assumption that the growth

    of the passivating film obeys the same rate law as

    Langmuir adsorption. Both models give similar

    results, although only ours allows for the onset of

    passivation well after the commencement of leaching,

    and only ours is amenable to a closed-form integral

    solution.

    Fig. 1(d) shows the effect of temperature on the

    ratio of Fe(III) to total dissolved iron. The ratios of

    [Fe(III)]/[FeT] decreased with time at the beginning,

    then increased, indicating that the iron in pyrite is

    most likely oxidized to Fe(III) directly under these

    conditions, and not Fe(II). Fe(III) is subsequently

    reduced to Fe(II). While the rate of Fe(II) oxidation

    increased with increasing temperature, more than 40%

    Fe(II) was detected at the end of the test at 170jC andabout 20% Fe(II) remained even after complete oxi-

    dation of the pyrite at 230jC.In order to clarify the role of pyrite in the oxidation

    of Fe(II), a blank test was run under the standard

    conditions with 0.6 g Fe(II) added to the initial

    solution. The results of this test are shown in Fig. 6.

    While Fe(II) was oxidized (slowly) in the absence of

    pyrite, again, nearly 20% Fe(II) remained in solution

    after pressure oxidation. Hence, the high percentage

    of Fe(II) remaining in solution after pressure oxidation

    of pyrite is not due to the disappearance of pyrite

    (although pyrite definitely catalyzes Fe(II) oxidation)

    but appears to be simply an equilibrium effect. (Of

    course, high levels of Fe(II) in the autoclave discharge

    following acidic pressure oxidation of refractory gold

    ores are to be avoided, and may necessitate an alkaline

    pre-aeration step prior to gold cyanidation, since ironforms strong complexes with cyanide.)3.4. Effect of particle size

    The effect of particle size (Tests 2 and 5 through 7)

    on pyrite conversion is shown in Fig. 2(a). Again,

    reaction times were calculated from the slopes of the jvs. time lines shown in the inset of Fig. 2(b). A plot of

    log 1/s vs. log d0, shown in Fig. 7, yields a straightline with a slope of 1.027, which represents thereaction order with respect to initial particle diameter.

    This is very close to the value of 1.0 expected fromthe shrinking sphere model.

    Plots of s(dr/dt) vs. r2 for the different particlesizes are shown in Fig. 2(c). Similar passivation

    behavior may be observed, although the (somewhat

    random) spread in the values of the passivation slope

    m is much narrower at the single temperature of these

    Fig. 6. Fe(II) oxidation kinetics in the absence of pyrite (800 rpm,

    210jC, 690 kPa O2, 0.6 g/L Fe).9) on pyrite conversion is shown in Fig. 3(a) and on r

  • then one would expect a reaction order with respect

    to oxygen partial pressure of 0.25. (For an enlighten-

    ing discussion of the orders of multistep electrochem-

    ical reactions, interested readers are referred to

    Section 9.1 of Bockris and Reddy, 1970.)

    Now that we have confirmed that the pyrite oxi-

    dation rate is 1/2-order with respect to oxygen partial

    pressure, it is possible to correct the Arrhenius acti-

    vation energy for the temperature effect on the satu-

    ration concentration of dissolved oxygen. According

    to the recent model of Tromans (1998) (shown here in

    slightly modified form), the (molal) saturation con-

    centration of dissolved oxygen may be calculated

    relative to the oxygen partial pressure (in atm) at

    any temperature T (in K) thus:

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349 345in Fig. 3(b). A plot of log 1/s vs. log PO2, shown inFig. 8, yields a straight line with a slope of 0.47

    (nearly 0.5), which represents the reaction order with

    respect to oxygen partial pressure. This finding is

    consistent with the anodic dissolution model of pyrite

    in acidic media suggested by Mishra and Osseo-Asare

    (1988), which assumes the first charge transfer step,

    involving the deprotonation of adsorbed water, to be

    the rate-controlling step of the anodic process:

    FeS2 H2O ! FeS2 OH H e 15

    It is also consistent with the cathodic reduction model

    of oxygen on pyrite in acidic media put forward by

    Biegler (1975), which assumes the first charge trans-

    Fig. 7. Log log plot for determining reaction order with respect to

    particle size.fer step, involving the reduction of adsorbed oxygen

    molecules, to be the rate-controlling step of the

    cathodic process:

    O2 e ! O2 16

    Assuming symmetry factors of approximately 0.5,

    then the first charge transfer step must be the rate-

    controlling step in both half-cell reactions, or else the

    reaction order with respect to oxygen partial pressure

    would report as some value other than 0.5. If, for

    example, the second charge transfer step of the anodic

    process was the rate-controlling step, as Biegler and

    Swift (1979) surmised, then one would expect a

    reaction order with respect to oxygen partial pressure

    of 0.75. Alternatively, if the second charge transfer

    step of the cathodic process were rate-controlling,O2PO2

    wexp

    0:046T2 203:35T lnT 1430:55T 68669

    8:3143T

    17

    where w is a concentration-dependent parameterwhich, for sulfuric acid alone in molal concentration

    units, is expressed as:

    w 11 2:01628H2SO41:253475

    ( )0:16895418

    As may be seen from the results of Tromans model

    shown in Fig. 9, the saturation concentration of

    Fig. 8. Log log plot for determining reaction order with respect tooxygen partial pressure.

  • H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349346dissolved oxygen is a fairly strong increasing function

    of temperature over the range relevant to this study.

    Hence, one would expect the temperature dependence

    of oxygen saturation to obscure significantly the

    true temperature dependence of the pyrite oxidation

    reaction.

    Since s is inversely proportional to the square rootof [O2], then replacing ln 1/s with ln 1/s 1/2ln[O2]on the y-axis of the Arrhenius plot gives the true

    activation energy of pyrite oxidation, which is 33.2

    kJ/mol (7.9 kcal/mol) as shown in Fig. 5. This is an

    important distinction to make, since in general the

    dissolved oxygen concentration in a full-scale auto-

    Fig. 9. Tromans model for oxygen solubility as a function of

    temperature and [H2SO4].clave will not correspond to the saturation concentra-

    tion due to the finite rate of gasliquid mixing. This is

    especially true in the critical first compartment of any

    continuous autoclave. Hence, in order to model such a

    situation rigorously, the rate law of pyrite leaching

    must be expressed in terms of dissolved oxygen

    concentration, and not of partial pressure.

    Plots of s(dr/dt) vs. r2 for the different oxygenpartial pressures are shown in Fig. 3(c). Here, the

    passivation slope m is a virtual constant, suggesting

    that the rate of passivation is not a function of oxygen

    partial pressure. However, the point at which passiv-

    ation begins, rp2, appears to be a linear function of

    oxygen partial pressure, at least at lower pressures.

    A partial explanation for the dependence of the

    onset of passivation on oxygen partial pressure may

    be found in the plots of the [Fe(III)]/[FeT] ratio shown

    in Fig. 3(d). At the lowest pressure, the [Fe(III)]/[FeT]ratio falls to very low values shortly after oxidation

    begins, corresponding to the earliest onset of passiv-

    ation. Hence, it may be that a decrease in the electro-

    chemical potential at the mineral surface, reflected by

    the decrease in the [Fe(III)]/[FeT] ratio, initiates the

    formation of a passivating film. However, the rate at

    which this film then grows would seem to have little

    to do with the surface potential.

    3.6. Effect of pulp density

    The effect of pulp density (Tests 2 and 10) on

    pyrite conversion at 1 and 20 g/L, under the standard

    conditions, is shown in Fig. 4(a) and on r in Fig.4(b). Increasing pulp density has a fairly dramatic

    beneficial effect on the rate of pyrite oxidation.

    However, perhaps the most significant effect of

    increasing pulp density is that it seems to prevent

    passivation, as shown most clearly in the plots of

    s(dr/dt) vs. r2 in Fig. 4(c). The reason for thisbehavior is not completely clear. However, it may

    depend on the fact that the solubilities of both Fe(II)

    and Fe(III) at these temperatures are very low (on the

    order of 1 g/L) (Zakharchenko and Tsitsorin, 1949).

    Hence, given the increase in catalytic (pyrite) surface

    area relative to the concentration of iron in solution, it

    may be that the potential near the pyrite surface is

    maintained at levels high enough to avoid the onset

    of passivation.

    3.7. Effect of copper addition

    The effect of copper ions (Tests 2, 11 and 12),

    added as 5 g/L anhydrous CuSO4 (thus giving 2 g/L

    Cu), on pyrite conversion and r are also shown inFig. 4(a) and (b), respectively, at both low (1 g/L

    FeS2) and high (20 g/L FeS2) pulp densities under

    the standard conditions. The effect of copper addi-

    tion at low pulp density is fairly modest, causing a

    slight increase (roughly 14%) in the oxidation rate,

    and no substantial change in the passivation behav-

    ior, as shown in Fig. 4(c). However, the effect at

    high pulp density is much more dramatic, increasing

    the oxidation rate by over 50%.

    It is doubtful whether copper is actually involved

    in any galvanic interactions at the pyrite surface.

    Instead, its catalytic effect probably stems from thefact that the reaction between Cu(I) and dissolved

  • (7) and (8). For the particular pyrite sample studied,

    at a pulp density of 1 g/L and in the absence of

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349 347oxygen under these conditions is virtually instanta-

    neous (Kimweri and Dixon, 2001), while that between

    Fe(II) and dissolved oxygen is very slow, as shown in

    Fig. 6. Also, it is well known that Cu(II) acts as a

    redox catalyst for dissolved oxygen in the oxidation of

    Fe(II) (Dreisinger and Peters, 1989). Hence, in the

    presence of copper, a larger proportion of Fe(II)

    oxidation can occur homogeneously, in the bulk

    solution rather than at the pyrite surface. This effec-

    tively decreases the number of electron transfer steps

    involved in the anodic process, which in turn serves to

    increase the potential at the pyrite surface. This

    increase in potential is also reflected by an increase

    in the terminal [Fe(III)]/[FeT] ratio at low pulp density

    from about 80% in the absence of copper to almost

    95% in the presence of copper, as shown in Fig. 4(d).

    At high pulp density, given the higher ratios of pyrite

    surface area to dissolved iron, the catalytic effect of

    copper is amplified.

    3.8. Explanation of the passivation phenomenon

    Several investigators have postulated the mecha-

    nism of pyrite oxidation as involving the formation of

    sulfoxy intermediates, most notably thiosulfate (Con-

    way et al., 1980; Goldhaber, 1983; Morse et al., 1987;

    Kelsall and Yin, 1996). Although the mechanism is

    quite complex, involving multiple adsorption, proton-

    ation, hydration, and desorption steps, the overall

    stoichiometry of the strictly electrochemical compo-

    nent of pyrite oxidation can probably be represented

    thus:

    4FeS2 7O2 4H ! 4Fe3 4S2O23 2H2O19

    Thiosulfate is highly unstable in acid and dispropor-

    tionates rapidly into elemental sulfur and sulfite:

    S2O23 ! S0 SO23 20

    The resulting sulfite is further oxidized by Fe(III) to

    sulfate:

    SO23 2Fe3H2O ! SO24 2Fe2 2H 21

    The sulfur, most likely in the form of polysulfidesadsorbed onto the pyrite surface, is also oxidized,dissolved copper, the time constant s may be calcu-lated thus:

    1

    s 6:50 10

    4O21=2d0

    exp 3993s

    22

    where s is in minutes, [O2] is the molal concentrationof dissolved oxygen as given by Tromans modelpossibly through one or more thionate intermediates.

    However, if the rate of this oxidation were to

    become too slow, owing perhaps to a decrease in

    potential near the pyrite surface due to the local

    buildup of Fe(II) as indicated in Figs. 1(d), 2(d), 3(d)

    and 4(d), then elemental sulfur would tend to accu-

    mulate on the pyrite surface, ultimately forming a

    passivating film.

    4. Conclusions

    The oxygen pressure oxidation kinetics of pyrite in

    sulfuric acid media were studied in the temperature

    range of 170230jC. A number of conclusions maybe drawn from the results:

    1. Pyrite oxidation during acidic oxygen pressure

    leaching at high temperature exhibits 1/2-order de-

    pendency on dissolved oxygen concentration, sug-

    gesting an electrochemical leaching mechanism.

    2. Fe(III) appears to be produced directly from the

    oxidation of pyrite by dissolved oxygen, with a portion

    being subsequently reduced to Fe(II), most likely

    during the oxidation of reduced sulfur intermediates.

    3. At low pulp density, pyrite forms a passivating

    film which impedes oxidation, and which may cause it

    to cease altogether, depending on the extent of oxi-

    dation at the onset of passivation and the rate of

    passivation. The former would appear to be a function

    of electrochemical potential at the pyrite surface (as

    indicated by the ratio of Fe(III) to total iron in

    solution), while the latter is mostly a function of

    temperature. Increasing the pulp density appears to

    prevent the onset of passivation.

    4. At least until the onset of passivation, pyrite

    oxidation conforms to the shrinking sphere model,

    denoting surface reaction control, as given in Eqs.(Tromans, 1998), d0 is in microns and T is in Kelvin.

  • (9)(13).

    R gas constant (8.314 J/gmol K)

    u parameter of the passivation shrinking sphere

    194122-96) for their financial support.

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349348References

    Argall, O.G., 1986. Perseverance and winning ways at McLaughlin

    Gold. E&MJ 187 (10), 26.

    Bailey, L.K., Peters, E., 1976. Decomposition of pyrite in acids by

    pressure leaching and anodization: the case for an electrochem-

    ical mechanism. Can. Metall. Q. 15 (4), 333.

    Biegler, T., 1975. Oxygen reduction on sulphide minerals: Part II.model

    w parameter of Tromans model of oxygensaturation

    Acknowledgements

    The authors express their sincere gratitude to

    Barrick Gold Corporation, SNC-Lavalin and the

    Natural Science and Engineering Research Council

    of Canada (Industrially Oriented Research Grantt time, min

    T temperature, jC or KX pyrite conversion

    Greek letters

    k parameter of the passivating shrinking spheremodel

    r dimensionless particle sizes time required for complete reaction, minNomenclature

    d particle size, Amd0 initial particle size, AmE Arrhenius activation energy, kJ/mol

    m passivation slope

    PO2 oxygen partial pressure, kPaHowever, the actual value of the time constant is

    dependent on the pyrite pulp density and the presence

    of catalytic ions such as Cu(II) (and presumably on the

    source of the pyrite sample as well).

    5. Beyond the onset of passivation, the rate of

    pyrite oxidation is adequately described by a pas-

    sivating shrinking sphere model as given in Eqs.Relation between activity and semiconducting properties of pyr-

    ite electrodes. J. Electroanal. Chem. 70, 265.

    Biegler, T., Swift, D.A., 1979. Anodic behaviour of pyrite in acid

    solutions. Electrochim. Acta 24, 415.

    Bockris, J.OM., Reddy, A.K.N., 1970. Modern Electrochemistry,

    vol. 2. Plenum, New York, pp. 9911017.

    Conway, B.E., Ku, J.C.H., Ho, F.C., 1980. The electrochemical

    surface reactivity of iron sulfide, FeS2. J. Colloid Interface

    Sci. 75 (2), 357.

    Cornelius, R.J., Woodcock, J.T., 1958. Pressure leaching of a man-

    ganese ore. Aust. Inst. Min. Metall. Proc. 185, 65.

    Crundwell, F.K., Godorr, S.A., 1997. A mathematical model of the

    leaching of gold in cyanide solutions. Hydrometallurgy 44, 147.

    Dreisinger, D.B., Peters, E., 1989. The oxidation of ferrous sul-

    phate under zinc pressure leaching conditions. Hydrometal-

    lurgy 22, 101.

    Garrels, R.M., Thompson, M.E., 1960. Oxidation of pyrite by iron

    sulfate solution. Am. J. Sci. 258A, 57.

    Gasparrini, C., 1983. The mineralogy of gold and its significance in

    metal extraction. CIM Bull. 76 (851), 144.

    Gerlach, J., Haehne, H., Pawlek, F.Z., 1966. Pressure leaching of Fe

    sulfides: II. Kinetics of the pressure leaching of pyrite. Z. Erz-

    bergbau Metall. 19 (2), 66.

    Goldhaber, M.B., 1983. Experimental study of metastable sulfur

    oxyanion formation during pyrite oxidation. Am. J. Sci. 283

    (3), 193.

    Hiskey, J.B., Schlitt, W.J., 1982. Aqueous Oxidation Pyrite. In:

    Schlitt, W.J., Hiskey, J.B. (Eds.), Interfacing Technologies in

    Solution Mining Proceedings, 2nd SME-SPE Intl. Solution Min-

    ing Symp. SME-AIME, Littleton (Colorado), p. 55.

    Jeffery, G.H., Bassett, J., Mendham, J., Denney, R., 1989. Vogels

    Textbook of Quantitative Inorganic Analysis, 5th ed. Longman,

    New York, p. 382.

    Kelsall, G.H., Yin, Q., 1996. Electrochemical oxidation of pyrite

    (FeS2) in acidic aqueous electrolytes I. In: Woods, R., Doyle,

    F.M., Richardson, P. (Eds.), Proceedings of the Fourth Interna-

    tional Symposium on Electrochemistry in Mineral and Metal

    Processing, vol. 96-6, p. 131.

    Kimweri, H.T.H., Dixon, D.G., 2001. Unpublished results.

    Lowson, R.T., 1982. Aqueous oxidation of pyrite by molecular

    oxygen. Chem. Rev. 82 (5), 461.

    McKay, D.R., Halpern, J., 1958. A kinetic study of the oxidation

    of pyrite in aqueous suspension. Trans. Metall. Soc. AIME

    212, 301.

    Mishra, K.K., Osseo-Asare, K., 1988. Aspects of the interfacial

    electrochemistry of semiconductor pyrite (FeS2). J. Electro-

    chem. Soc. 135, 2502.

    Morse, J.W., Millero, F.J., Cornwell, J.C., Rickard, D., 1987. The

    chemistry of the hydrogen sulfide and iron sulfide systems in

    natural water. Earth Sci. Rev. 24 (1), 1.

    Papangelakis, V.G., Demopoulos, G.P., 1991. Acid pressure oxida-

    tion of pyrite: reaction kinetics. Hydrometallurgy 26, 309.

    Tozawa, K., Sasaki, K., 1986. Effect of coexisting sulphates on

    precipitation of ferric oxide from ferric sulphate solutions at

    elevated temperature. In: Dutrizac, J.E., Monhemius, A.J.

    (Eds.), Iron Control in Hydrometallurgy. Wiley, New York,

    p. 454.

  • Tromans, D., 1998. Oxygen solubility modeling in inorganic sol-

    utions: concentration, temperature and pressure effects. Hydro-

    metallurgy 50, 279.

    Uhlig, H.H., Revie, R.W., 1985. Corrosion and Corrosion Control,

    2nd ed. Wiley, New York, p. 363.

    Zakharchenko, G.A., Tsitsorin, V.A., 1949. Zhur. Priklad. Khim.

    22, 703 (as reported in Linke, W.F., 1958. Solubilities

    Inorganic and MetalOrganic Compounds, 4th ed. American

    Chemical Society, Washington, DC, p. 1053).

    H. Long, D.G. Dixon / Hydrometallurgy 73 (2004) 335349 349

    Pressure oxidation of pyrite in sulfuric acid media: a kinetic studyIntroductionExperimentalMaterialsApparatusExperimental conditionsTest procedureAnalytical methods

    Results and discussionEffect of agitation speedEffect of temperatureThe passivating shrinking sphere modelEffect of particle sizeEffect of oxygen partial pressureEffect of pulp densityEffect of copper additionExplanation of the passivation phenomenon

    ConclusionsAcknowledgementsReferences