You are on page 1of 82

STEADY STATE, GEOCHEMICAL BOX

MODEL OF A TAILINGS IMPOUNDMENT:


Application to Impoundment 1, Kristineberg,
Sweden, and prediction of effect of remediation
S.U. Salmon and M. Malmstrm
Water Resources Engineering,
Department of Civil and Environmental Engineering,
Royal Institute of Technology,
Brinellvgen 32, S-100 44 Stockholm, Sweden

MiMi Report
March 2000

Summary
The aim of these investigations was to develop a geochemical box model that could help
improve understanding of the processes occurring in a tailings impoundment, to reproduce the
important aspects of impoundment geochemistry and to predict the effect of remediation of
the impoundment. Rate expressions for slow processes that are controlled by chemical
kinetics, such as sulphide oxidation and silicate dissolution, were selected from the literature
and compared. Using these rate expressions, regions of mechanism dominance for sulphide
and ferrous iron oxidation were framed in terms of master variables.
Available field data from Impoundment 1, Kristineberg, was analysed and used to constrain
as far as possible a general conceptual model. The slow rates expressions were coupled to
fast, equilibrium processes, such as aqueous speciation and solubility equilibrium with
secondary minerals, in order to create a mathematical model of a generalised tailings
impoundment. Modelling with field input data gave results that compared favourably with site
data. The model was furthermore calibrated to the site using two scaling factors; model results
after calibration were even closer to field conditions. However, it should be noted that the
calibration factors were small; it is possible to account for the major deviations from the field
conditions by considering the uncertainty in parameters and constants in the expressions used.
This implied that the field geochemistry could be explained with an essentially abiotic model.
The response of the impoundment geochemistry to changed conditions, such as can be caused
by remediation, was modelled. Change in the partial pressure of oxygen had a much greater
effect than individual variation in infiltration rate or water content; decreased partial pressure
of oxygen led to decreased concentrations and fluxes of all modelled elements as well as
increased pH. Addition of lime led to increased pH but also increased sulphate concentrations,
reflecting an increased sulphide oxidation rate.
The hypothesis that with decreasing partial pressure of oxygen, pyrite oxidation by ferric iron
would become increasingly important and maintain high pyrite oxidation rates was not
supported by the model results. A theoretical investigation showed that this process only
becomes important when the pH decreases below four.

Contents

Introduction............................................................................................................1

Modelling biogeochemical processes in mill tailings deposits .............................2


2.1 Geochemical databases - general................................................................................ 3
2.1.1
2.1.2

Equilibrium processes .........................................................................................................3


Empirical rate expressions ..................................................................................................4

2.2 Sulphide oxidation reactions....................................................................................... 5


2.2.1
2.2.2
2.2.3

Abiotic oxidation of pyrite ..................................................................................................5


Microbially mediated oxidation of pyrite............................................................................7
Mechanism predominance for pyrite oxidation.................................................................10

2.3 Ferrous iron oxidation............................................................................................... 12


2.3.1
2.3.2
2.3.3
2.3.4

Abiotic homogeneous oxidation........................................................................................12


Surface mediated oxidation...............................................................................................13
Microbially mediated oxidation ........................................................................................14
Mechanism predominance ................................................................................................15

2.4 pH buffering reactions............................................................................................... 16


2.4.1
2.4.2
2.4.3

Carbonate weathering .......................................................................................................16


(Hydr)oxide weathering ....................................................................................................17
Silicate weathering............................................................................................................18

Model description.................................................................................................19
3.1 The STEADYQL Code.............................................................................................. 19
3.1.1
3.1.2

Code description ...............................................................................................................19


Comments on the STEADYQL code ................................................................................20

3.2 The model applied to Impoundment 1 at the Kristineberg site............................. 21


3.2.1
3.2.2
3.2.3

Base case conceptual model ..........................................................................................21


Geochemical equilibrium and kinetic database used in the model....................................26
Input data ..........................................................................................................................26

3.3 Calibration.................................................................................................................. 28
3.3.1
3.3.2

Calibration factors and sensitivity.....................................................................................28


Calibration results .............................................................................................................30

Variations on the base case .................................................................................32


4.1 Absence of ferric hydroxide ..................................................................................... 32
4.2 Equilibrium with gibbsite ......................................................................................... 33
4.3 CO2 partial pressures ................................................................................................ 34
4.4 Equilibrium with gypsum ......................................................................................... 34
4.4.1
4.4.2

Gypsum equilibrium..........................................................................................................34
Influx of calcium...............................................................................................................36

4.5 Presence of calcite ...................................................................................................... 37


4.5.1
4.5.2
4.5.3

Assumed equilibrium with calcite.....................................................................................37


Kinetic dissolution of calcite.............................................................................................38
Influx of a water in contact with calcite ............................................................................38

4.6 Oxygen transport limitation on sulphide oxidation................................................ 40


4.6.1
4.6.2

Varying the effective diffusion coefficient .......................................................................42


Varying water content .......................................................................................................43

4.7 Heterogeneous ferrous iron oxidation...................................................................... 43

Base case results and modelling of remediation.................................................44


5.1 The base case: Modelling of Impoundment 1.......................................................... 44
5.1.1

Oxygen consumption ........................................................................................................44

5.1.2
5.1.3
5.1.4

Iron....................................................................................................................................45
Proton balance...................................................................................................................45
Sulphide oxidation mechanism dominance.......................................................................46

5.2 Investigation of the effect of remediation ................................................................ 47


5.3 Results of investigations ............................................................................................ 48
5.3.1
5.3.2

Overall rate........................................................................................................................48
Pyrite oxidation mechanism dominance ...........................................................................51

Discussion ............................................................................................................53

Conclusions ..........................................................................................................54

References ............................................................................................................56

Appendices............................................................................................................59
Appendix A. Pyrite oxidation rate expressions ............................................................... 59
Appendix B. Pyrite oxidation mechanism predominance. ............................................. 60
Appendix C. Ferrous iron oxidation mechanism predominance................................... 66
Appendix D. Geochemical database: Slow processes ..................................................... 73
Appendix E. Geochemical database: Equilibrium processes......................................... 77

1 Introduction
Waste products from the mining of sulphidic, base metal ores are often in the form of finely
ground mill tailings. Oxidation of metal sulphides in tailings impoundments, combined with
other biogeochemical processes, can lead to the production of acidic, metal-laden drainage
water (acid mine drainage, or AMD) over extended periods of time. This can result in serious
consequences for the environment receiving the AMD.
Processes occurring in a mill tailings impoundment take place over different time scales.
Reactions such as precipitation, dissolution, and aqueous speciation often occur at a
sufficiently fast rate that equilibrium conditions can be assumed. Other processes are limited
by chemical or transport kinetics, such as homogeneous redox reactions, heterogeneous
weathering of metal sulphides and diffusion of gases. The final geochemistry of the
impoundment is a result of the interactions between all processes and the master variables pe
and pH.
In this study we have investigated these phenomena through the construction of a steady state
box model, based on a conceptual model that includes both equilibrium reactions and
reactions limited by slow chemical kinetics. To select appropriate rate laws for pyrite and
ferrous iron (Fe(II)) oxidation we have compared various empirical expressions reported in
the literature. Using the selected expressions we have framed regions in terms of master
variables pH and Po2 where different reaction mechanisms dominate.
Model results have been compared with site data in order to investigate the possibility of
prediction of impoundment geochemistry. In this part of the study, parameters such as
physical dimensions, mineralogy and infiltration rates have been taken from pre-remediation
data from Impoundment 1, Kristineberg (see compilation by Malmstrm et al., 1999).
Remediation strategies to decrease AMD generation often involve attempts to decrease the
partial pressure of oxygen (Po2) in the impoundment. The water content and infiltration rate
are also often manipulated either individually or at the same time, to help hinder the diffusion
of oxygen or decrease the amount of leachate produced. The effect of remediation on an
impoundment was investigated by varying these three parameters in the base case model.
Of particular interest to this study was the possibility that under certain conditions, the
dominant oxidation process for metal sulphides may involve an oxidant other than oxygen,
such as ferric iron, Fe(III). Such conditions may occur for example due to groundwater
fluctuations, or remediation of an impoundment, where the intention is often to decrease
oxygen concentrations in the impoundment. In the latter case, the presence and action of other
oxidants than oxygen may result in less effective remediation than expected.
Other aspects of the conceptual model investigated in this study include:
the presence or absence of secondary minerals ferrihydrite (Fe(OH)3(am)), gibbsite
(Al(OH)3(am)), calcite (CaCO3(s)) and gypsum (CaSO4.2H2O(s)), as solubility equilibrium
with these minerals can affect porewater geochemistry by controlling aqueous
concentrations of the associated ions,
the possibility of biological catalysis of sulphide and ferrous iron oxidation reactions,
the effect of the input of oxygen being limited by transport kinetics (diffusion), rather
than being assumed to be at equilibrium, and
the possible significance of catalysis of oxidation of ferrous iron species adsorbed onto
mineral surfaces.

This report has been produced within the MiMi project, Mitigation of the environmental
impact from mining waste (MiMi, 1997), which is financed by the Swedish Foundation for
Strategic Environmental Research (MISTRA).

2 Modelling biogeochemical processes in mill tailings


deposits
Oxidation of metal sulphides is generally considered to be one of the dominant processes
affecting the geochemistry of AMD-generating tailings impoundments. However, the final
leachate composition is a function of all physical and chemical processes occurring within an
impoundment. The geochemistry in a tailings impoundment will also depend upon processes
such as:
other slow reactions controlled by chemical kinetics, such as the dissolution of silicate
minerals with the associated buffering effect, and the homogenous oxidation of ferrous
iron to ferric iron;
reactions that occur sufficiently fast to be considered to be at equilibrium, such as
aqueous speciation in the porewater, and precipitation and/or dissolution of various
secondary phases; and
physical processes within and around the impoundment.
These processes can affect and be affected by the master variables pH and pe. A model that
successfully represents the interactions between the main processes in an impoundment can
assist in the understanding of the relative importance of the various processes, as well as in
the prediction of the outcome of changed conditions, for example, due to remediation.
The modelling process begins with a conceptual model, which is a qualitative description of
the processes that are considered important in the generation of AMD. One means of then
testing whether the conceptual model is an accurate representation of the system is to
mathematically quantify the processes in such a way that the resulting system of mathematical
equations can be solved for state variables, usually using numerical computer-based
techniques. If model results compare favourably with field data, the choice and representation
of processes in the model is supported. If model results deviate significantly from field
measurements, this may be an indication that the conceptual model or process representation
is in some way incorrect. In such a way the model, and understanding of the processes at
work, is iteratively refined.
Mathematical descriptions of biogeochemical processes may be in the form of rate equations
for reactions limited by kinetics, or mass action expressions with thermodynamic constants
for processes assumed to be at equilibrium. Mathematical formulas and constants can be
obtained by theoretical calculation, observations in the field, experimentation and/or data and
expressions reported in the literature. The collection of expressions and constants used in a
model is known as the geochemical database for that model (see Section 2.1). In the case of
AMD, the most important processes are sulphide oxidation, ferrous iron oxidation, and pHbuffering processes; these processes are described in Sections 2.2-2.4.

2.1 Geochemical databases - general


2.1.1

Equilibrium processes

The general form of a reaction can be written:

aA + bB cC + dD

(2-1)
where A and B represent reactants, C and D represent products, and a, b, c and d are
stoichiometric coefficients. The general mass action equation that describes the equilibrium
composition is:
c
d
{
C} {D}
K '=
{A}a {B}b

(2-2a)

where K is the equilibrium constant (compare to Equation 3-1), and {A} denotes activity.
Activity is related to concentration, [A], via an activity coefficient, fA, where fA= {A}/[A].
Hence an equilibrium expression based on concentration,

K ' f Aa f Bb [C ] [D ]
=
f Cc f Dd
[A]a [B ]b
c

K=

(2-2b)

can be defined. When Equation 2-2 is used to describe solubility equilibria, the activity of the
solid phases is set to 1.
Ionic strength, I, is defined by:

I=

1
2

Ci zi2

(2-3)

where C is concentration and z is the charge of the ion. Equilibrium constants can be
corrected for field ionic strength, for example, using the Davies equation (Stumm and
Morgan, 1996, p103):

I
2
log f i = Az i
0 .2 I
1+ I

(2-4)

where f is the activity coefficient as given above. A in this case is a parameter determined by
the expression 1.82 x 106(T)-3/2, where = is the dielectric constant and T the temperature. The
Davies equation is applicable for ionic strengths less than approximately 0.5 M. Geochemical
processes in mining wastes can lead to solutions with very high concentrations and high ionic
strengths.
Equilibrium constants can be corrected for field temperature using the Vant Hoffs equation
(Stumm and Morgan, 1996, p.52):

K H r 1 1
ln
=

R T T
K

(2-5)

where K is the known equilibrium constant at temperature T, K is the equilibrium constant


for the reaction at temperature T, H r is the reaction enthalpy and R is the gas constant.
Thermodynamic constants for the geochemical database can be obtained from literature
sources. Some examples of databases often used for AMD applications include Nordstrom et
al. (1990), WATEQ (Ball and Nordstrom, 1991) and MINTEQ (Allison et al., 1991).

2.1.2

Empirical rate expressions

For the general reaction given by Equation 2-1 above, an empirical rate expression may be of
the form:
m
n
R = k [ A] [B ]
(2-6)
where k is the rate constant and m and n are reaction orders that must be experimentally
determined. A general rate expression for kinetically controlled reactions is given in Section
3.1.1 (Equation 3-2). Rate expressions for reactions limited by chemical kinetics can be found
in the literature, often as empirical expressions obtained from analysis of experimental results
or field measurements.
Although the data available for equilibrium processes is not complete or necessarily
consistent, there is still more data available for equilibrium processes than for processes
limited by chemical kinetic. It should be noted that rate constants and reaction orders
determined experimentally can be associated with considerable uncertainty, which can affect
model results.
There is a shortage of quantitative rate expressions that agree with each other and are
applicable for field conditions, particularly in the case of AMD in cold climates, where
conditions of low temperature and high ionic strength prevail. This is particularly the case for
metal sulphides other than pyrite; for example, no expressions were found in the literature for
oxidation of chalcopyrite, galena, or sphalerite by oxygen. Also lacking is agreement between
laboratory-determined rate expressions and rates determined at field scale (Malmstrm et al.,
1999a; see also Section 3.3).
Despite a number of experiments and publications discussing the effect of biological catalysis
of sulphide weathering reactions, there is a particular shortage of quantitative biological rate
expressions. Most publications are concerned with the bacteria Thiobacillus ferrooxidans; for
other species there are few publications (e.g. Schrenk et al., 1998) and no rate expressions that
we are aware of. Mixtures of different bacterial species are reported to have different effects
to pure cultures, but only qualitatively, and the effect of low temperature is often not
explored. Varying conditions, such as changing redox state and oxygen concentration, may
also effect the type and activity of bacteria.
It has been reported in the literature that galvanic effects can exist between different
sulphides, accelerating the rate of oxidation of the sulphide with the lower potential in the
electromotive series by up to three orders of magnitude (e.g. Scharer et al., 1994, Kwong
1995). Galvanic effects can also be accelerated or hindered in the presence of bacteria and
vice versa, however, a systematic quantification of the effect on reaction rates is not given,
only experimental results for a few conditions (e.g. see review in Herbert, 1999).
Most of the rate expressions reported in the literature are empirical rather than mechanistic in
nature, coming from equations fitted to experimental or field data. For example, most kinetic
expressions for oxidation of sulphides by ferrous iron involve expressions with free Fe(III)
concentration to some exponent, such as in Equation 2-11 below. This is in contradiction with
at least one suggested mechanistic model for pyrite oxidation where it is the first hydrolysis
species of Fe(III), Fe(H2O)5(OH)2+ (often written FeOH2+) that interacts with the pyrite
surface, rather than free Fe(III) (Moses et al., 1987).
Rate constants can be corrected for field temperature using the Arrhenius equation (Stumm
and Morgan, 1996, p.73):

k = Ae

Ea
RT

where A is in this case the pre-exponential factor and Ea is the activation energy.

(2-7)

Described in the sections below are the representations for some typical processes in tailings
impoundments, namely abiotic and biologically mediated rate expressions for oxidation of
sulphides and ferrous iron, and rate and/or equilibrium expressions for pH-buffering reactions
such as dissolution of silicate minerals, carbonates and secondary hydroxides.

2.2 Sulphide oxidation reactions


Two parallel mechanisms have been considered for the weathering of metal sulphides, with
either aqueous oxygen or ferric iron as oxidant, here exemplified by oxidation of the iron
sulphide pyrite:

7
FeS 2 ( s ) + O2( aq ) + H 2O Fe 2+ + 2 SO42 + 2 H +
2
FeS 2 ( s ) + 14 Fe 3+ + 8H 2O 15Fe 2+ + 2SO42 + 16 H +

(2-8)
(2-9)

Under conditions where oxidant availability is not limiting, e.g. by slow transport, these
reactions are usually limited by chemical reaction kinetics. Abiotic rate expressions exist in
the literature for reaction of pyrite with both oxidants, and bacterial agents may also mediate
the reactions. For all reaction types, rate expressions from the literature are discussed below.
Focus in this section is on pyrite, as this mineral has been the subject of a greater number of
investigations than any other sulphide and hence there is more information available.
Furthermore, in the field data for the site which we have compared model results with, pyrite
is reportedly the sulphide present in the largest quantities, such that oxidation of this mineral
is the dominant process in that case.

2.2.1

Abiotic oxidation of pyrite

Experimentally determined empirical rate expressions can be found in the literature for
processes limited by chemical kinetics, such as the following examples of expressions for the
oxidation of pyrite:

[O2 (aq)] 0.5

By O2 (Williamson and Rimstidt, 1994):

Rate = koxy

By Fe(III) (Rimstidt and Newcomb, 1993):

Rate = k fer Fe 3+

[H ]

+ 0.11

0.62

(2-10)

(2-11)

where koxy and kfer are rate constants. Given in Appendix A are the expressions that were
selected from the literature as being relevant for the conditions of interest for this
investigation, namely sulphide oxidation occurring in a tailings impoundment with low pH,
low temperature and high ionic strength. The expressions given by different authors vary in
the species they depend upon and the reaction orders for these species. For example, of the
expressions for oxidation of pyrite by oxygen, only one involves dependence on proton
concentration (Equation 2-10).
Figure 2-1 compares the values of the different rate expressions under the same conditions for
oxidation of pyrite by oxygen. The figure shows that the reported rate laws predict surface
area normalised rates of pyrite oxidation that differ by up to two orders of magnitude. The
majority of the rate laws, however, agree within one order of magnitude and predict a rate of
10-9-10-10 mol m-2s-1 at atmospheric conditions (log [O2(aq)] = -3.6 at 25C).

The expression reported by Williamson and Rimstidt (1994) (Equation 2-10) reputedly
applies over a wider pH range (2-10) than the other expressions; for this reason this
expression was chosen to represent pyrite oxidation by oxygen in this study.

log rate (mol m-2s-1)

-8
Williamson & Rimstidt,
1994, at pH 2
Williamson & Rimstidt,
1994, at pH 5
Nicholson et al., 1990,
valid for pH 7.6-8.6
McKibben & Barnes, 1986,
valid for pH 2-4
Nicholson et al., 1988,
valid for pH 6.7-8.5

-9

-10

-11

-12
-7

-6

-5
log O2 (mol l-1)

-4

-3

Figure 2-1. Rates of pyrite oxidation by oxygen as given in the literature; valid at the given
pH and 25-30C. The rate expression given by Williamson and Rimstidt (1994) is the only
expression with pH dependence. Rate expressions are given in Appendix A.
Figure 2-2 compares expressions for oxidation of pyrite by ferric iron at pH 4. For this study
the expression found in Rimstidt and Newcomb (1993) (Equation 2-11) was chosen to
represent pyrite oxidation by ferric iron, as it seemed closest to a mean rate value over a range
of pH values and Fe(III) concentrations (e.g. at pH 4, Figure 2-2). The reported rate laws for
ferric iron oxidation of pyrite predict surface area normalised rates that differ by many orders
of magnitude at low ferric iron concentrations. Note that the expressions tend to agree best
over the usual range of experimental conditions, e.g. for oxidation by Fe(III), for -5<log
[Fe(III)]<-2.
Rimstidt & Newcombe,
1993

-2 -1

log rate (mol m s )

-6

Wiersma & Rimstidt,


1984

-8
McKibben & Barnes,
1986
Williamson & Rimstidt,
1994 (absence of
dissolved O2)

-10

Williamson & Rimstidt,


1994 (presence of
dissolved O2)

-12
-8

-6

3+

-4

-2

log [Fe ]

Figure 2-2. Pyrite oxidation by ferric iron at pH=4; log [Fe(II)]=2, 25-30. Rates varied
somewhat with varying pH; experimental pH for the selected expressions was reported to be
between pH 0.5-3. Rate expressions are given in Appendix A.
Under appropriate conditions, the rate of oxidation of metal sulphides by either mechanism
may be accelerated by biological catalysis, either due to catalysis of the direct oxidation of
pyrite (see Section 2.2.2), or catalysis of the oxidation of Fe(II) to Fe(III) (see Section 2.3.3).
Weathering of metal sulphides may also be accelerated by galvanic interactions between
sulphides, however, as mentioned in Section 2.1.2, no rate expressions which include the
quantification of this effect were found in the literature.

2.2.2

Microbially mediated oxidation of pyrite

Microbial mediation is a potentially important path for sulphide oxidation and several authors
report biotic oxidation rates that exceed reported abiotic rates (e.g. see review in Herbert,
1999).
Although a catalytic effect of microbes has often been reported in the literature, systematic,
quantitative studies of the oxidation rate are sparse. Additionally, in many studies of the biotic
oxidation of pyrite the surface area has not been reported which hence limits the quantitative
use of these data. In order to be able to include the effects of microbial action in models, the
effect of different microbes and cultures on different sulphides must be assessed. Data to base
such models on are presently not available.
This section briefly reviews and compares two different approaches to modelling microbial
sulphide oxidation that have been presented in the literature. The interactions between
microorganisms and sulphides are complex and the magnitude of catalysis provided by this
mechanism depends on many parameters, for only a few of which the effects have been
quantified. However, in the literature, the oxygen concentration, pH, and the temperature have
been recognised as key parameters for limiting the biotic oxidation rate.
Jaynes et al. (1984) suggested an empirical rate law for the pyrite oxidation rate:
0 X 1

Rb = kX T X pH X O 2

(2-12)

where XT, XpH, XO2 give the temperature-, pH- and oxygen concentration-dependence of the
oxidation rate, Rb. Hence, k is the oxidation rate at optimal conditions, that is, when
XT=XpH=XO2=1. Jaynes et al. (1984) fitted empirical expressions to data reported in the
literature and derived the following expressions:

X T = 1.2 10 5 T 3 4.4 10 4 T 2 + 0.66T 0.25


X pH = 0.35 pH 2 + 2.3 pH 2.7

X O2 = 1
X O2 =

PO2
0.01

(2-13)

(2-14)

for

PO2>0.01

(2-15)

for

PO2<0.01

(2-16).

Jaynes et al. (1984) do not give actual values of k and hence their approach can only be used
to compare rates under different conditions.

In the RATAP-code, Scharer et al. (1994) used the rate law:

m RTa [O2 ]
1
e
2.5 pH
K o + [O2 ] 1 + 10
+ 10 pH 4
Yx / s
E

Rbio = b

(2-17)

where b is a biological scaling factor, Ea is the activation energy, Ko is the half saturation
constant for oxygen, YX/S is the growth yield, m is the specific growth rate, and is the
specific surface coverage. Equation 2-17 explicitly accounts for the dependence of the biotic
oxidation rate on temperature, oxygen concentrations and pH through the terms e
Arrhenius equation),

[O2 ]
K o + [O2 ]

(Michaelis-Menten kinetics), and

Ea
RT

(the

1
1 + 10 2.5 pH + 10 pH 4

respectively.
Scharer et al. (1994) suggested that the specific growth rate depends upon several parameters
such as PCO2, water content, and the availability of nutrients. However, no reference is given
and no numerical value is suggested for this parameter, or for the growth yield or the specific
surface coverage. As these parameters are not available, it seems reasonable to simplify
Equation 2-17 to:

Rbio = b' k Bio


0
where k Bio = k Bio e

[O2 ]
1
2.5 pH
K o + [O2 ] 1 + 10
+ 10 pH 4

(2-18)

Ea
RT

, k0Bio is a rate constant at a defined standard state, and b is a scaling


factor that accounts for the dependence of Rbio on factors not explicitly handled by Equation
2-18, i.e. deviation in microbial activity from the defined standard state. By defining the
conditions of the experiments reported in the Table 2 in Scharer et al. (1994) as our standard
state (b=1), their results can be used to estimate kBio(T=30oC). For pyrite, chalcopyrite and
sphalerite, we estimate kBio to be 9.4 x 10-8, 9.5 x 10-9 and 1.1 x 10-8 mol m-2s-1, respectively.
Figure 2-3 shows that in the temperature range 5<T<30 oC, the modelling approaches used by
Scharer et al. (1994) and Jaynes et al. (1984) give similar temperature dependence of the
biotic oxidation rate. According to Jaynes et al. (1984), no data are available for lower
temperatures. At higher temperature the models deviate, where the model of Jaynes et al.
(1984) shows a decline in rate with temperature. This reflects that the optimum temperature
for the microbes has been exceeded. The model of Scharer et al. (1994), on the other hand,
shows the exponential increase of rate with temperature that is expected for an abiotic
chemical reaction. The different approaches give similar dependence on oxygen
concentrations and pH (see Figure 2-3 b,c). For the dependence on oxygen concentration the
models agree best if a value of 1 x 10-6<Ko<1 x 10-5 is assumed in Equation 2-18.

a)

b)

c)
Figure 2-3. The dependence of the biotic pyrite oxidation rate on a) temperature; b)
pH; and c) oxygen concentration according to models by Jaynes et al. (1984) and
Scharer et al. (1994) (see text).

2.2.3

Mechanism predominance for pyrite oxidation

By combining the abiotic rate expressions for pyrite oxidation by O2 (Equation 2-10) and by
Fe(III) (Equation 2-11), a fraction diagram was generated for the different abiotic pyrite
oxidation mechanisms (see Figure 2-4). This diagram shows that at low pH, oxidation of
pyrite by Fe(III) contributes significantly to the overall pyrite oxidation rate, or even
dominates over the rate of oxidation by oxygen, depending on the partial pressure of oxygen
and the ferric iron concentration.

Rate of ox'n of pyrite by Fe


/Overall pyrite ox'n rate

3+

1.2
NB: Only hydrolysis
complexes accounted for.

Fe(OH)3 eqm

0.8

3+
M
[Fe ]tot=1

0.6
0.2 atm

0.4

0.02 atm
0.002 atm

0.2
0
1

4
pH

Figure 2-4. Relative importance of the abiotic ferric iron oxidation mechanisms as a function
of pH for different ferric iron concentrations and partial pressures of oxygen (see Appendix B
for details).

We have also attempted to frame regions of pH and oxygen concentration where the abiotic
and the biotic reaction paths dominate by comparing Equations 2-10, 2-11 and 2-18 (see
Appendix B). However, these calculations should be considered preliminary, as quantitative
data on the biotic kinetics, from which relevant rate laws and constants can be derived, are
sparse in the literature. As noted in Section 2.2.1, abiotic rate laws from the literature predict
process rates that differ over orders of magnitude, implying that the choice of the rate law is
critical for all model results. It should also be noted that kinetic expressions derived by
different methods may be inconsistent, which may flaw comparison.
Figure 2-5 shows the derived predominance areas of the various biotic and abiotic
mechanisms. Consistent with what has previously been reported in the literature, this
comparison indicates that for a wide range of conditions, the microbially mediated
mechanism may dominate. Note, however, that the dominance area is strongly dependent on,
for example, the scaling factor b and the assumed value of Ko.

10

a)

Figure 2-5. Mechanism predominance diagram for pyrite oxidation at T=1oC and
Ko=1 x 10-6 M. The full and dotted lines denote b=1 and b=0.01, respectively. The arrows
indicate the response to an increase in the adjacent printed parameter. a) Ferric iron
concentration assumed to be controlled by an amorphous Fe(III)-hydroxide;
b) [Fe(III)]tot=1 x 10-6 M (only hydrolysis species accounted for). See Appendix B for details.

11

2.3 Ferrous iron oxidation


In the presence of oxygen, ferrous iron produced by sulphide oxidation or other reactions can
be oxidised to ferric iron in a process that is limited by chemical kinetics:

1
1
Fe 2+ + O2( aq ) + H + Fe 3+ + H 2 O
4
2

(2-19)

where hydrolysis of the iron ions is not shown in Equation 2-19. Oxidation may occur via
three different paths; abiotic homogeneous oxidation, microbial oxidation and surface
catalysed oxidation. The Fe(III) produced may subsequently form complexes with other ions
in solution. If the resulting solution is oversaturated with respect to a secondary mineral, for
example, amorphous ferric hydroxide, the ferric iron concentration in solution will be
controlled by the precipitation of this phase:
Fe 3+ + 3H 2 O Fe(OH ) 3( am ) + 3H +
(2-20)
where this reaction is often assumed to be at equilibrium.

2.3.1

Abiotic homogeneous oxidation

The kinetics of the homogeneous oxidation of ferrous iron has been extensively studied in the
literature. Lowson (1982) reviewed the kinetics over a wide range of conditions and Millero
(1985) summarised data valid for natural waters. As given by Millero (1985), the reaction rate
is zero order with respect to hydroxyl ions and first order with respect to both ferrous iron and
oxygen concentrations at pH<2:
-

d Fe 2 +
= k [Fe( II )][O2 (aq )]
dt

(2-21).

At 2<pH<5 the rate is first order with respect to ferrous iron, hydroxyl ions, and oxygen:

d Fe 2 +
= k [Fe( II )] OH [O2 (aq )]
dt

(2-22)

and at 5<pH<8, the rate is first order with respect to ferrous iron and oxygen but second order
with respect to hydroxyl ions:
-

d Fe 2 +
= k [Fe( II )] OH
dt

] [O (aq)]
2

(2-23).

Wehrli (1990), based on Millero (1985), interpreted the rate dependence of the reaction as
being due to oxidation of different ferrous iron species, that is, the aqua ion at low pH and the
first and second hydrolysis species at higher pH. This implies oxidation of the different
species via parallel mechanisms, for which the total rate, Rhom, is given by the sum of the
contributions from each mechanism:

] [

Rhom = k o Fe 2 + + k1 FeOH + + k 2 [Fe(OH ) 2 (aq )] [O2 (aq )]

(2-24).

where ki are rate constants belonging to the different aqueous species.


Figure 2-6 shows the rate of reaction, the relative importance of the different terms in
Equation 2-24, and the speciation of ferrous iron as a function of pH. This figure shows the
importance of the hydrolysis species for the dramatic increase in oxidation rate with
increasing pH.

12

Millero (1985) discussed how homogeneous reaction rates depend on the medium in which
the reaction occurs. He showed that effects are due to either non-specific ion interactions,
dependent on ionic strength, or on specific ion interactions, e.g. complex formation. This
implies that empirical rate laws should be expressed in terms of ion activities rather than
concentrations. For ferrous iron oxidation in mine wastes in particular, this implies that the
rate of oxidation is slower at high ionic strength or in the presence of ligands that complex
ferrous iron, e.g. sulphate. On the other hand, Cu2+, has been shown to have a catalytic effect
on the reaction (see e.g. review by Pesic et al., 1989).

-4

Fe(OH)2 (aq)

-6

FeOH

2+

Fe

-8
-10

0.75

-12
2

pH

0.5

2+

Fe

1
0.25
+

FeOH

0
2

pH

pH

Figure 2-6. Ferrous iron oxidation and speciation in solution as a function of pH. Left:
Relative importance of the different species for the overall rate of reaction. Upper right: The
total rate of reaction as function of pH. Lower right: The speciation of ferrous iron in
absence of complexating ligands.

2.3.2

Surface mediated oxidation

Based on the observations of, for example, Tamura et al. (1976), that autocatalytic oxidation
of ferrous iron occurs subsequent to ferric iron precipitation (see also Sung and Morgan,
1980), Wehrli (1990) pointed out that surface-mediated oxidation may be an important path
for oxidation. Such oxidation is assumed to occur via a three-step mechanism:
Adsorption:

>FeIIIOH + Fe2+

Electron transfer:

>FeIII-O-FeII +

>FeII-O-FeIII +

(2-26)

Desorption:

>FeII-O-FeIII + +H+ +H2O >FeIIOH + FeIII(OH)2+

(2-27)

>FeIII-O-FeII + + H+

(2-25)

where the catalytic effect is due to the fast electron transfer in the surface complex (Wehrli,
1990). Adsorbed Fe(II) oxidation competes with homogeneous oxidation as a parallel
mechanism, where the total rate of reaction is determined by the amount of Fe(II) adsorbed as
well as by the rate constant for the heterogeneous reaction. The adsorption of Fe(II) can be
simplistically described by the equilibrium expression:
K=

{Fe O Fe }[H ]
{Fe OH }[Fe ]
II +

III

2+

III

13

(2-28)

where curved brackets refer to surface concentrations. The adsorption of ferrous iron on
minerals has not been extensively studied in the literature. However, Zhang et al. (1992)
report experimental results and a surface complexation model for the adsorption of ferrous
iron onto lepidocrocite (-FeOOH(s)).
Based on the results of Tamura et al. (1976), Wehrli (1990) suggested the rate expression

rSurf = k Surf Fe III O Fe II + [O 2 (aq )]

(2-29)

and gave a tentative rate constant.


For a heterogeneous system with a solid phase concentration of A/V [m2 l-1], Equation 2-29
can be converted to:
R Surf =

A
k Surf Fe III O Fe II + [O 2 (aq )]
V

(2-30)

where A refers to surface area (m2) and V to solution volume (l).

2.3.3

Microbially mediated oxidation

The third mechanism for ferrous iron oxidation involves microbial mediation (see e.g. reviews
by Evangelou and Zhang, 1995, Nordstrom and Southam, 1997; and Nemati et al., 1998).
Although several bacteria have been reported to be active ferrous iron oxidisers, Thiobacillus
ferrooxidans has been most extensively studied. For fully aerated conditions at pH of about 2
and temperature between 20 and 35 oC, several authors have suggested a modified MichaelisMenton-type rate expression for the ferrous iron oxidation (e.g. Lacey and Lawson, 1970;
Nyavor et al., 1996; Nemati and Webb, 1997). Nyavor et al., (1996) report a switch to
pseudo-first-order kinetics at high cell concentrations and an inhibiting effect of ferric iron.
Pesic et al. (1989) quantified the ferrous iron oxidation rate in the presence of various
concentrations of bacteria in sulphate media and reported empirical rate laws that can be
written as:
Rbio = k bio Cbact . Fe 2 + [O2 (aq)] H +

[ ]

for pH>2

(2-31)

for pH<2

(2-32)

'
Rbio = k bio
C bact . Fe 2 + [O 2 (aq)]

where Cbact is the concentration of bacteria in solution (mg l ), k Bio = k


-1

'0
k 'Bio = k Bio
e

Ea
RT

0
Bio

Ea
RT

and

where Ea is the activation energy, and k0Bio and k0Bio can be obtained from

the data in Pesic et al. (1989)1 by accounting for the Henrys law solubility of oxygen in
water.

Note that Pesic et al. (1989) give inconsistent rate laws for pH<2 and pH>2. Therefore we have used
rate constants obtained by linear regression of the data in their Figures 8, 11 and 12. Also note that
Po2=0.21 atm in all experiments and thus that the first-order dependence on oxygen depicted by
Equations 2-31 and 2-32 has not been confirmed in the experiments.

14

2.3.4

Mechanism predominance

Equation 2-24 can be combined with Equation 2-30 and used to compare the surface mediated
ferrous iron oxidation rate with the homogeneous reaction at abiotic conditions. Such a
comparison (Figure 2-7, solid lines) shows that in the pH range 4<pH<7, the surface catalysed
reaction dominates over the homogeneous reaction even at low bulk surface site
concentrations.
At low pH, the importance of the heterogeneous path is small due to insignificant ferrous iron
surface adsorption (see Appendix C; see also Sung and Morgan, 1980). At high pH, the
homogeneous reaction rate increases due to formation of ferrous iron hydrolysis species, and
the heterogeneous reaction becomes less important. Note, however, that the mechanism
predominance diagram is based on the adsorption constant for ferrous iron on lepidocrocite
only, and that the heterogeneous rate constant for the associated oxidation reaction is
uncertain by several orders of magnitude.
Equation 2-31 has been used to compare the biotic with the abiotic homogeneous and
heterogeneous ferrous iron oxidation rates (Equations 2-24 and 2-30, respectively) and to
frame regions where the different mechanisms dominate. Figure 2-7 shows a mechanism
predominance diagram for different concentrations of bacteria and dissolved ferrous iron.
This diagram indicates that the microbial path is dominant at low pH, whereas the abiotic
mechanisms (heterogeneous and homogeneous) dominate at higher pH. The pH where the
switch from biotic to abiotic control occurs depends on the bacteria concentration as well as
on the concentration of surface sites. For the reasons mentioned above and in Section 2.2.3,
this diagram should only be seen as tentative.

Figure 2-7. Mechanism predominance diagram for ferrous iron oxidation at T=25oC (only
hydrolysis species accounted for). Solid lines represent the abiotic system. Dotted lines
denote the switch between abiotic and biotic control for various bacteria and ferrous iron
concentrations. The lines for 10-7<Cbact<10-4 g l-1 have been truncated in order to facilitate the
reading of the figure. See Appendix C for details.

15

2.4 pH buffering reactions


Oxidation of iron sulphides can lead to the release of protons, as indicated by Equation 2-8. In
a tailings impoundment a series of pH-buffering reactions occur with the gangue materials
present, which can include carbonate, hydroxide and silicate minerals (Blowes and Ptacek,
1994). First to dissolve are carbonates, which react at a greater rate than sulphides, as shown
in Figure 2-8. Carbonates react at a great enough rate that solubility equilibrium with aqueous
solution can often be assumed. Dissolution of carbonate minerals buffers pH around neutral,
where many metal hydroxides can precipitate due to low solubility, with the associated
removal of heavy metals from solution and production of protons (for example, see Equation
2-20, ferric hydroxide solubility equilibrium).
With continued acid production, depletion of carbonates leads to a drop in pH, after which the
hydroxides begin to dissolve and buffer pH; these are also eventually depleted. Remaining are
silicate minerals, which usually have a greater buffering capacity but react more slowly than
carbonates, hydroxides and sulphides. The solution is thus still buffered, but at a lower pH.

-2

log (rate) mol m -2s-1

-4
calcite (atmospheric Pco2)*
calcite (Pco2=30%)*

-6
-8

pyrite (Po2=0.1 atm)**

albite (1)

-10

anorthite (2)
biotite (3)

-12

chlorite (4)
K-feldspar (5)

-14

muscovite (6)
2

6
pH

10

Figure 2-8. Weathering rates of silicates, pyrite and calcite at 5C. Anorthite and K-feldspar
have almost identical rates for the pH range shown. References:
1)
2)
3)
4)

Wollast &Chou, 1985


Amrhein & Suarez, 1988
Malmstrm & Banwart, 1997
Malmstrm et al., 1995

2.4.1

5)
6)

Helgeson et al., 1984


Knauss & Wolery, 1989

*
**

Chou et al., 1989


Williamson & Rimstidt, 1994

Carbonate weathering

Carbonate weathering can be exemplified by the dissolution of calcite as given below, here
shown at equilibrium:
CaCO3( s ) Ca 2+ + CO32
(2-33)
Protonation of the carbonate ion at pH 10 leads to increase in solution pH.

16

As can be seen in Figure 2-9, dissolution rates of the carbonates calcite, dolomite
(CaMg(CO3)2(s)) and magnesite (MgCO3(s)) increase with decreasing pH, becoming then
transport limited. At the pH of natural waters, dissolution and precipitation are surfacecontrolled; however, at the lower pH of a typical tailings impoundment, solubility equilibrium
may be a reasonable assumption, until the available carbonate is consumed.
Calcite is an efficient buffer of pH in the neutral pH range, buffering the pH around 6.5 - 7.5
(Blowes and Ptacek, 1994). The related increase in pH can lead to the precipitation of
hydroxides and carbonates with lower solubility than calcite, for example siderite, FeCO3(s),
the dissolution of which in turn buffers solution pH around 4.8-6.3 once the calcite is
consumed.

Figure 2-9. Rates of carbonate dissolution, 25C (modified from Wollast, 1990).

If the assumption of equilibrium is applicable, the solubility of calcite will depend on the
solubility product, Ksp =[Ca2+][CO32-]. Aqueous speciation and the precipitation of secondary
minerals such as gypsum, CaSO42H2O(s), can affect free calcium concentrations; free CO32concentrations are dependent upon the partial pressure of CO2(g) and pH. If these factors are
taken into account, the relationship between field pH and Ca2+ concentrations, as well as the
measured (or estimated reasonable) Pco2, can be used as a rough guide as to whether calcite
is at equilibrium with porewater concentrations.
If equilibrium cannot be assumed, rate expressions for calcite dissolution are available. Chou
et al. (1989) gave an expression for the surface area normalised rate of dissolution:

rate = k1aH + + k 2 a H CO* + k3 a H 2O k 3 aCa 2+ aCO2


2

(2-34)

where ki are rate constants and aj are activities (see also Appendix E). The first three terms
represent the forward rate of Equation 2-33, that is, dissolution; the fourth term represents the
rate of the reverse reaction, precipitation, where the reverse reaction only becomes significant
for pH > 8.

2.4.2

(Hydr)oxide weathering

High concentrations of many metals are often found in tailings impoundments, for example,
iron as a result of iron sulphide oxidation, and Al due to weathering of aluminosilicates.
Suitable conditions of pH and concentration can lead to the precipitation of secondary
hydroxides such as amorphous Al(OH)3(s), amorphous Fe(OH)3(s), ferrihydrite (poorly
crystalline form of hydrous ferric oxide/hydroxide), gibbsite (Al(OH)3(c)) and goethite
(FeO(OH)(s)), as well as less soluble hydroxysulphate minerals such as jarosite
(KFe3(SO4)2(OH)6(s)) and alunite (KAl3(SO4)2(OH)6(s)). After depletion of carbonate minerals

17

and the subsequent drop in pH, dissolution of amorphous Al(OH)3(s) can buffer solution pH
around 4.0-4.3. Consumption of aluminium hydroxide leads to a drop in solution pH to below
3.5 and dissolution of ferric hydroxide (Blowes and Ptacek, 1994).
As indicated by Equation 2-20, dissolution and precipitation of amorphous hydroxides is
often considered sufficiently fast that solubility equilibrium control over aqueous
concentrations can be assumed, according to the following equation:

K=

[Fe ]
[H ]
3+

(2-35)

+ 3

This equilibrium affects and is affected by the pH of solution. Solubility of secondary


minerals is increased by complex formation in the aqueous phase; in particular free ferric iron
concentrations are strongly affected by hydrolysis of the Fe(III) ion.

2.4.3

Silicate weathering

The reaction for chlorite dissolution is given below as an example of silicate weathering:

(Mg

4.5

Fe0II.2 Fe0III.2 Al AlSi3 O10 (OH ) 8 ( s ) +

77 +
H
5

4.5Mg 2+ + 0.2 Fe 2+ + 0.2 Fe(OH ) 3 ( s ) + 2 Al 3+ + 3SiO2 ( s ) +

57
H 2O
5

(2-36)

As shown, weathering of silicates leads to the consumption of protons and release of metal
cations. As shown in Figure 2-8, most silicates display pH-dependent weathering with a
minimum rate around neutral pH. The rate of reaction is often expressed as the sum of three
terms

[ ]

R = k1 H + + k 2 + k3 OH

(2-37)

where the first and last term represent acid and base catalysis of the reaction, respectively. As
mentioned above, silicate minerals tend to have a greater buffering capacity than carbonates
and hydroxides, but weather at a slower rate due to reaction kinetic limitations.
See Appendix D for rate expressions obtained from the literature and utilised in the current
model, including those for silicate weathering.

18

3 Model description
The current model for the Kristineberg tailings impoundment is based on modelling by
Strmberg and Banwart (1994), where the steady state, box model code STEADYQL was
applied to waste rock heaps at the Aitik site in northern Sweden. In Section 3.1 the code is
briefly described and discussed; Section 3.2 covers how the conceptual model for
Impoundment 1 in Kristineberg was formed and the mathematical expressions and database
used; Section 3.3 describes calibration to field data.

3.1 The STEADYQL Code


3.1.1

Code description

STEADYQL (Furrer et al., 1989, 1990) is a code for box models where the system being
modelled is treated as a completely mixed flow-through reactor. Processes can occur over
three time-scales: i) fast processes, controlled by chemical equilibrium, ii) slow processes,
limited by kinetics, and iii) relatively very slow processes, such as movement of fronts and
consumption of minerals, which are considered not to occur over the time period that the
modelled steady state applies for.
Equilibrium processes are accounted for by the mass action equation:

C( j)

C (i ) = K (i )

a (i , j )

(3-1)

where C(i) is the free concentration of specie i [mol dm-3], K(i) is the conditional stability
constant of specie i, and a(i,j) is the stoichiometric coefficient of component j in species i
(compare with Equations 2-1 and 2-2). Adsorption reactions are accounted for by a simple
surface complexation model, by including an immobile component in Equation 3-1. Solubility
equilibria can be expressed by a modified version of Equation 3-1, where activity of the solid
phase(s) is fixed, forcing the affected free aqueous concentrations into equilibrium with the
phase(s). If the system without the solubility restriction would have been over- or
undersaturated, this approach conceptually corresponds to precipitation or dissolution,
respectively, of the solid phase.
The rate of any slow process is expressed as a power law of a number of parameters and free
concentrations of chemical species:
R (l ) =

P ( m)

w( l , m )

C (i)

n ( l ,i )

(3-2)

where R(l) [mol dm-2s-1] is the rate of the process, P(m) is the value of parameter m, w(l,m) is
the exponent of parameter m in process l, and n(l,i) is the reaction order the chemical species i
in the rate expression for process l.
For any component the total flux J [mol dm-2s-1] is given by the equation:
J (tot ) =

R(l ) s (l , j ) v
l'

a(i ' , j ' )C (i ' )

(3-3)

i'

where l represents all processes but the outflow, i and j denotes mobile components, s(l,j)
denotes the stoichiometric coefficient of component j in process l, and v is the outflow

19

velocity [dm s-1] (referred to as infiltration rate q by Strmberg and Banwart, 1994, and in the
following sections). In Equation 3-3 the second term represents the outflow. Note that the
model mathematically describes flows in terms of area normalised fluxes.
The model runs to steady state, such that for all components Equation 3-3 equals zero. The
STEADYQL program is dimensioned for up to 60 species, 20 components, 30 processes and
33 kinetic parameters. Figure 3-1 shows a schematic example of how the code can be applied;
the model shown is of Impoundment 1 at the Kristineberg site, as described in the following
sections.

Figure 3-1. Steady state box model of a tailings impoundment.

3.1.2

Comments on the STEADYQL code

Although real systems are rarely at steady-state, this may be a more accurate approach than
the common assumption of full equilibrium, which also lacks time resolution but makes no
allowance for reactions limited by chemical kinetics. In order to account for changing
conditions over a long time, such as consumption of minerals, it is possible to compare steady
states representing time periods where different processes or conditions dominate. For
example, in order to model the effect of the consumption of amorphous Fe(OH)3(s) in a system
where dissolution of this specie is buffering pH, the progression can be considered as two
subsequent steady states, a) in the presence of, and then b) in the absence of, the hydroxide.
In many computer codes, reactions are chosen automatically by the code from an
accompanying database after the input of desired components by the user. In STEADYQL,
the reactions to be represented must be specifically chosen, as does the entire database,
leaving it up to the user to ensure that the dominant reactions are accurately entered and
represented. For example, unless an equilibrium is assumed and specifically included in the
database, precipitation of secondary minerals will not occur, even if the solution is
oversaturated with respect to the mineral(s). On the other hand, this aspect of the code gives
the user full control over which reaction will occur in each simulation, and which rate
expression is used.

20

Models based on the STEADYQL code can represent advective transport; however, there is
no possibility to represent diffusive transport. As a box model the system is assumed to be a
continuously stirred tank reactor (CSTR). There is thus no time or space resolution; it is not
possible to simulate movement of fronts. One possibility, for example to represent diffusion
of a gas into the box, is to simulate the kinetic addition of the gas to the box; however, there is
no coupling between the rate of diffusion into the box and the oxygen concentration gradient
created by the consumption of oxygen below the box (see Section 4.6). A possible
representation of zones with different characteristics could be via linking inflow and outflow
of sequential boxes, and repeated calculation of steady state (e.g. Berg, 1997). Such an
approach has also been used to represent the movement of fronts in the commercial tailings
impoundment model RATAP (Scharer et al., 1994).
Moreover, it is not possible to represent Monod-type kinetics, such as the Michaelis-Menten
expressions often used to represent biological kinetics (see e.g. Equation 2-17). Instead, such
expressions have to be approximated by a fractional dependence on the variable parameter, or
simple scaling factors applied to the reactions affected by the biological catalysis.
Additionally, surface speciation reactions can be represented, but only with simple mass
action expressions.
The advantage of this type of steady-state model, for example over a pure geochemical
equilibrium model, is the simple coupling of porewater geochemistry with slow, kinetically
limited processes, showing the interdependence between these processes, and in particular the
effects of and on the system master variable pH.

3.2 The model applied to Impoundment 1 at the


Kristineberg site
Based on the general concepts of the processes mentioned in Section 2, as well as field data, a
model was built up of a tailings impoundment. Field data was obtained from Impoundment 1
at the Kristineberg site, Sweden, before remediation, which occurred in 1996 (Lindvall et al.,
1999). The model represents the top 1 m of the impoundment, which is approximately the
extent of the pre-remediation weathered zone. The main aim of the modelling was to explain
the main ion composition of the groundwater, in order to then use the model for the prediction
of the effect of remediation. Described in the sections below is the base case conceptual
model, as well as the geochemical database and input parameters that make up the
mathematical formalisation of the conceptual model. In Section 3.3, the process in which the
model was calibrated to site data using two scaling factors is described. Variations on the
base case conceptual model, tested by the addition or removal of various processes, are
described in Section 4.

3.2.1

Base case conceptual model

The conceptual model for the site was created through consideration of sources and possible
sinks for the aqueous ions. The processes represented in the base case model were selected
as described below. The model was developed through an iterative process, in which
comparisons between variations on the base case were used to refine the conceptual model.
Variations tested in the conceptual model are described in the Section 4.
Pre-remediation field data has been taken from Malmstrm et al. (1999b) who summarised
and compared data from du Rietz (1953), Qvarfort (1983, 1989), Ekstav and Qvarfort (1989),
Axelsson et al. (1986, 1991) and others. Available data from pre-remediation studies of the
impoundment includes analysis of the chemical composition of a few solid phase profiles,

21

mineralogical determination of the gangue material, mineralogy of the Kristineberg mine2,


and groundwater analyses from within the impoundment.
The 1 m weathered zone at the top of the impoundment has been modelled, where the average
groundwater level was 1 m with an annual variation of 1 m (Axelsson et al., 1986). The total
depth of the impoundment ranged from a few meters up to 11 m. The average depth of the
weathered zone in Impoundment 1 was also 1 m (see compilation in Malmstrm et al. 1999b).
As concentrations in the saturated zone did not vary greatly with depth (Ekstav and Qvarfort,
1989), the unsaturated zone is conceptualised as the major source of ions in solution (Banwart
and Malmstrm, 1999). The volume fraction of minerals in the modelled box were calculated
by combining the reported mineralogy with the average of the solid phase compositions at the
top and bottom of the 1 m weathered zone. According to the aqueous phase analyses, the
major ions in solution in the field were SO42-, Fe, Zn, Ca and Mg (Figure 3-2).

1.E+00

6.5

1.E-01

1.E-02

5.5

pH

Concentration (mol/l)

average

1.E-03

1.E-04

4.5

1.E-05

1.E-06

3.5
SO42-

Fe

Mg

Ca

Zn

Al

Na

Cu

pH

Figure 3-2. Groundwater composition in Impoundment 1, Kristineberg showing minimum,


maximum and average values. Data are taken from Ekstav and Qvarfort (1989), and are
averages from a time series of measurements over 5 years, 1989-1983, from Piezometers 3, 4,
and 4H. The depth of sampling was 1.5m below the groundwater table, which lay on average
1m below the surface of the impoundment (see Malmstrm et al., 1999b).

Preliminary geochemical modelling with the program PHREEQC (Parkhurst, 1995) using the
MINTEQ database (see Section 2.1.1) was carried out on groundwater analyses from
Impoundment 1, in order to obtain an indication of which primary and secondary minerals
may have been close to thermodynamic equilibrium with porewaters, and thus controlling
aqueous concentrations. However, care must be taken in interpretation of geochemical
equilibrium modelling, due to conditions of high ionic strength and incomplete chemical
analysis; for example, no anion analyses other than sulphate are available. Moreover, preremediation aqueous concentrations were reported only as average, minimum and maximum
values for each specie; no entire analysis was reported. Geochemical equilibrium modelling
with average values may not represent the groundwater geochemistry accurately; for example,

Tailings in Impoundment 1 came not only from the Kristineberg mine but also mines in the
surrounding region.

22

average values of the sulphate concentration and pH may be used, but the sulphate
concentration may have been low in analyses with high pH, or vice versa.

3.2.1.1

Sulphate, Zinc and Copper

The main ore minerals of the Kristineberg mine are reported to be pyrite, chalcopyrite and
sphalerite (du Rietz, 1953). Oxidation of these minerals was thus considered to be the source
of the aqueous ions SO42-, Zn2+ and Cu2+. The Kristineberg mine itself was reported to be poor
in pyrrhotite, however the tailings in Impoundment 1 come from surrounding mines as well,
so it is possible that pyrrhotite is present. Comparison of aqueous Fe(tot) and SO42concentrations from field data indicated that the SO42- to Fe(tot) ratio was too low to be due to
the stoichiometric oxidative dissolution of pyrite by oxygen alone. One possibility was a
portion of the iron sulphide is present as pyrrhotite, however, as mentioned, pyrrhotite was
not named in the original mineralogical assay. Additionally, analysis of solid phase
compositions showed that the ratio of iron to sulphur bound in iron sulphides is 0.92, which is
greater than the theoretical ratio of 0.87 for pyrite, but lower than the expected >1.7 for
pyrrhotite (Malmstrm et al., 1999b). In lieu of other information, this suggested that the
majority of the iron sulphide is in the form of pyrite rather than pyrrhotite.
Two other possibilities to explain the low aqueous SO42- to Fe(tot) ratio were a) the
immobilisation of sulphate through, for example, precipitation of gypsum, and/or b) the
dissolution of a solid Fe-bearing minerals, such as chlorite and/or amorphous ferric hydroxide
(see Section 3.2.1.2 below). Geochemical modelling indicated that gypsum was close to
solubility equilibrium in the porewaters, and hence may have affected Ca2+ and SO42concentrations (see Sections 4.4 and 4.5 for testing of the model with gypsum present).
However, as the presence of gypsum could not be confirmed, it was not included in the base
case.

3.2.1.2

Iron

According to the reported mineralogy and the analysis of the solid phase chemical
composition, possible sources of the very high iron concentrations in the aqueous phase were
oxidation of iron sulphides pyrite and chalcopyrite, as well as dissolution of the silicate
mineral chlorite. Other possible sources included the dissolution of secondary minerals.
Possible sinks of ferrous iron considered were precipitation of secondary minerals, for
example via oxidation of ferrous to ferric iron and subsequent precipitation of amorphous
Fe(OH)3(s).
As no redox measurements were made in the impoundment before remediation, the
distribution of iron between ferric and ferrous forms is unknown. Ferric iron is the more
thermodynamically stable redox form of aqueous iron at the oxygen concentrations assumed
for the modelling, Po2=0.1atm. However, the high iron concentration at the high pH found in
the impoundment (up to pH 6.15) suggested that most iron is in the form of ferrous iron, as
ferric hydroxide becomes highly insoluble as pH approaches neutral. High iron concentrations
thus indicate that there may be a kinetic barrier to oxidation of Fe(II) to Fe(III) due to slow
reaction kinetics, hence the importance of representing this process in the model by a kinetic
expression. Additionally, it has been reported in the literature that the rate of abiotic oxidation
of aqueous ferrous iron to ferric iron by oxygen decreases greatly with decreasing pH, but that
at low pH the reaction can be accelerated by bacterial catalysis, if conditions are otherwise
suitable; however, this catalytic effect was tested in the model and found to be unlikely in the
field (see Section 2.3).

23

Although redox values for the impoundment are not reported, the modelling showed that
within pe values of 4-7 3 amorphous ferric hydroxide may be at solubility equilibrium, and
hence exerting control on aqueous ferric iron concentrations. According to pre-remediation
field reports (Qvarfort, 1983), the upper layers of the impoundment displayed the
characteristic colour of ferric (oxy-)hydroxides, supporting the theory that some form of
secondary Fe(III) mineral was present. For the base case, it was assumed that ferric hydroxide
was present and exerting solubility control on the ferric iron concentration. The model was
also tested without the hydroxide present, see Section 4.1.
3.2.1.3

Base cations

High aqueous concentrations of Ca2+ and Mg2+, as well as an average pH of 4.87, suggested
that some form of buffering reaction was occurring in the impoundment. As shown in Table
3-1, mineralogical reports indicate that calcite and dolomite constitute 10 vol-% of the gangue
material at Kristineberg, and these were the only reported minerals containing calcium.
However, the chemical composition of the tailings indicated that the carbonate content was
lower than this, and that at least half the calcium present was bound in non-carbonate minerals
(Malmstrm et al., 1999b). Additionally, investigations with the model showed that
equilibrium with calcite could not result in field pH and calcium concentrations, even with
elevated CO2(g) partial pressures (see Section 4.5.1). Possible reservoirs of Ca2+ other than
carbonates could be an unreported Ca-bearing silicate such as plagioclase
((Na,Ca)Al(Al,Si)Si2O8(s)), or secondary gypsum, where gypsum could act as either source or
sink for calcium, depending on the over-/undersaturation of solution with respect to this
mineral.
Table 3-1. The average composition of the waste rock fraction in the
Kristineberg ore (data from Qvarfort, 1983).
Mineral
Chemical composition
vol-%
Chlorite
(Fe,Mg,Al)4-6(Si,Al)4O10(OH)8(s)
50
Talc
Mg3Si4O10(OH)2(s)
Quarts
SiO2(s)
25
Sericite
KAl2(AlSi3)O10(OH)2(s)#
15
Calcite
CaCO3(s)
10
Dolomite
CaMg(CO3)2(s)
#
Different definitions given in the literature; resembles an Fe(III)-substituted
muscovite (Malmstrm et al., 1999b, interpreting data from du Rietz, 1953)

Investigations with the model showed that three of the potential sources of Ca2+ were feasible;
kinetic dissolution of a small volumetric fraction of calcite (see Section 4.5); gypsum
dissolution (see Section 4.4), or dissolution of an unreported Ca-bearing silicate (see Section
4.5). Thus release of Ca2+ in the field could be the result of one or a combination of the above
possibilities. However, the prevailing source could not be identified. To allow for a
conservative buffering capacity of the weathering of the unknown calcium source, total
magnesium concentrations, originating from chlorite weathering in the model, were allowed
to be equal to the sum of Mg2+ and Ca2+ concentrations in the field.
High concentrations of Mg2+ in the aqueous phase indicated weathering of dolomite and/or
Mg-bearing silicate minerals. However, in the mineralogical analysis of gangue material,
separate volumetric fractions were not given for calcite and dolomite (see Table 3-1), and
calculations indicated that solubility equilibrium with dolomite was not feasible. It was
assumed that the source of Mg2+ was silicate weathering.
3

Typical conditions encountered within the mid-upper layers of an impoundment, e.g.


Appleyard and Blowes (1994); pe can vary between 1-14 for an entire impoundment. The
resulting pe in our model varied from 8 at atmospheric Po2, to below one at 0.001atm.
24

Mg2+-bearing silicate minerals in the gangue material were reported to be talc and chlorite
(see Table 3-1). However, separate volumetric fractions were not given. Given that only one
tracer (Mg2+) was available for both minerals, that the stoichiometric ratio of H+ consumption
to Mg2+ release for both minerals is similar (see also discussion of aluminium below), and that
there is no rate expression for talc weathering available, the weathering of talc and chlorite
was represented by a single process. This process was assumed to be the weathering of a
chlorite with a composition corresponding to that of chlinochlore, found in the mineralogy of
the mine.
Pre-remediation solid phase analyses of K+ or Na+ are not available; however K+
concentrations in the aqueous phase and the reported mineralogy suggested the weathering of
e.g. muscovite. As no source for Na+ was given in the mineralogy and explanation of the
relatively low Na+ concentrations was not the main aim of the investigation, no process to
account for Na+ release was included in the model.
3.2.1.4

Si and Al

Pre-remediation solid phase analyses of Al or Si were not available. However, high aqueous
Al concentrations suggest considerable aluminosilicate weathering; in the model Al is
released by the weathering of muscovite and chlorite, as given in Table 3-1. Aluminium
concentrations in the model are actually higher than the field concentrations; this may be due
to the representation of the reported talc-chlorite mixture entirely by chlorite, for reasons
discussed in Section 3.2.1.3, as talc does not contain aluminium. In a quick test with talc and
no chlorite, assuming the same weathering rate as for chlorite, the results were the same for
pH and all components other than aluminium, which was too low. As other model results
were insensitive to the assumption of either talc or chlorite as the source of Mg2+, and no rate
law is available in the literature for talc weathering, Mg-aluminosilicate weathering was
represented by chlorite dissolution for the base case.
Equilibrium with the secondary mineral gibbsite was tested as a variation upon the base case,
as gibbsite was close to solubility equilibrium in the preliminary geochemical modelling; on
the basis of this testing and lack of field data, gibbsite was not included in the base case (see
Section 4.2).
Aqueous Si concentrations were not measured prior to remediation, but high aluminium and
base cation concentrations indicated silicate weathering. Dissolved silicon concentrations
were taken from post-remediation measurements (H. Holmstrm, personal communication). It
was assumed that concentrations had not changed greatly since remediation, given the pH
dependency of silicate weathering and that the pH has not changed significantly. Geochemical
modelling, including these Si concentrations, suggested equilibrium with a SiO2 secondary
mineral such as chalcedony, SiO2(am) or quartz. In the model, equilibrium with SiO2(am) is
assumed. However, this assumption is not as significant as other secondary phase
assumptions, as equilibrium with a SiO2-phase does not affect pH or alkalinity.
3.2.1.5

Equilibrium with gas phase

Consistent with a gradientless, steady state formulation, fixed oxygen and carbon dioxide
partial pressures in the gas phase were conceptualised as being in equilibrium with the
aqueous phase. Tested variations on the base case included the absence of CO2 equilibrium,
which had little effect on results (see Section 4.3), and the slow addition of oxygen to
approximate diffusion (see Section 4.6). The test with oxygen diffusion, using the base case
water content of 0.15, resulted in aqueous oxygen concentrations that were similar to those in
equilibrium with the base case partial pressure of oxygen (see Table 4-3), which supported the
chosen representation for the base case.

25

3.2.1.6

Aqueous speciation model

The reactions representing the aqueous speciation of the reported main ions in solution were
taken from literature sources (see Section 3.2.2, and Appendix E for the aqueous speciation
model used).
3.2.1.7

The final base case conceptual model

Figure 3-1 above is a representation of the conceptual geochemical model for the Kristineberg
site. Slow kinetic processes are the weathering of sulphides and silicates and the oxidation of
aqueous ferrous iron; aqueous speciation is assumed to be at equilibrium, and solubility
equilibrium is assumed for secondary minerals SiO2(am) and Fe(III)(hydr)oxide. Variations on
the base case that were tested are summarised in Section 4 below. The reactions, rate and
equilibrium expressions and constants used in the model are given in Appendices D and E.

3.2.2

Geochemical equilibrium and kinetic database used in the model

For reactions limited by chemical reaction kinetics, empirical rate expressions were chosen
from literature where available, or where not available, expressions were assumed by analogy
with similar minerals. The latter was the case for oxidation of chalcopyrite and sphalerite by
oxygen, where analogy was made with oxidation of pyrite (see Appendix D for the
expressions used). Rate constants were corrected to field temperature by the Arrhenius
equation (Equation 2.7) using activation energies reported in the literature (see Appendix D).
Thermodynamic constants for the geochemical database in the Kristineberg model were
obtained from, in order of descending preference, Nordstrom et al. (1990) and the WATEQ
and MINTEQ databases (see Section 2.1.1, and Appendix E for the aqueous speciation model
used). Constants were corrected for field temperature using the Vant Hoffs equation
(Equation 2-5) and for ionic strength using the Davies equation (Equation 2-4), where the
ionic strength was taken as an average value from preliminary geochemical equilibrium
calculations using field data and the model PHREEQC, I=0.27.

3.2.3

Input data

Values and sources of some of the parameters used in the model are given in Table 3-2 (see
also Appendix D). Parameters 1-4 were obtained from pre-remediation field reports.
Parameters 5-7 were estimated based on information in the pre-remediation field reports. The
compact density was compiled from the estimated mineralogy of the impoundment, as given
by Malmstrm et al. (1999b). The water content ( = volume water/total volume) was
estimated from the porosity of the impoundment to be 0.1-0.25 for the unsaturated zone
(Malmstrm et al., 1999b). A middle value of 0.15 was used for in the base case. For
impoundment temperature, the average air temperature of 1C was used, as no measurements
were available.
Oxygen concentrations before remediation were not reported. Atmospheric conditions existed
at the surface (0.21 atm) and probably anoxic conditions at the groundwater table, which lay
at a depth of approximately 1 m, at the bottom of the weathered zone. As the base case
mineralogy was taken as the average composition of the oxidised zone, the base case oxygen
partial pressure assumed to be at equilibrium with the aqueous phase was 0.1 atm.
Replacement of the assumed equilibrium with kinetic addition of oxygen to the box,
approximating diffusion, was tested as a variation upon the base case (see Section 4.6).

26

Equilibrium was assumed with carbon dioxide, for which there also were no field
measurements. The partial pressure taken as the base case value was atmospheric partial
pressure. The base case was also tested without this equilibrium, and with higher partial
pressures of carbon dioxide (see Section 4.3).
Table 3-2: Model parameters. See also text for explanation.
Parameter
Base case value
1) Annual effective infiltration rate, q
7.7 x 10-9 ms-1
2) Average depth of unsaturated zone, h
1m
0.25
3) Porosity,
4) Total specific surface area of the
0.1 m2g-1
tailings material, As,tot
3.3 x 106 gm-3
5) Compact density,
0.15
6) Water content,

7) Impoundment temperature, T

8) Partial pressure of oxygen, Po2


9) Partial pressure of CO2, Pco2
10) Surface site concentration on
secondary ferric minerals, [>FeOH]
Uncertain parameter; see Section 4.7

1C

0.1 atm
10-3.5 atm (atmospheric)
10-9 M

Source
Axelsson et al., 1986
Axelsson et al., 1986
Axelsson et al., 1986
Qvarfort, 1983
(0.1-0.25 m2g-1)
Malmstrm et al., 1999b
Malmstrm et al., 1999b
(0.1-0.25)
estimated from Axelsson
et al., 1986 in Malmstrm
et al., 1999b
estimated by the authors
estimated by the authors
Zhang et al., 1992

The surface area for each mineral, Amin, was calculated by the authors from:
Amin = min x As,tot x specific density
(3-4)
where the volumetric fraction of each mineral, min, was taken to be the values estimated in
Table 3-3, As,tot was as given in Table 3-2 and the specific density is (1-) = 2.48 x 106 gm-3.
Table 3-3. Minerals included in the model. Mineralogy corresponding to the average of
reported chemical compositions at the highest and lowest levels in the weathered zone (depths
of approximately 0 and 1 m, respectively). The remaining material is considered to be
unreactive quartz.
Sulphides
vol.%
Silicates
vol.%
Pyrite
4
Chlorite
53
Chalcopyrite
0.2
Muscovite
16
(sericite)
Sphalerite
0.1

Investigations of the sensitivity of the model to variations in the infiltration rate (q), the water
content () and the partial pressure of oxygen (Po2), as well as variability in the mineralogical
composition (i), are detailed in Section 5. Variations in the partial pressure of carbon dioxide
(Pco2) and the surface site concentration of secondary ferric minerals ([>FeOH]) are discussed
in Sections 4.4 and 4.7, respectively.

27

3.3 Calibration
Model results were compared to the average of monthly averages of the field groundwater
analyses taken from November 1983 to September 1988 from three piezometers (Piezometers
3, 4H and 4, see Table 3-4). Model results were in general remarkably similar to field
measurements. A notable exception to this is that the predicted concentration of Zn2+ deviated
considerably from field values.
Table 3-4. Comparison of field measurements 4 with uncalibrated and calibrated model
results. All concentrations in mol l-1. Field data from Ekstav and Qvarfort (1989) (see also
Figure 3-2).
Range of field
Field monthly Uncalibrated model,
Calibrated model,
values
average
base case
base case
pH
4.05-6.15
4.87
5.41
4.53
SO420.07-0.14
0.1
0.12
0.1
Fe(tot)
0.06-0.09
0.08
0.06
0.07
Cu2+
8 x 10-6- 2 x 10-4
6 x 10-5
3 x 10-5
6 x 10-5
Zn2+
0.004 - 0.007
0.006
5 x 10-6
5 x 10-5
2+

Ca
0.002 - 0.012
0.005
Mg2+
0.007 - 0.017
0.011
0.04
0.017
2+
2+

Mg +Ca
0.009-0.029
0.016
K+
3 x 10-6 2 x 10-4
6 x 10-5
3 x 10-4
1 x 10-4
-4
Al
9 x 10 0.003
0.002
0.02
0.008

As no sources of Ca2+ were included in the model, we allowed Mg2+model = ( Mg2+ + Ca2+)field
to allow for a conservative buffering capacity of minerals that release Ca2+ in the field.

For pH, SO42-, Fe(tot), Cu2+ and K+, the uncalibrated model gave concentrations within the
range of the field values; Mg2+model was higher than the average (Mg2+ + Ca2+)field by a factor
of 2.5 and Al was an order of magnitude greater than the field average. In general these
deviations between model results and field observations are small and can be well accounted
for by the uncertainties in the empirical rate laws employed.
The model was calibrated to average field results. Following the approach outlined by
Strmberg and Banwart (1994), three factors were considered for scaling of kinetic processes:
Xreact, =and .

3.3.1 Calibration factors and sensitivity


3.3.1.1 Calibration factors
The scaling factor Xreact was used to allow for the commonly observed scale dependence of
mineral weathering rates. Field rates are commonly up to three orders of magnitude lower
than rates measured in the laboratory, where empirical rate laws are normally determined; this
may be due to, for example, differences in temperature, particle size distribution, porewater
pH, mineral content and water flow patterns between the scales (Malmstrm et al., 1999a).
In the current model we have attempted to explicitly account for some of the parameters that
have been identified as contributing to such scale dependence. The rate equations taken from
the literature were from laboratory experiments with conditions as similar as possible to the
4

Note that the top 0-1 m of the impoundment, that is, the average depth of the weathered zone, has
been modelled, where the groundwater lay on average at a depth of 1 m. However, the only available
groundwater analyses, shown in Table 3-4 and Figure 3-2, were taken 1.5 m below the groundwater
surface, that is, approximately 2.5 m below the tailings surface.

28

acidic AMD environment. Rate and equilibrium constants were corrected for temperature;
mineralogy was also taken into account, and the resulting pH from the uncalibrated model
was within the range of field values recorded. Xreact is intended to calibrate the model for
additional discrepancies between the scales.
The scaling factor was included to account for direct microbial catalysis of sulphide
oxidation, which has been reported to accelerate the rate of oxidation by up to an order of
magnitude (e.g. see review in Herbert, 1999). In Section 2.2.2 we showed how specific rate
laws can account for microbial kinetics. However, the STEADYQL code does not allow for
such formulations, hence the use of a simple scaling factor.
In a similar fashion as for , and for the same reasons, we have accounted for microbial
mediation of ferrous iron oxidation with another scaling factor, . Microbial mediation is
reported to have the potential to increase the rate of oxidation of Fe(II) by more six orders of
magnitude (Singer and Stumm, 1970; see also Section 2.3.3). However, the base case values
could be brought close to average field values without the use of this factor, and in fact were
only taken further from average field values when this factor was anything other than unity,
hence the value of 1 in the base case.

3.3.1.2 Sensitivity to calibration factors


Model results were quite sensitive to the calibration factors; Figure 3-3 shows the results of
variation of Xreact and ,. It can be seen that change in these scaling factors by less than an
order of magnitude gave concentrations that were outside the range of field values. That we
did not find another solution to the model for field conditions with other values of the scaling
factors suggests that the scaling factors used in the base case are unique within the tested
intervals, where the tested intervals varied by several orders of magnitude above and below
the values shown in Figure 3-3.
5.5

0.20

0.10

4.5

0.00

4
0

0.2

0.4

0.6 0.8
Xreact

0.30

5.5

0.20

0.10

4.5

pH

Tot concentration (M), Fract.Fe

Mg2+
Fe2+
SO42Fract.Fe
pH

pH

Tot. Concentration (M), Fract.Fe

0.30

0.00

A)

B)

Figure 3-3. Effect of variation in a) Xreact and b) , separately, within the base case. In each
case, one factor is varied and the other factor has been held constant at the base case value,
namely Xreact=0.4 and =2.5. Fract.Fe indicates the contribution of pyrite oxidation by
ferric iron to the overall pyrite oxidation rate; compare with Figure 3-4, and see Sections 2.2.3
and 5.3.2.

29

For increase in by three orders of magnitude (from 1 to 1000), concentrations similar to the
4base case could be obtained (Figure 3-4). However, with increasing from 1 to 1000, pH
decreased by over half a pH unit and was then less than the minimum value found in the field.
Noteworthy is that at =1000, the rate of oxidation of pyrite by ferric iron is 30% of the
overall pyrite oxidation rate, compared to 5% when =1, as in the base case (see Section 5.3.2
for discussion of mechanism predominance results). For = 10000, that is, when the rate of
ferrous iron oxidation was increased by a factor of more than 104, concentrations greatly
increased above field conditions, pH dropped to 3, and oxidation was dominated by the ferric
iron mechanism.
Possible conclusions from this include that if conditions changed to favour acceleration of
ferrous iron oxidation in Impoundment 1, for example, due to biological mediation, the effect
could be of a magnitude of up to 1000 and still not affect concentrations of dissolved ions.
This would, however, lead to a slightly lower pH than the field value. Additionally, catalysis
increasing ferrous iron oxidation by more than 2 orders of magnitude would increase the
significance of the oxidation of pyrite by ferric iron compared to the base case; the
mechanisms are equally important when =2000. In the base case, pyrite oxidation by oxygen
dominated at all times (see Section 5.3.2).
5

2.0
Mg2+
Fe2+
SO42+
Fract.Fe
pH

1.5

pH

Tot. concentration (M), Fract.Fe

2.5

1.0
3
0.5

0.0
1

10

100

1000

2
10000

Figure 3-4. Effect of accelerating ferrous iron oxidation by increasing the scaling factor in
the base case. The base case value of =is 1. Fract.Fe indicates the contribution of pyrite
oxidation by ferric iron to the overall pyrite oxidation rate; compare with Figure 3-3, and see
Sections 2.2.3 and 5.3.2.

3.3.2 Calibration results


Calibration was performed by iteratively altering Xreact and . The values arrived at for the
scaling factors are given in Table 3-5.
Table 3-5: Scaling factors used in calibration of the model.
Factor
Description
Xreact
Scale dependency
Biological catalysis of oxidation of sulphides

Biological catalysis of Fe(II) oxidation to Fe(III)

30

Base case value


0.4
2.5
1

As mentioned above, the model gave results similar to field values for pH, SO42-, Cu2+ and K+
even before calibration. Calibration improved Fe(tot) and Mg2+ concentrations and took SO42concentrations and pH closer to average field values (Table 3-4).
Copper concentrations in the model agree reasonably well with the field concentrations,
suggesting that the assumed analogy between rate expressions of pyrite and chalcopyrite
oxidation, in the absence of other expressions for chalcopyrite weathering, may be suitable.
However, this would require further investigation; field Cu concentrations may also be
affected by other processes not accounted for in the model, such as adsorption on surfaces or
precipitation of secondary minerals. Zinc concentrations, on the other hand, were 2-3 orders
of magnitude lower than field concentrations, suggesting that the assumed analogy between
sphalerite and pyrite oxidation rate expressions is incorrect. One possible explanation for the
underestimation could be a galvanic effect, where contact between different sulphides can
accelerate the weathering in the sulphide with the lower potential in the electromotive series
(e.g. Kwong, 1995). This was not accounted for in the model as no expressions quantifying
this effect were found.
Before calibration, Al concentrations were an order of magnitude higher than the average
field value, and the solution was oversaturated with respect to gibbsite; after calibration
concentrations were still higher by a factor of four, though the solution was undersaturated
with respect to gibbsite. Note that field concentrations were close to solubility equilibrium
with Al(OH)3(c), but undersaturated with respect to amorphous aluminium hydroxide. The
representation of Mg-bearing silicate entirely by chlorite may cause the high dissolved Al
concentrations, as talc, which is present in the field but not accounted for in the model,
contains Mg2+ but no Al (see Section 3.2.1.4).
Note that the sulphide weathering rates are not actually affected by the calibration, as they
were scaled by the product Xreact x = 0.4 x 2.5 = 1. Silicate weathering reactions, on the
other hand, were scaled by Xreact only, 0.4. Note also that this was the calibration required to
obtain Mg2+model = ( Mg2+ + Ca2+)field, which was allowed as a conservative estimate of the
probable acidity attenuation caused by the same process that releases calcium. The final Mg2+
concentration was still within the range of concentrations measured in the field. The
calibration required to obtain Mg2+model = Mg2+field was Xreact = 0.25, = 4. The resulting
calibration was still unity for sulphide weathering rates, but the resulting pH was lower, 4.44.
This suggests that there is an additional process in the field besides chlorite dissolution that
has a pH-buffering effect, possibly the same process that releases calcium; however, due to
uncertainty concerning the calcium source as a result of deficiencies in the field data, this
process could not be included in the model. The source of magnesium itself was also
uncertain, as discussed in Section 3.2.1.3.
By varying parameters in the pyrite oxidation expressions over the uncertainties given in the
source literature (see Appendix A), Fe(II) and SO42- concentrations and pH could be varied
over almost the entire range of field conditions, though Mg2+ and Al concentrations were still
too high. Investigation of similar uncertainty in the silicate rate expression parameters may
show that, when compared to the variation in the field and laboratory measurements, the
model does not require the calibration used here at all.

31

4 Variations on the base case


The conceptual model was constructed on the basis of analysis of the available field data. As
the data was limited in relation to the complexity of the system being modelled, it was
necessary to test variations on the base case in order to constrain the conceptual model. For
example, the source of calcium in the field was unclear as porewaters were found to not be at
equilibrium with calcite, and gypsum was not confirmed as present; however, preliminary
geochemical modelling suggested gypsum may have been at solubility equilibrium with the
aqueous phase. As sulphate fluxes have been used as a measure of sulphide weathering, it was
necessary to test the plausibility of the different possible sources of calcium, and in
association with this, the likelihood that equilibrium solubility of gypsum would significantly
affect sulphate fluxes.
All variations on the base case were tested with the same parameters and calibration as used
in the base case, unless otherwise mentioned. A summary of the variations and the
conclusions reached is given in Table 4.1.
The effects of decreased pH, by whatever means, include decreased pyrite oxidation, due to
the inverse dependency of the rate of pyrite oxidation by oxygen on proton concentration (see
Equation 2-10). This leads to decreased concentrations of Fe(II) and SO42-, as well as
decreased release of protons. This also leads to decreased oxidation of Fe(II) to Fe(III), as the
rate of this reaction depends upon pH and the concentration of Fe(II). Decreased Fe(III)
concentrations resulting from decreased ferrous iron oxidation can be offset by increased
solubility of amorphous Fe(OH)3(s), if this mineral is present. Drop in pH also leads to
increased silicate dissolution (see Figure 2-8), and hence increase in base cation production.
These changes in concentrations affect in turn the rate equations, which in turn affect pH;
model results are given when the system reaches steady state.

4.1 Absence of ferric hydroxide


The possibility of the presence of ferric hydroxide, Fe(OH)3(am), was supported by the initial
geochemical modelling within a reasonable range of pe values (see Section 3.2.1.2), and field
reports of the characteristic colour of iron hydroxides (Qvarfort, 1983). However, field redox
conditions were unknown, and proximity to thermodynamic equilibrium does not confirm the
presence of the mineral. The base case was tested without solubility equilibrium with ferric
hydroxide, to see what effect this assumption had on the results. Most of the variations on the
base case were tested both with and without the ferric hydroxide equilibrium.
Without ferric hydroxide present, ferric iron was undersaturated with respect to ferric
hydroxide at base case conditions for Po2>0.02 atm, that is, even in the base case (Po2=0.1
atm). This, in addition to base case results such as the iron and proton balances given in
Figure 5-1, suggested that in the base case, ferric hydroxide is dissolving to give equilibrium
aqueous concentrations of Fe(III). Ferric iron is supplied and the pH is buffered at a higher
level than when the equilibrium is not present (see Table 4-2). The pH without ferric
hydroxide was 3.80, compared to 4.53 with the mineral present and 4.87 in the field, with the
associated effects of lower pH as discussed in Section 4.1 above; this suggests that either
Fe(OH)3(am) dissolution or some other pH-buffering process is present in the field. In the case
without Fe(OH)3(am) present and acting as a source of Fe(III) ions, Cu2+ and Zn2+
concentrations were much lower, as the rate of oxidation of these sulphides by ferric iron in
the presence of ferric hydroxide is of the same magnitude or greater than the rate of oxidation
by oxygen, as shown later in Table 5-2.

32

The assumption of equilibrium with Fe(OH)3(am) in the base case implies that conditions were
suitable for Fe(OH)3(am) precipitation during some earlier period of time, and have in the base
case changed to favouring dissolution. The comparison of the models with and without ferric
hydroxide may be interpreted as successive steady states. Originally the ferric hydroxide was
present, that is, had precipitated earlier, and was dissolving at steady state; eventually the
mineral is depleted and a new steady state is reached in the absence of ferric hydroxide, with
lower pH and slightly different porewater composition (see Table 4-2).
Table 4-2.Resulting total concentrations (M) for variations on the base case.
No equilibrium
Eqm with
Field monthly
Base case:
with Fe(OH)3(am) gibbsite and
average*
Equilibrium
Fe(OH) 3(am)
with Fe(OH)3(am)

Eqm with
gibbsite,
no
Fe(OH) 3(am)
4.42
0.09
0.05
3 x 10-5
6 x 10-6
0.017
2 x 10-4
0.02

pH
4.87
4.53
3.80
4.54
SO420.1
0.1
0.08
0.1
Fe(tot)
0.08
0.07
0.04
0.07
Cu2+
6 x 10-5
6 x 10-5
4 x 10-5
Zn2+
0.006
5 x 10-5
6 x 10-6
5 x 10-5
2+

Ca
0.005
-Mg2+
0.011
0.017
0.02
0.017
Mg2++Ca2+
0.016
+
K
6 x 10-5
2 x 10-4
2 x 10-4
2 x 10-4
Al
0.002
0.008
0.01
0.009
*
See also Table 3-4 for the range of field values.

As no minerals containing Ca2+ were included in the model, we allowed


Mg2+model=( Mg2+ + Ca2+)field to allow for the probable buffering capacity of minerals that
release Ca2+in the field (see Section 3.2.1.3).

Dissolution and precipitation of Fe(OH)3(am) is sensitive to pH, as was shown in


Equation 2-35; for example, in the uncalibrated model results (Table 3-4) with higher pH and
lower total iron concentrations, this mineral is precipitating.

4.2 Equilibrium with gibbsite


According to the preliminary speciation modelling with field data, gibbsite (Al(OH)3(c)) and
bohemite (AlOOH(s)) were close to equilibrium, though Al(OH)3(am) was slightly
undersaturated. Equilibrium with gibbsite was thus tested in the model, with and without
ferric hydroxide present; over-/undersaturation of Al and Fe(III) was checked by comparison
with the relevant case without the presence of the appropriate secondary mineral. Blowes and
Ptacek (1994) suggested that the dissolution of Al(OH)3(s) after the consumption of carbonates
buffers AMD solution pH between 4.0 and 4.3, and that Al(OH)3(s) dissolution occurs before
ferric hydroxide begins to dissolve. However, it has also been suggested that at low pH and
high sulphate and potassium concentrations, gibbsite is unstable to Al-hydroxysulphates,
which may then control aqueous aluminium concentrations in the AMD environments
(Nordstrom and Alpers, 1999). Jurbanite (Al(SO4)(OH)5H2O(s)) and basaluminite
(Al4(SO4)(OH)105H2O(s)) were oversaturated, according to the preliminary geochemical
modelling, suggesting that these minerals were not controlling aqueous concentrations. The
implications of the presence of aluminosulphates have not been investigated here.
As in the case with ferric hydroxide alone, that is, the base case (see Section 4.1), for
Po2>0.02 atm results were undersaturated with respect to aluminium hydroxide when
equilibrium with this mineral was not included. This suggested that in the case with forced
equilibrium the hydroxide was dissolving to provide equilibrium concentrations. This
dissolution consumed protons, buffered pH, and released Al. Equilibrium with gibbsite gave

33

aluminium concentrations that were slightly higher than in the base case, which was already a
factor of four greater than the average field concentration. However, as can be seen in Table
4-2, the resulting concentration of Al is quite sensitive to pH; higher pH in the model due to
some other process may lower Al concentrations, which then agree with the field
concentrations. However, as there was too much uncertainty associated with assuming
equilibrium with gibbsite, this process was not included in the base case.

4.3 CO2 partial pressures


The Pco2 assumed in the base case was the partial pressure in the atmosphere (0.032%) as
there were no measured values reported. Partial pressures of 30% and 0% were also tested. As
expected, the absence of CO2 had only a slight effect on pH, and only in simulations that had
resulted in higher pH, such as those with solubility equilibrium with calcite. Similarly,
increasing Pco2 to 30% had no effect on the base case results, though as mentioned below, in
cases such as those with calcite present the chosen Pco2 could affect model results.

4.4 Equilibrium with gypsum


4.4.1 Gypsum equilibrium
Preliminary geochemical modelling suggested that gypsum and anhydrite were both close to
solubility equilibrium with solution, implying that a secondary calcium sulphate specie may
have affected aqueous calcium and sulphate concentrations in the field; this was tested in the
model.
There was no source of Ca2+ in the base case, as the presence of calcite or dolomite could not
be confirmed or excluded (see Section 4.5), and no non-carbonate calcium-bearing minerals
were given in the mineralogical description of the gangue material. The base case was thus
undersaturated with respect to gypsum; forcing aqueous calcium and sulphate concentrations
to be in equilibrium with gypsum implied that gypsum was present and dissolving. Solution
pH and sulphate concentrations were not significantly affected, and calcium was released at a
rate equal to the higher flux measured in the field, 10-7 mol m-2s-1 (that is, total aqueous
concentrations of 0.01M at an influx rate of 10-8 m.s-1; see Figure 4-1 below). Hence, we
cannot exclude that the source of calcium in the field is dissolution of previously precipitated
gypsum.
Simulations where gypsum was the source of calcium and Mg2+ was calibrated to the field
concentration, rather than (Mg2+ + Ca2+)field as in the base case, resulted in a pH that was too
low, and Al concentrations were still too high, despite decreased silicate weathering. This
does not eliminate gypsum as a possible source of Ca2+, as the difference between field and
model proton balance may lie in the chosen composition of silicate minerals (see Section
3.2.1), as well as uncertainties in parameters and rate constants. However, the results suggest
that the source of calcium in the field is associated with a slight buffering capacity, rather than
gypsum dissolution being the only source.

34

Table 4-1. Summary of and conclusions from variations on the base case.
Variation
Absence of
Fe(OH)3(am) eqm

Results
-Undersaturated w.r.t. Fe(OH)3(am) for Po2>0.02
atm
-pH below field value

Gibbsite eqm

-Without eqm, undersaturated for Po2>0.02 atm


-Buffers pH, e.g. if Fe(OH)3(am) not present
-Absence or 30% partial pressure of CO2 had no
effect on base case
NB: Pco2 did affect cases with calcite and
elevated pH.
-No source of Ca2+ in base case; thus gypsum
eqm implied dissolution
- Ca2+ released at upper rate of flux seen in the
field (10-7mol m-2s-1)
-pH, [SO42-] not significantly affected
-without gypsum eqm:
When flux same as the field (10-8 mol m-2s-1),
system was oversaturated w.r.t. gypsum

CO2 eqm

Gypsum

Ca2+ addition
(as a slow,
kinetic process )

-with gypsum eqm


-Sulphate affected when Ca2+ flux >10-8.
-Flux > 10-6, [SO42-] lower than field
concentrations
-No other component affected until flux >10-1

Presence of
calcite:
-Calcite, siderite,
gypsum eqm*
-Kinetic
dissolution of
calcite and
gypsum eqm*
-Influx of water
in contact with
calcite, and
gypsum eqm#
Oxygen
diffusion:
-Varied Deaq@

-Varied (water
content)
Heterogeneous
Fe(II) oxidation

Cannot achieve field pH, [Ca2+] and reasonable


Pco2 at the same time

If
the volumetric fraction of calcite was
sufficiently small for pH to agree with field,
[Ca2+] within field range; [SO42-] was not
affected by gypsum precipitation
-[SO42-] and pH affected with inlet flux >
10-8 mol m-2s-1 i.e. inlet [Ca2+]=[CO32-]=10-3 M.
-Inlet [Ca2+]=[CO32-]=10-1 M, conditions
outside field range (e.g. pH> 6.15).
-Deaq 5 x 10-6 m2s-1, O2 diffusion limits
-Deaq >5 x 10-6 m2s-1, chemical kinetics limit
-Constant Deaq=5 x 10-6 m2s-1, results sensitive
for all sulphide oxidation rates
Deaq affected, and hence most model results

Conclusion
-pH was too low, suggesting Fe(OH)3(am) was present
and dissolving in the field
-The case with no Fe(OH)3(am) is appropriate for the
case where the mineral has been consumed
Solubility equilibrium with gibbsite was not included,
too uncertain.
Atmospheric Pco2 was assumed in the base case

A sole source of Ca2+ could be dissolution of


previously precipitated gypsum, but could not be
confirmed. Not included in the base case.
If there is another source of Ca2+ than gypsum, gypsum
is probably present and precipitating to achieve
solubility equilibrium.

Field conditions can be reproduced with


10-8 <Ca2+ flux < 10-6, e.g. by weathering of 10%
anorthite. A source of Ca2+ greater than 10-6mol m-2s-1
would affect sulphate concentration and hence may
affect model results.
NB: No buffering was included explicitly in the Ca
influx, but that Mg2+model=( Mg2+ + Ca2+)field allowed
for a slight buffering capacity.
Calcite can not be at solubility equilibrium with
solution in Impoundment 1.

Kinetic dissolution of calcite is one of the possible


sources of Ca2+, if the volumetric fraction of calcite is
less than 10-7.
Influx of water that has been in contact with calcite,
and as a result, [Ca2+]=[CO32-]>10-3 M, will raise pH
and affect concentrations.
Mixed diffusive and reaction kinetic control possible;
should be further investigated in a more comprehensive
model.
Model results are sensitive to degree of water
saturation for >0.15 with porosity = 0.25.
The surface site concentration was assigned
10-9 M in the base case, as it was an uncertain
parameter. More information required.

-Surface site concentration varied


-At the only reported site concentration (10-3 M),
became the dominant Fe(II) oxidation process
-At lower concentrations, small effect on pH,
even less on [Fe(II)]
eqm equilibrium
w.r.t. with respect to
*The system was oversaturated with the

I.e., if the volumetric fraction


mentioned minerals if the equilibrium was not present
-7
of calcite was <10 , i.e. much lower than reported in mineralogy or solid phase analysis #When
the
flux was high enough to affect other components, the system was oversaturated with gypsum if the
@
equilibrium was not present
Deaq diffusion coefficient in the aqueous phase

35

0.14

Field values:
maximum
average
minimum

Fe(OH)3, pH 4.53 (base case)


0.12

Total concentration (M)

gypsum, Fe(OH)3, pH 4.53


0.10
0.08
0.06
0.04
0.02
0.00
Mg2+

Ca2+

SO42-

Fe2+

Figure 4-1. Model results, base case and base case with gypsum equilibrium.

4.4.2 Influx of calcium


Investigation was conducted into the case where calcium was added to both the base case and
the case with equilibrium with gypsum, to see which rate of addition of Ca2+ from an
unspecified source, such as an unreported Ca-bearing mineral, would affect model results.
When Ca2+ was added to the base case system (without gypsum equilibrium) at the same
apparent rate of production as the net rate in the field, that is, 10-8 to 10-7 mol m-2s-1 (total
aqueous concentrations of 0.001M-0.01M at the given infiltration rate), the solution
concentrations approached field conditions and oversaturation with respect to gypsum.
With addition of Ca2+ to the system with gypsum equilibrium present, sulphate concentrations
were affected when the rate of addition was greater than 10-8 mol m-2s-1, and decreased below
the minimum field sulphate concentrations when the influx increased above 10-6 mol m-2s-1.
No other component was affected until pH began to decrease, which occurred when the influx
of Ca2+ was greater than 10-1 mol m-2s-1. This suggested that unless gypsum is prevented from
precipitating, precipitation will begin to occur when the influx of calcium is greater than
10-8 mol m-2s-1; additionally, an influx in the field of less than 10-6 mol m-2s-1 does not
drastically affect model results. It should be noted that calcium release by processes other
than gypsum dissolution may also affect the acidity and alkalinity of the solution, such as has
been allowed for by Mg2+model=( Mg2+ + Ca2+)field, and hence affect model results.
Calcite cannot be confirmed as a source of Ca2+, in the presence of gypsum or otherwise, as
discussed in the section below. A possible source of Ca2+ may be an unreported silicate, for
example, Figure 2-8 shows the rate of dissolution of the calcium-bearing mineral anorthite. At
the field pH and temperature, the figure shows that the surface area normalised rate of release
from this source would be approximately 3 x 10-13 mol m-2s-1. Estimates showed that the
weathering of anorthite, if 10 vol-% were present, would correspond to the critical Ca2+ source
strength of 10-8 mol m-2s-1 (area normalised flux) mentioned above. Ca-bearing gangue
minerals actinolite and apatite were reported to be common in the Kristineberg mine (du
Rietz, 1953), and as the tailings come from other mines in the region, other Ca-bearing
minerals may be present.

36

As mentioned above, the source of calcium in the field is unknown and could be the
dissolution of gypsum that had precipitated at an earlier point in time, and/or the dissolution
of calcite or an unknown silicate mineral. As the presence of gypsum equilibrium under
conditions of Ca2+ input at the same rate as field production does not seem to affect other
results than calcium concentrations and it is unknown whether gypsum is actually present, the
mineral has been excluded from the base case. However, this mineral should be considered in
future model developments, and field characterisation should involve determination of the Ca
source, as this may be crucial to the proton balance.

4.5 Presence of calcite


4.5.1

Assumed equilibrium with calcite

The over-/undersaturation of calcite in the field could not be tested in the preliminary
geochemical modelling, as no field carbonate analyses were available. Calculations indicated
that if solubility equilibrium with calcite were present it would be accompanied by
unreasonably high partial pressures of carbon dioxide, if field calcium concentrations and pH
were to be maintained. To further support this, equilibrium with calcite was tested in the
model, with varying Pco2. Calcite equilibrium resulted in oversaturation of solution with
respect to Fe(OH)3(am) and gypsum, equilibrium with these minerals were assumed. Solution
was also oversaturated with respect to siderite (FeCO3(s)); equilibrium with this mineral was
tested as well.
Results obtained suggested that solubility equilibrium with calcite is not the case in the field,
as aqueous calcium concentrations were too high and sulphate concentrations too low in all
cases (see Figure 4-2); pH within the field range was only obtained in the case with siderite
and elevated Pco2, while concentrations were still lower than field conditions. These results
suggest that calcite is not at solubility equilibrium in the field.

0.14
calcite, gypsum, Fe(OH)3 pH 7.50

Total concentration (M)

0.12
0.10
0.08

Field values:
maximum
average
minimum

calcite, gypsum, Fe(OH)3,


30%Pco2, pH 6.12
calcite, siderite, gypsum, Fe(OH)3,
pH 7.94
calcite, siderite, gypsum, Fe(OH)3,
30% Pco2, pH 5.99

0.06
0.04
0.02
0.00
Mg2+

Ca2+

SO42-

Fe2+

Figure 4-2. Model results with calcite equilibrium, gypsum and calcite equilibrium, and
gypsum and calcite equilibrium at elevated Pco2.

37

4.5.2

Kinetic dissolution of calcite

If calcite was present and dissolving, the results from the previous section suggested that the
dissolution was in some way kinetically limited from reaching thermodynamic equilibrium. In
order to test if calcite was present and dissolving slowly, a rate expression for the kinetic
dissolution of calcite was added to the model in the place of equilibrium (see Section 2.4.1
and Appendix E). The rate of the reverse reaction, that is, precipitation, was not included, as
this is reported to be insignificant below pH 8 (Chou et al., 1989).
The original volumetric fraction of carbonate in the gangue material in the tailings was
reported to be 10 volume-%, that is, 7.5% of the total volume (see Table 3-1). However, only
with a volumetric calcite content 5 orders of magnitude less than this, ca=10-7, did pH and
sulphate and calcium concentrations approach average field conditions (see Figure 4-3).
base case, pH 4.53
Field values:
maximum
average
minimum

0.14

Total concentration (M)

0.12

kinetic calcite dissolution,


gypsum eqm, vol. fract.10^-7,
pH 4.73

0.10
0.08
0.06
0.04
0.02
0.00
Mg2+

Ca2+

SO42-

Fe2+

Figure 4-3. Model results with kinetic calcite dissolution. The volume fraction of calcite, ca,
is 10-7.

As shown in Figure 4-3 , a sufficiently low volumetric fraction of calcite results in pH and
Ca2+ concentrations within the range of field conditions. The solution is oversaturated with
respect to gypsum, but the addition of equilibrium with this mineral did not greatly affect
sulphate concentrations. Model results could also be calibrated to field Mg2+ concentrations,
resulting in almost the same alternative calibration as for the base case (Section 3.3.2). The
resulting pH was the same as in the base case; calcium concentrations were a factor of two
greater than field conditions, but still within the range of field measurements. Thus the slow
dissolution of a small fraction of calcite may be the source of Ca2+ in the field, or
alternatively, dissolution of calcite with limited availability (see e.g. Strmberg, 1997).
However, this could not be determined from the available field data.
4.5.3

Influx of a water in contact with calcite

As spreading of lime on the surface of tailings impoundments is a common method of


attempting to neutralise porewaters, cases where the influx of water had been in contact with
calcite were tested. This was performed by adding varying total concentrations of calcium and
carbonate in a 1:1 ratio to the inlet water in the base case, the case without ferric hydroxide
equilibrium, and the case with gypsum equilibrium; this in order to determine which level of

38

influx was required to affect model results. Atmospheric carbon dioxide partial pressures
were assumed.
In the cases without gypsum present, calcium and carbonate concentrations in the incoming
water of greater than 10-2 M each were required to raise the pH by 0.1 units, that is, a flux of
10-7 mol m-2s-1, and higher concentrations had greater effect. At this flux of 10-7 mol m-2s-1, the
solution became oversaturated with respect to gypsum. In the case with gypsum equilibrium
assumed, sulphate concentrations and pH began to be affected already when inlet
[Ca2+]=[CO32-]=10-3 M, that is, with fluxes of 10-8 mol m-2s-1. At an inlet concentration of
10-1 M, that is, a flux of 10-6 mol m-2s-1, pH was greater than the highest field value (pH 6.53
compared with a maximum of 6.15 in the field; see Figure 4-4).

0.30

Total concentration (M)

base case, pH 4.53


0.25
0.20

gyps eqm, influx of 10^-6 of


Ca, CO32-; pH 6.53

Field values:
maximum
average
minimum

0.15
0.10
0.05
0.00
Mg2+

Fe2+

Ca2+

SO42-

Figure 4-4. Model results with influx of 1:1 of total calcium and carbonate concentrations.

Investigations were performed with a geochemical equilibrium model to see whether


concentrations in incoming waters that were in contact with calcite could be high enough to
disturb model results, that is, greater than 10-3 M. Calcite dissolution leads to the release of
equal total moles of Ca2+ and CO32-. In the case of equilibrium between the solid calcite and
pure rainwater, the maximum possible total concentrations will vary with the Pco2 in solution
as shown in Figure 4-5. That is, with elevated Pco2 it is possible to have Ca2+ and CO32concentrations at the given infiltration rate that can disturb the model geochemistry.
In summary, model results indicate that influx of water in contact with dissolving calcite can
raise pH and affect concentrations in the impoundment.

39

[H+ ] T O T =

0.10 M

Log {CaCO 3 (c)} =

0.00

-1

CO2 (g)

Log Solubl.

CaCO 3 (c)
-3

-5

-7

-9
-5

-4

-3
Log PCO

-2

-1

Figure 4-5. Total aqueous concentrations of components CaCO3(c) and CO2(g) resulting from
equilibrium of pure water with calcite, as a function of partial pressure of carbon dioxide (at
25C).

4.6 Oxygen transport limitation on sulphide oxidation.


Due to slow oxygen transport in mine tailings, an alternative control on sulphide oxidation,
besides (bio)chemical kinetics (See Section 2.2), is the availability of oxygen. The oxygen is
transported by advection and diffusion, where in many cases the latter process is more
important.
The box-model we use accounts for advective transport of solutes to and from the box using
Equations 3-3. Diffusive transport of oxygen follows Ficks law:
f O2 =

De C
z

(4-1)

where fO2 is the area-normalised flux of oxygen, De is the effective diffusion coefficient, C is
the oxygen concentration gradient, and z is the diffusion length. Simplistically, this would for
our box model of the unsaturated zone correspond to:
f O2 =

Deaq (Cin C )
l/2

(4-2a)

where Cin is the aqueous oxygen concentration at the top of the box, C is the steady-state
oxygen concentration in the box and Deaq is the effective diffusion coefficient for aqueous
concentration gradients. The diffusion length is assumed to be half of the height, l, of the box.
Note that the STEADYQL model is not designed to handle concentration gradients, and,
hence, that our assessment is to be seen as tentative.
Because the STEADYQL-code can not handle kinetic expressions that include differences
such as (Cin-C), we formulated Equation 4-2a as:
f O2 = f + f

(4-2b)

40

where

f+ =

Deaq Cin
and f
l/2

Deaq C
. The processes f+ and f- were then introduced in the
l/2

model as two independent processes. This is numerically unfortunate in cases where f+ and fboth are large and the resulting fO2 is small, as the accuracy of the model results is limited. An
additional shortcoming of this way of representing diffusion processes is that we lack the
possibility to account for diffusion of oxygen to deeper parts of the impoundment, that is,
those underlying the box being modelled. This process, however, is only important when the
steady state concentration of oxygen in the box is large.
The effective diffusion coefficient is dependent upon, among other things, water saturation.
Empirical expressions relating these quantities have been reported by, for example, Reardon
and Moddle (1985) and Elberling and Nicholson (1996):
3
0.05
De = 3.98 10
T 2
0.95

1.7

De = D0 (1 S )

SDw
H

(4-3)

(4-4)

where DO is the diffusivity of oxygen in air (DO=1.8 x 10-5 m2s-1), Dw is the diffusivity in
water (Dw=2.1 x 10-9 m2s-1), H =

1
where KH is the Henrys law solubility of oxygen in
RTK H

water (KH=10-2.68 M atm-1), is the water content, is the porosity, and S =

. and are

empirical constants (=3.28 and =0.273; Elberling and Nicholson, 1996). The aqueous
phase-based diffusion coefficient, Deaq, is related to the gas phase-based effective diffusion
coefficient, De, by:
Deaq =

De
K H RT

(4-5).

Figure 4-6 compares the effective diffusion coefficients as given by Equations 4-3 and 4-4
for a porosity of =0.25.

Figure 4-6. The effective diffusion coefficient as given by Equation 4-4 (Elberling and
Nicholson, 1996) and Equation 4-3 (Reardon and Moddle, 1985) as a function of at =0.25.

41

4.6.1 Varying the effective diffusion coefficient


Using the base case, we have tested the sensitivity of model results to diffusion of oxygen by
removing the constraint of Po2=0.1 atm and introducing oxygen to the box through advection
(Equation 3-3) as well as diffusion (Equation 4-2). With Cin=4.0 x 10-4 M, which corresponds
to Henrys law solubility of oxygen in water at a partial pressure of oxygen equal to that in the
atmosphere, Deaq was varied; the response in total sulphate concentrations, [SO42-]T, [O2(aq)]
and pH at steady state is shown in Figure 4-7. For Deaq 5 10 6 m2s-1, which correspnds to
the = in the base case, the model results show decreasing sulphate and oxygen concentrations
and increasing pH with decreasing Deaq, indicating that oxygen diffusion is limiting the
weathering processes. However, at higher effective diffusion coefficients, the concentrations
are insensitive to changes in Deaq, indicating that chemical kinetics is limiting the reactions.

Figure 4-7. Model results for various effective oxygen diffusion coefficients at steady state.
The left axis shows the free aqueous oxygen concentration and the total sulphate
concentration; the right axis shows pH.

In the same way, we varied the sulphide oxidation rate (by changing ) with
Deaq=5.0 x 10-6 m2s-1. Model results, see Figure 4-8, are sensitive to over the entire range of
values studied, 0.4<<100 (note that the base case value was 2.5). This indicates that the
processes are at least partly controlled by chemical kinetics.
From this comparative sensitivity analysis we conclude that the control on the rate of sulphide
oxidation may be mixed, and that both chemical kinetics and oxygen diffusion must be
assessed in a more comprehensive model.

42

Figure 4-8. Model results with Deaq=5.0 x 10-6 m2s-1 and varied . The axis to the left shows
the free oxygen concentration and the total sulphate concentration whereas the axis to the
right shows pH at steady state.

4.6.2 Varying water content


Varying the water content and using the corresponding De (from Equation 4-4), showed, as
expected, that model results depend drastically on for >0.15 (see Table 4-3).
Table 4-3. Model results at various water contents in the case of diffusive transport of
oxygen. Other parameters as in the base case. Concentrations are given in units of M.
=0.10
=0.15
=0.19
=0.20
pH
4.38
4.42
4.93
6.54
[SO42-]T
0.13
0.11
0.051
0.031
[Fe(II)]T
0.097
0.080
0.025
0.014
[Cu2+]T
9x 10-5
7 x 10-5
2 x 10-5
6 x 10-6
[Zn2+]T
9 x 10-5
8 x 10-5
1 x 10-6
6 x 10-7
2+
[Mg ]T
0.017
0.017
0.015
0.013
[O2(aq)]free
3.5 x 10-4
2.6 x 10-4
4.7 x 10-5
7.8 x 10-6

4.7 Heterogeneous ferrous iron oxidation


In Section 2.3.2, heterogeneous, surface mediated ferrous iron oxidation is described. At
higher surface oxidation site densities such as reported for lepidocrite in the literature
(3 x 10-6 mol m-2; Zhang et al., 1992; converted to mol l-1 with the specific surface area and ),
this process was the dominant mechanism for Fe(II) oxidation. No reaction activation energy
was available to account for field temperature, and the effect of ionic strength on surface
reactions is unknown.
When the surface site concentration was lowered by 6 orders of magnitude, pH increased by
only 0.21 units and total iron concentrations increased slightly. Other concentrations were not
significantly affected, and model results did not vary with further decrease in surface site
concentration.
Due to the uncertainty in the parameters involved, the lower surface site concentration was
utilised in the base case.

43

5 Base case results and modelling of remediation


5.1 The base case: Modelling of Impoundment 1
The aim of the modelling of Impoundment 1 was to reproduce the main ions reported in field
geochemistry, namely SO42-, Fe(tot), Mg2+, Ca2+ and Zn2+, and pH, as closely as possible,
given the field data available.
The processes involved in the base case were shown schematically in Figure 3-1. The
uncalibrated model gave results that were within the given range of field conditions for pH,
SO42-, Fe(tot), Cu2+ and K+. Magnesium concentrations were within a factor of two and Al
concentrations were within an order of magnitude of average field conditions (compare Table
3-4). Calibration to field conditions involved the use of two scaling factors (see Section
3.3.1); one allowing for the scale dependence of the laboratory-derived rate expressions that
were used in the model (Xreact=0.4), and a second allowing for biological mediation of
sulphide oxidation reactions (=2.5). A third calibration factor was considered, allowing for
the biological mediation of ferrous iron oxidation (), but was not required to obtain results
close to field conditions.
It should be noted that the calibration factors were small, and for the sulphide oxidation
reactions, cancelled each other out (0.4 x 2.5 = 1). Xreact was only a small correction, most
likely because we had already accounted for some of the factors that have been suggested to
contribute to the scale dependence of weathering reactions. Compilation of rates from the
literature indicate that the direct biological catalysis of sulphide reactions can be in the order
of a factor of 10 (e.g. see review in Herbert, 1999). At the low average temperature in the
field (1C), a factor of 2.5 may be realistic. However, as mentioned in Section 3.3, it may be
possible to account for the calibration factors by considering the reported uncertainty
associated with parameters and constants, in which case the geochemistry can be explained by
abiotic processes.
Calibration brought SO42-, Fe(tot) and Mg2+ concentrations and pH closer to average field
values, while not adversely affecting Cu2+ or K+ concentrations. Zn2+ concentrations remained
more than two orders of magnitude too low, both before and after calibration; this was
probably due to the representation of sphalerite oxidation, which for oxidation by oxygen was
taken from analogy with pyrite, as a rate law was not available in the literature. This analogy
appears to be inappropriate. Sphalerite dissolution may be affected by galvanic effects
between sulphide minerals (see Section 2.1.2).

5.1.1 Oxygen consumption


Aqueous oxygen concentrations were assumed to be at equilibrium with a constant oxygen
partial pressure of 10% in the gas phase. By summating the flux of oxygen consumed by the
relevant slow kinetic reactions and that lost in the outflow, the overall area normalised oxygen
consumption rate was calculated to be 1.29 x 10-6 mol m-2s-1 (see Table 5-1). If we consider
the distribution of oxygen between the oxygen-consuming processes, pyrite oxidation is
clearly dominant.
The total rate of oxygen consumption is equivalent to approximately 40 mol m-2year-1. This
agrees with the field measurements of Elberling and Nicholson (1996) at Falconbridge,
Ontario, Canada, which ranged from 0.1-250 mol m-2year-1.

44

Table 5-1. Rates of oxygen consumption for various processes in the base case.
Rate of O2
Process
consumption
(mol m-2s-1)
Fe(II) oxidation
8.87 x 10-10
Pyrite oxidation by O2
1.29 x 10-6
Chalcopyrite oxidation by O2
9.65 x 10-10
Sphalerite oxidation by O2
4.94 x 10-11
Outflow
1.57 x 10-9
Total consumption of O2
1.29 x 10-6

5.1.2

Iron

Dominant forms of iron in solution in the model were ferrous iron species Fe2+ and FeSO4(aq).
From Figure 5-1a it can be seen that the major source of ferrous iron is pyrite oxidation; also
that oxidation of ferrous iron is relatively slow. The figure also shows that chlorite
(chlinochlore) weathering does not contribute significantly to the Fe(II) mass-balance. Pyrite
oxidation through the ferric iron path is, although unimportant for the overall rate of pyrite
oxidation and the sulphate mass-balance (as shown in Equation 2-9), an important source of
Fe(II). This fact may help explain the low SO42- to Fe(tot) ratio found in the aqueous phase in
the impoundment (see Section 3.2.1). The resulting redox potential for base case at Po2=0.1
atm was 7.6.
Figures 5-1b and c show that the oxidation of pyrite by ferric iron consumes ferric iron and
produces protons, as indicated in Equation 2.9. However, when the source of the ferric iron is
dissolution of a solid ferric hydroxide, the overall reaction consumes protons
FeS 2 ( s ) + 14 Fe(OH ) 3( s ) + 26 H + 15Fe 2+ + 2SO42 + 34 H 2 O
(5-1)
The amount of Fe(OH)3(am) dissolving in the base case was calculated to be 0.20g kg tailings-1
year-1. In the absence of Fe(OH)3(am), oxidation of pyrite by ferric iron had relatively little
effect, on either ferrous iron concentrations or the proton balance (see Figures 5-1b and c).
Without Fe(OH)3(am), pe was lower, 5.6; ferrous iron concentrations differed only by a factor
of 2 from the base case, but the free ferric iron concentration was lower by up to three orders
of magnitude. Comparison with the base case indicated that if the ferric hydroxide is
consumed in the impoundment, pH and pe will drop, as will all concentrations other than
those associated with silicate weathering. Conditions of lower pH may assist the action of
acidophilic bacteria; however, among other factors controlling microbial mediation of
reactions, field temperature may be too low to allow significant activity (see Figure 2-3).

5.1.3

Proton balance

Figure 5-1c shows that the only process releasing large quantities of protons is pyrite
weathering by oxygen. Processes buffering the released acidity are mainly chlorite
weathering, and in the base case, dissolution of the ferric hydroxide. In the presence of a
process releasing Ca2+ and consuming protons (e.g. calcite and/or Ca-silicate dissolution; not
represented in the base case model), the proton balance may be affected to a small extent (not
shown). Figure 5.1c also suggests that only a small portion of the protons that are released
through pyrite weathering leaves the unsaturated zone. However, the high ferrous iron
concentration and flux implies potential additional acidity; protons are produced via oxidation
of ferrous iron and associated precipitation of ferric iron, subsequent to mixing of AMD with
oxic surface waters.

45

base case
1

10

10

10

10

11

11

11

11

-6.E-07

0.E+00

Fe(II)

6.E-07

A)

-2.E-07

0.E+00

Fe(III)

2.E-07 -1.E-06

0.E+00
+

B)

1.E-06 -1.E-06

C)

no Fe(OH)3

0.E+00

SO4

2-

1.E-06
C)
D)

Figure 5-1. Fluxes (mol m-2s-1) of A) Fe(II), B) Fe(III), C) H+ and D) SO42- due to various
processes* for the base case, pH=4.53, and the case with no Fe(OH)3(am) present, pH=3.80.
*Processes:
1) Ferrous iron oxidation
7) Sphalerite oxidation by ferric iron
2) Pyrite oxidation by oxygen
8) Dissolution of chlinochlore
3) Pyrite oxidation by ferric iron
9) Dissolution of muscovite
4) Chalcopyrite oxidation by oxygen
10) Flux out of the unsaturated zone
5) Chalcopyrite oxidation by ferric iron
11) Dissolution of Fe(OH)3(am) in the base
6) Sphalerite oxidation by oxygen
case

5.1.4

Sulphide oxidation mechanism dominance

Oxidation of metal sulphides occurs with either oxygen or ferric iron as oxidant. As shown in
Table 5-2, oxidation by oxygen dominates for pyrite, the two mechanisms have equal
importance for chalcopyrite, and oxidation by ferric iron dominates for sphalerite. Note that
expressions for chalcopyrite and sphalerite oxidation by oxygen were taken from analogy
with pyrite oxidation, as none were found in the literature; see Appendix D for the
expressions used. Zinc concentrations in the base case were two orders of magnitude lower
than field concentrations, suggesting the analogy is incorrect for sphalerite, or that other
factors may be affecting Zn concentrations.
It can also be seen that oxidation of pyrite is the most significant sulphate-producing process
in the current model.
Table 5-2. Relative importance of oxidation of sulphides by oxygen and ferric iron, as
indicated by the sulphate flux for each process.
Log
Flux due to
Sulphate flux
Mineral
Formula
Oxidant
rate
O2/total flux
(mol m-2s-1)
O2
-6.43
3.7 x 10-7
pyrite
FeS2(s)
0.95
-8
Fe(III)
1.1 x 10
-7.97
O2
2.4 x 10-10
-9.62
chalcopyrite CuFeS2(s)
0.53
Fe(III)
2.2 x 10-10
-9.66
-11
O2
2.5 x 10
-10.61
sphalerite
ZnS(s)
0.06
Fe(III)
3.6 x 10-10
-9.45

46

5.2 Investigation of the effect of remediation


The effect of variation of input parameters in the base case was investigated. The aim was to
investigate how the system would behave under changed conditions, for example due to
remediation.
The aim of remediation is often to decrease the rate of sulphide oxidation by hindering the
diffusion of oxygen into the impoundment. Two common methods are dry covering, where
the goal is to decrease infiltration of air and water in to the impoundment, and wet covering,
where saturating the impoundment reduces oxygen diffusion. The three relevant parameters,
namely the partial pressure of oxygen, the infiltration rate and the water content, were varied
separately in the model; in reality, the infiltration rate and water content will affect the
availability of oxygen, as indicated by tests with oxygen diffusion, see Section 4.6.
Another common method of treatment of impoundments is the addition of lime, either as the
sole remediation measure or in combination with other methods of remediation. The presence
of calcite at equilibrium, with kinetically limited dissolution, or in contact with water entering
the impoundment, was discussed in Section 4.5. The effect of decreasing partial pressure of
oxygen on the case with equilibrium with calcite was tested as the extreme case of calcite
dissolution, as part of the model investigations which are presented in the following sections.
The effect of varying pyrite fraction (py) was also tested, as this is generally accepted to have
a significant effect on tailings impoundment geochemistry, both in terms of the total AMD
generation capacity of an impoundment, and the effect on AMD generation as the sulphides
are weathered in the upper layers of the impoundment. Consecutive simulations with
decreasing sulphide concentrations can be used to model the consumption of sulphide in the
tailings. Remediation techniques can also involve decreasing the sulphide content, or effective
sulphide content, of the tailings, for example, by pre-disposal treatment or by surface-coating
additives (Evangelou, 1995).
The partial pressure of oxygen (Po2), infiltration rate (q), water content () and volumetric
fraction of sulphides were individually varied as given in Table 5-3.
Table 5-3. Parameters varied during simulations.
Parameter
Description
Base case value
Varied over range:

0.1
Po2
Set partial pressure of O2 in the
1 x 10-4 0.2
gas phase in equilibrium with the
aqueous phase (atm)
Degree of water saturation
0.15
0.1-0.25&

-1
-8 &
q
Rate of water infiltration (m s )
7.7 x 10
7.7 x 10-7- 7.7 x 10-10
Volumetric percent of pyrite in
4%*
1-20% #
py
the tailings
Estimated value
& Axelsson et al., 1986
*Qvarfort, 1983
#spalerite and chalcopyrite fractions varied proportionally with py

Of particular interest is which pyrite oxidation mechanism dominates under which conditions.
A hypothesis we tested was that when the partial pressure of oxygen in the impoundment is
lowered due to remediation, oxidation of sulphides by ferric iron becomes dominant over
oxidation by oxygen. If this is the case, remediation via the decrease of oxygen flux into the
impoundment may have less effect than expected.

47

5.3 Results of investigations


As indicated in Table 5-3 above, four parameters were individually varied to investigate the
effect of remediation in the base case. Decreasing Po2 was also investigated in the case of no
equilibrium with Fe(OH)3(am), and the case with equilibrium with calcite. The effect of this on
the overall rate of pyrite oxidation is given below, as well as the effect on pyrite oxidation
mechanism dominance.

5.3.1 Overall rate


5.3.1.1

Partial pressure of oxygen

Figure 5-2 shows the effect of decreased Po2 on the total sulphate flux due to pyrite oxidation,
which, as indicated in Table 5-2, is the major source of sulphate ions. Table 5-4 shows the
resulting rate of oxygen consumption. For each fixed Po2 and in all three cases tested, the total
sulphate flux decreased with decreasing Po2 at constant infiltration rate. However, with
decrease in Po2 by two orders of magnitude below the base case value of 0.1 atm, the overall
rate of sulphide oxidation for the base case did not reach less than 10% of the original rate.
Figure 5-3 shows that decreasing Po2 led to increasing pH; for Po2 < 0.002atm, pH was above
neutral.
3.5

2.5

2-

normalised SO4 flux

3.0

base case
no Fe(OH)3
calcite, siderite, gypsum eqm

2.0
1.5
1.0
0.5
0.0
0.0001

0.001

0.01
Po2 (atm)

0.1

Figure 5-2. Area-normalised sulphate flux due to pyrite oxidation5, normalised to the base
case value, vs. Po2, for i) the base case, ii) no equilibrium with Fe(OH)3(am) and iii)
equilibrium with calcite, siderite and gypsum6.

Normalised to base case value, 7.7 x 10-7 mol m-2s-1, or concentrations of 0.1 M at the given flowrate
of 7.7 x 10-9 ms-1.
6
The case with no Fe(OH)3(am) equilibrium is oversaturated with respect to Fe(OH)3(am) for Po2<0.02
atm, and the case with calcite becomes undersaturated with respect to gypsum under the interval
0.01 atm<Po2<0.001 atm.

48

The sulphate fluxes shown indicate the rate of pyrite oxidation, which is inversely
proportional to pH. From comparison of Figures 5-2 and 5-3 it can be seen that at each fixed
Po2, the case with no Fe(OH)3(am) present had lower pH and sulphate flux than the base case.
Table 5-4. Area-normalised oxygen consumption (mol m-2s-1) as a function of Po2, for base
case, the case with no Fe(OH)3(am) and the case with equilbrium with calcite, siderite and
gypsum. Also shown are values normalised to the base case rate, that is, at Po2 = 0.1 atm (the
assumed average partial pressure assumed for the unsaturated zone).
Po2
0.2
0.1
0.01
0.001
0.0001

base case
rate
norm
1.8 x 10-6
1.36
1.3 x 10-6
1.00
6.0 x 10-7
0.46
2.6 x 10-7
0.20
1.1 x 10-7
0.09

no Fe(OH)3(am)
rate
norm
1.4 x 10-6
1.08
1.1 x 10-6
0.85
6.3 x 10-7
0.48

calcite etc.
rate
norm
3.9 x 10-6
3.0
2.8 x 10-6
2.12
8.8 x 10-7
0.67
2.8 x 10-7
0.21
8.8 x 10-8
0.07

The case with equilibrium with calcite had higher pH than the base case, but also a greater
sulphate flux. In all three cases, the sulphate concentration decreased and pH increased with
decreasing Po2. Thus at a fixed Po2, addition of calcite, or other means of buffering the pH at a
higher level, will increase pH, but also the rate of sulphide oxidation; decreasing oxygen
availability will both increase pH and decrease the rate of sulphide oxidation. Neutral pH was
reached in the base case when oxygen concentrations were just above 0.001atm (see Figure 53).

base case
no Fe(OH)3
calcite,siderite, gypsum eqm

pe

pH

base case pe

5
1
4
3
0.0001

0.001

0.01
Po2 (atm)

0.1

-1
0.0001

0.01
Po2 (atm)

Figure 5-3. pH and pe for various cases with decreasing Po2. The case without equilibrium
with Fe(OH)3(am) becomes oversaturated with respect to this mineral for Po2 < approximately
0.02 atm, and the case with calcite and gypsum is undersaturated with respect to gypsum for
Po2 < approximately 0.001 atm .

49

5.3.1.2

Infiltration rate and water content

Independently varied infiltration rate (q) and water content () had very little effect on the
overall rate; in reality, or in a model with diffusive transport, these parameters may affect Po2
(e.g. see Section 4.6). Variation in infiltration rate (q) over two orders of magnitude caused a
change of only 0.1 pH units, and 2 pe units, and had less effect on the overall rate of sulphide
oxidation than Po2 (Figure 5-4), though concentrations were directly affected as a function of
dilution. Variation of the water content had almost no effect on model results (Figure 5-5).
10
1.00
pH
pe
sulphate flux

pH, pe

8
7

0.50
6
5
4
1.00E-11

1.00E-10

1.00E-09

1.00E-08

normalised sulphate flux

0.00
1.00E-07

Infiltration rate, q (m s-1)

Figure 5-4. Effect of varied infiltration rate (q) on pH, pe and sulphate flux (normalised to
base case values), where flux = q x [SO42-]total. The infiltration rate in the base case was
7.7 x 10-9 m s-1.

pH, pe

4
0.05

pH
pe
sulphate flux

0.50

normalised sulphate flux

1.00

0.00

0.1

0.15

0.2

0.25

0.3

theta

Figure 5-5. Effect of varied water content () on pH, pe and the sulphate flux (normalised to
base case values). The base case value of theta was 0.15.

50

5.3.1.3

Pyrite fraction

Also considered was the effect of mineralogy, in particular, pyrite concentration (py), on
mechanism predominance (CuFeS2 and ZnS were proportionally scaled with py; gangue was not
changed). See Table 5-3 for the range over which the pyrite content was tested. Sulphate flux,
indicating the overall sulphide reaction rate, varied proportionally with the volumetric fraction
of pyrite (py) when py increased or decreased by an order of magnitude (see Figure 5-6).

Figure 5-6. Effect of varied volumetric fraction of pyrite (py) on sulphate flux normalised to
the base case value, pH and pe.

5.3.2

Pyrite oxidation mechanism dominance

Although it is theoretically possible that oxidation of pyrite by ferric iron can dominate under
certain conditions (see Figure 2-4), in this model oxidation by oxygen dominates for all Po2,
infiltration rate, water content and mineralogical compositions investigated (see Figure 5-7 for
the effect of Po2 and mineralogy). The contribution of oxidation by oxygen to the overall
pyrite oxidation rate was 95% at atmospheric Po2; dominance of this mechanism increased
with decreasing Po2, as the result of increasing pH and the resulting lower free concentration
of ferric iron.
The only case in which pyrite oxidation by ferric iron became significant to the overall rate
was when the oxidation of ferrous iron was increased by a factor of 102-103 (see Figure 3-4),
as can be the case when the reaction is microbially mediated. However, as the ferric iron
mechanism becomes more significant, the pH drops below field conditions, suggesting this
acceleration is not taking place in Impoundment 1.

51

Figure 5-7. Dominance of oxygen or ferric iron as pyrite oxidant as a function of pH, oxygen
concentration and volumetric fraction of pyrite. Impoundment 1 was reported to have 4%
pyrite; simulations shown with solid triangles.

52

6 Discussion
The ability of a model to represent a real system depends upon identification and accurate
representation of dominant processes, and availability of sufficient input data. During the
process of assembling the current model we discovered various shortcomings in the literature.
Kinetic weathering expressions missing in the literature include expressions for silicates, such
as talc and chlorite, and for metal sulphides, particularly for sulphides other than pyrite, such
as chalcopyrite and sphalerite. No expressions were found which quantified galvanic effects
occurring during the oxidation of complex sulphide mixtures.
Reported rates of biotic oxidation of pyrite and ferrous iron were compared to abiotic rates.
The comparison showed that, in accordance with general consensus, the biotic path may
dominate over a wide range of condition; however, quantitative rate expressions to base such
comparisons on are sparse in the literature.
It is not possible to produce a characterisation of a tailings impoundment that will imply in
detail to the entire impoundment, due to heterogeneity in, for example, chemical,
mineralogical, and hydrological factors. However, more detailed and/or accurate modelling of
Impoundment 1 requires additional field data to constrain the conceptual model, and to give
the necessary site-specific parameters. Desirable additional data, ideally in profiles, includes
complete aqueous phase analyses, among other reasons because it is from these analyses that
we are able to identify and trace the main processes occurring in the field. Full field
mineralogy is also important when building the conceptual model, including gangue fraction
composition and knowledge of which secondary phases are present. Similarly, knowledge of
redox conditions and gas phase composition can affect which processes are included and
hence how successfully the model will represent the impoundment geochemistry. Parameters
such as impoundment temperature, mineral surface area and variation in water content with
depth will also affect results. Species present, concentrations, and spatial distribution of
bacteria within the impoundment may give an indication as to whether microbial mediation of
reactions is occurring, and if so, where it is important.

53

7 Conclusions
The aim of these investigations was to develop a geochemical box model that could help
improve understanding of the processes occurring in a tailings impoundment, to reproduce the
important aspects of impoundment geochemistry and to predict the effect of remediation of
the impoundment. Field data from Impoundment 1, Kristineberg, was analysed to form a
general conceptual model. Rate expressions for slow processes controlled by chemical
kinetics, such as sulphide oxidation and silicate weathering, were compared and selected from
the literature; these were coupled to fast, equilibrium processes such as aqueous speciation
and solubility equilibrium with secondary minerals, to create a mathematical model of a
generalised tailings impoundment.
Input parameters were taken from field data and model results were compared to field
geochemistry. For pH, SO42-, Fe(tot), Cu2+ and K+, the uncalibrated model gave concentrations
within the range of the field values; Mg2+ concentrations were within a factor of 2 of field
averages and Al was an order of magnitude greater than the field average. The model was
further calibrated to field values, using two scaling factors, allowing for a) the scale
dependence of the rates of heterogeneous processes and b) biological mediation of sulphide
oxidation reactions. However, it should be noted that the deviation of model results from field
observations is small and can probably be accounted for by the uncertainties in the constants
and parameters used in the empirical rate laws employed. In general the essentially abiotic
model successfully reproduced most of the significant aspects of the impoundment
geochemistry.
Base case results include that pyrite oxidation was the major sink for oxygen, and that the
modelled total rate of oxygen consumption, 1.3 x 10-6 mol m-2s-1, was within the range of field
measurements reported for other sites in the literature. The oxygen path was also the
dominant mechanism for pyrite oxidation, and the major source of acidity; oxidation of pyrite
by ferric iron was however a significant source of ferrous iron, and a sink for protons. This
fact may help explain the low SO42- to Fe(tot) ratio found in the aqueous phase in the
impoundment. The major sink for protons was chlorite dissolution.
The presence of amorphous ferric hydroxide in the base case increased the overall rate of
pyrite oxidation, both by buffering pH at a higher level, which favoured pyrite oxidation by
oxygen, and increasing ferric iron concentrations. Comparison with the case without ferric
hydroxide present indicated that depletion of this mineral would lead to decreased pH and
Fe(tot) and SO42- concentrations. Application of these model results to the Kristineberg site
suggests that ferric hydroxide present in the impoundment at the point in time when the field
data was collected was dissolving, and that depletion of this mineral would have lead to
decreased Fe(tot) and SO42- concentrations and pH.
Sensitivity analyses using a simplified representation of oxygen diffusion indicated that the
pyrite oxidation rate may be controlled by either reaction or diffusion kinetics, suggesting that
coupling of oxygen diffusion and chemical kinetics would be required in a more
comprehensive model.
The base case model was used to simulate remediation of the impoundment. The parameters
likely to be affected by remediation that were separately varied were the partial pressure of
oxygen (Po2), the infiltration rate (q), the water content () and the volumetric fraction of
sulphides (sulphide). Model results were only sensitive to Po2 and sulphide, however, in reality
the infiltration rate and water content would affect the availability of oxygen. Decreasing Po2
led to decreasing total concentrations and increasing pH, with neutral pH achieved when Po2
was decreased by two orders of magnitude, to 0.001 atm. Equilibrium with calcite increased
pH but also sulphate concentrations, as the rate of pyrite oxidation is inversely proportional to
the proton concentration.

54

We tested the hypothesis that under the conditions created by remediation such as decreased
oxygen concentrations, the oxidation of pyrite by ferric iron would dominate over oxidation
by oxygen. However, simulations indicated that oxidation by oxygen will dominate for all
Po2, infiltration rates, degrees of saturation and mineralogical compositions investigated,
unless the oxidation of ferrous iron were to be accelerated by a factor greater than 103. Such
acceleration may be possible due to microbial catalysis of reactions, however, results suggest
that this is not the case at the Kristineberg site.
The current investigation has highlighted that more detailed and/or accurate modelling of
AMD requires additional qualitative and quantitative field data, in order to facilitate the
correct choice of processes in the conceptual model, and to give the necessary site-specific
parameters. Specifically for the case study of Impoundment 1 in Kristineberg, missing data
included detailed mineralogical information, such as observations confirming/rejecting the
presence of ferric hydr(oxides) and carbonate minerals, as well as aqueous groundwater and
pore gas concentrations of key components, such as oxygen and carbon dioxide. Additionally,
process representation such as rate expressions quantifying the weathering of sphalerite and
chalcopyrite by oxygen are essential for modelling of impoundment geochemistry. Lack of
this information resulted in limited interpretative and predictive capability of the model.
Extensions of the present investigation could include the coupling of different boxes to
represent zones of different conditions within the impoundment (c.f. Berg, 1997) and
validation of the model, by application to other field sites. Further investigations that could be
of interest include a deepened analysis of coupling of the geochemistry to diffusion of the gas
phase and/or hydrological transport models, and investigation of the importance of secondary
processes such as precipitation and adsorption onto mineral surfaces; for example, these
processes may have an important effect on trace element mobility. The effect of uncertainty in
parameters and empirical rate expressions, and the effect of heterogeneity in the
impoundment that results in variation in geochemical input parameters and impoundment
hydrology, could also be investigated.

55

References

Allison J.D., Brown D.S. and Novo-Gradac K.J. (1991) MINTEQA2/PRODEFA2, A geochemical
assessment model for environmental systems; Version 3.0 Users Manual, Environmental
Research Laboratory, Office of Research and Development, U.S. Environmental Protection
Agency, EPA/600/3-91/021, Athens, Georgia, 30605, 92p
Appleyard E.C. and Blowes D.W. (1994) Applications of mass-balance calculations to weathered
sulfide mine tailings. . In: Environmental geochemistry of sulfide oxidation, Alpers C.N and
Blowes D.W. (Eds) ACS symposium series 550:516-534
Amrhein C. and Suarez D.L. (1988) The use of a surface complexation model to describe the kinetics
of a ligand-promoted dissolution of anorthite. Geochimica Cosmochimica Acta 52:2785-2793
Axelsson C-L., Ekstav A., Jansson T. (1991) Provtagning av sand och grundvatten i sandmagasin 1, 1B
och 2 Kristineberg Fltrapport. Golder Geosystems AB, Rapport 917-1687, (in Swedish).
Axelsson, C.-L., Karlqvist L., Lintu Y. and Olsson T. (1986) Gruvindustrins restproduktupplag fltunderskningar med vattenbalansstudie i Kristineberg, Uppsala Geosystem AB, (in
Swedish, with Summary in English).
Ball J.W. and Nordstrom D.K. (1991) Users manual for WATEQ4F, with revised thermodynamic
database and test cases for calculating speciation of major, trace and redox elements in natural
waters, U.S. Geological Survey Open-File Report 91-183, 189p. (revised and reprinted August
1992).
Banwart, S. and Malmstrm M (1999) Hydrochemical modelling for preliminary assessment of
minewater contamination, submitted to Journal of Geochemical Exploration (in review)
Berg, A. (1997) The Response of soil weathering to climate change; Laboratory and modelling studies.
PhD thesis, Dept. of Chemistry, Inorganic Chemistry, Royal Institute of Technology,
Stokcholm, Sweden
Bigham J.M., Schwertmann U., Traina S.J., Winland R.L. and Wolf M. (1996) Schwertmannite and the
chemical modeling of iron in acid sulfate waters. Geochimica Cosmochimica Acta 60:21112121
Blowes D.W. and Ptacek C.J. (1994) Acid-neutralization mechanisms in inactive mine tailings. In:
Environmental Geochemistry of Sulphide Mine-Wastes J.L.Jambor and D.W. Blowes, (Eds)
Short course handbook, Mineral Association of Canada, 22:271-292
Chou L., Garrels R.M. and Wollast R. (1989) Comparative study of the kinetics and mechanisms of
dissolution of carbonate minerals. Chemical Geology 78:269-282
Ekstav A. and Qvarfort U. (1989) Metallbalans Kristineberg. Kvartrgeologiska avd., Uppsala
universitet, Uppsala (in Swedish)
Elberling B. and Nicholson R.V. (1996) Field determination of sulphide oxidation rates in mine
tailings. Water Resources Research 32(6):1773-1784
Evangelou V.P. (1995) Pyrite Oxidation and its control. CRC Press, Florida
Evangelou V.P. and Zhang Y.L. (1995) A review: Pyrite oxidation mechansims and acid mine drainage
prevention. Critical Reviews in Environmental Science and Technology 25(2):141-199
Furrer G., Westall J. and Sollins P. (1989) The study of soil chemistry through quasi-steady-state
models: I. Mathematical definition of model. Geochimica Cosmochimica Acta 53:595-601
Furrer G., Westall J. and Sollins P. (1990) The study of soil chemistry through quasi-steady-state
models: II. Acidity of soil solution. Geochimica Cosmochimica Acta 54:2363-2374
Helgeson H.C., Murphy W.M., Aagaard P (1984) Thermodynamic and kinetic constraints on reaction
rates among minerals and aqueous solutions II. Rate constants, effective surface area, and the
hydrolysis of feldspar. Geochimica Cosmochimica Acta 48:2405-2432
Herbert R. (1999) Sulphide oxidation in mine waste deposits: A review with emphasis on dysoxic
weathering. MiMi Report, Mitigation of the environmental impact from mining waste program
(MiMi), Stockholm (in press)
Jambor J.L. (1994) Mineralogy of sulfide-rich tailings and their oxidation products. In: Environmental
Geochemistry of Sulphide Mine-Wastes J.L.Jambor and D.W. Blowes, (Eds) Short course
handbook, Mineral Association of Canada, 22:59-102
Jaynes D.B. Rodowski A.S. and Pionke H.B. (1984) Acid mine drainage from reclaimed coal strip
mines 1. Model description. Water Resources Research 20(2):233-242
Knauss K.G. and Wolery T.J. (1989) Muscovite dissolution kinetics as a function of pH and time at
70C. Geochimica Cosmochimica Acta 53:1493-1501
Kwong Y.T.J. (1995) Influence of galvanic sulfide oxidation on mine water chemistry. Proceedings of
Sudbury '95, Conference on Mining and the Environment II, Sudbury, Ontario, May 28 June 1,
1995, 477-483

56

Lacey D.T., Lawson F. (1970) Kinetics of the liquid-phase oxidation of acid ferrous sulfate by the
bacterium Thiobacillus ferrooxidans. Biotechnol. Bioeng. 12:29-50.
Lindvall M., Eriksson N. and Ljungberg J. (1999) Decommissioning at Kristineberg Mine, Sweden.
Proceedings of Sudbury '99, Conference on Mining and the Environment II, Sudbury, Ontario,
September 13 - 15, 1999, 3:855-862
Lowson, R.T., (1982) Aqueous oxidation of pyrite by molecular oxygen. Chemical Reviews 82(5): 461497
Malmstrm M., Werner K., Salmon U. and Berglund S (1999b) Hydrogeology and geochemistry of
mill tailings impoundment 1, Kristineberg, Sweden: Compilation and interpretation of preremediation data. MiMi Report, Mitigation of the environmental impact from mining waste
program (MiMi), Stockholm (in revision)
Malmstrm M. and Banwart S. (1997) Biotite dissolution at 25C: The pH dependence of dissolution
rate and stoichiometry. Geochimica Cosmochimica Acta 61(14):2779-2799
Malmstrm M., Banwart S., Duro L., Wersin P. and Bruno J. (1995) Biotite and chlorite weathering at
25C: The dependence of pH and (bi)carbonated on weathering kinetics, dissolution
stoichiometry and solubility; and the relation to redox conditions in granitic aquifers. Technical
Report 95-01, The Swedish Nuclear Fuel and Waste company (SKB), Stockholm
Malmstrm M.E., Destouni G., Banwart S.A. and Strmberg B.H.E. (1999a) Resolving the scaledependence of mineral weathering rates. (Accepted for publication in Environmental Science
and Technology)
McKibben M.A. and Barnes H.L. (1986) Oxidation of pyrite in low temperature acidic solutions: Rate
laws and surface textures. Geochimica Cosmochimica Acta 50:1509-1520
Millero F.J. (1985) The effect of ionic interactions on the oxidation of metals in natural waters.
Geochimica Cosmochimica Acta 49:547-553
MiMi (1997) MiMi programme plan for the period 1998-2000, Mitigation of the environmental impact
of mining waste. MiMi-report, Mitigation of the environmental impact from mining waste
program (MiMi), Stockholm
Moses C.O., Nordstrom D.K., Herman J.S. and Mills, A.L. (1987) Aqueous pyrite oxidation by
dissolved oxygen and by ferric iron. Geochimica Cosmochimica Acta 51:1561-1571
Nagy K.L. (1995) Dissolution and precipitation kinetics of sheet silicates. In: Chemical weathering
rates of silica minerals. White A.F. and Brantley S.L. (Eds), Review of Minerals 31, Mineral
Society of America
Nemati M., Webb C. (1997) A kinetic model for biological oxidation of ferrous iron by Thiobacillus
ferroooxidans. Biotechnol., Bioeng. 53(5):478-486.
Nemati M., Harrison S.T.L., Hansford G.S., Webb C. (1998) Biological oxidation of ferrous sulphate
by Thiobacillus ferrooxidans: a review on the kinetic aspects. Biochem. Eng. J. 1:171-190.
Nicholson R.V. (1994) Iron-sulphide oxidation mechanisms: Laboratory studies. In: Environmental
Geochemistry of Sulphide Mine-Wastes J.L.Jambor and D.W. Blowes, (Eds) Short course
handbook, Mineral Association of Canada, 22:163-183
Nicholson R.V., Gillham R.W. and Reardon E.J. (1988) Pyrite oxidation in carbonate buffered
solution: 1 Experimental kinetics. Geochemica et Cosmochimica Acta 52:1077-1085
Nicholson R.V., Gillham R.W. and Reardon E.J., (1990) Pyrite oxidation in carbonate-buffered
solution: 2. Rate control by oxide coatings. Geochemica et Cosmochimica Acta 54:395-402
Nordstrom D.K. and Alpers C.N. (1999) Geochemistry of acid mine waters. In: Environmental
geochemistry of mineral deposits; Plumlee G.S. and Logsdon M.J. (Eds), Reviews in Economic
Geology, 6A, Society of Economic Geologists
Nordstrom D.K. and Ball J.W. (1984) The geochemical behavior of aluminum in acidified surface
waters. Science 232:54-56
Nordstrom D.K., Plummer N., Langmuir D., Busenberg E., May H.M., Jones B.F. and Parkhurst D.L.
(1990) Revised chemical equilibrium data for major water-mineral reactions and their
limitations. In: Chemical Modeling of aqueous systems II, Melchior D.C. and Bassett R.L.,
(Eds), ACS Symposium Series 416:398-413
Nordstrom D.K. and Southam G. (1997) Geomicrobiology of sulfide mineral oxidation. In: Reviews in
Mineralogy: Geomicrobiology: interactions between microbes and minerals, Banfield J.F. and
Nealson K.H. (Eds), Mineral Association of Canada, 35:361-390
Nyavor K., Egiebor N.O., Fedorak P.M. (1996) The effect of ferric iron on the rate of ferrous oxidation
by Thiobacillus ferrooxidans. Appl. Microbiol Biotechnol. 45:688-691.
Parkhurst D.L. (1995) Users guide to PHREEQC a computer program for speciation, reaction-path,
advective transport and inverse geochemical calculations. US Geological Survey WaterResources Investigations Report 95-4227, 143p

57

Pesic, B., Oliver, D.J. and Wichlacz, P. (1989) An electrochemical method of measuring the oxidation
rate of ferrous to ferric iron with oxygen in the presence of Thiobacillus ferrooxidans.
Biotech.Bioeng.33:428-439
Qvarfort U. (1983) En kartlggning av sandmagasin frn sulfidmalmsbrytning. Naturvrdsverkets
Rapport SNV PM 1699, The Swedish national environment protection board, Stockholm (in
Swedish).
Qvarfort U. (1989) Sandmagasin frn sulfidmalmsbrytningen: En kartlggning och inventering.
Naturvrdsverkets Rapport SNV 3587, The Swedish national environment protection board,
Stockholm (in Swedish).
Reardon E.J. and Moddle P.M. (1985) Gas diffusion coefficient measurements on uranium tailings
measurements: Implications to cover layer design. Uranium 2:11-131
du Rietz T. (1953) Geology and ores of the Kristineberg deposit, Vesterbotten, Sweden. Sveriges
Geologiska Underskning, Ser C. 524, rsbok 45(5) (1951)
Rimstidt J.D., Chermak J.A. and Gagen P.M. (1994) Rates of reaction of galena, sphalerite,
chalcopyrite and arsenopyrite with Fe(III) in acidic solutions In: Environmental geochemistry of
sulfide oxidation, Alpers C.N and Blowes D.W. (Eds) ACS symposium series, 550:2-13
Rimstidt J.D. and Newcomb W.D. (1993) Measurement and analysis of rate data: The rate of reaction
of ferric iron with pyrite. Geochimica Cosmochimica Acta 57:1919-1934
Scharer J.M., Nicholson R.V., Halbert B. and Snodgrass W.J. (1994) A computer program to assess
acid generation in pyritic tailings. In Environmental Geochemistry of Sulfide oxidation, Alpers
C.N. and Blowes D.W. (Eds), ACS Symposium Series 550:132-152
Schrenk MO, Eswards KJ, Goodman RM, Hamers RJ, Banfield JF, (1998) Distribution of Thiobacillus
ferrooxidans and Leptospirillum ferrooxidans: Implications for Generation of Acid Mine
Drainage. Science 279:1519-1522
Singer P.C. and Stumm W. (1970) Acid mine drainage: The rate-determining step. Science, 167:11211123
Strmberg B. (1997) Weathering kinetics of sulphidic mining waste: An assessment of geochemical
processes in the Aitik mining waste rock deposits. PhD thesis, Dept. of Chemistry, Division of
Inorganic Chemistry, Royal Institute of Technology, Stockholm
Strmberg B. and Banwart S. (1994) Kinetic modelling of geochemical processes at the Aitik mining
waste rock site in northern Sweden. Applied Geochemistry 9: 583-595
Stumm W. and Morgan J.J. (1996) Aquatic chemistry. 3rd edition, John Wiley and Sons, Inc., New
York
Sung W. and Morgan J.J. (1980) Kinetics and product of ferrous iron oxygenation in aqueous systems.
Environmental Science and Technology 14:561-568
Tamura H., Goto K. and Nagayama M. (1976) The effect of ferric hydroxide on the oxygenation of
ferrous ions in neutral solutions. Corrosion Science 16:197-207
Wehrli B. (1990) Redox reactions of metal ions at mineral surfaces. In: Aquatic chemical kinetics
Stumm W. (Ed), Wiley, New York, 311-336
Wiersma C.L. and Rimstidt J.D. (1984) Rates of reaction of pyrite and marcasite with ferric iron at pH
2. Geochimica Cosmochimica Acta 48:85-92
Williamson M.A. and Rimstidt J.D. (1994) The kinetics and electrochemical rate-determining step of
aqueous pyrite oxidation. Geochimica Cosmochimica Acta 58(24):5443-545
Wollast R. (1990) Rate and mechanism of dissolution of carbonates in the system CaCO3- MgCO3. In:
Aquatic chemical kinetics Stumm, W. (Ed), John Wiley & Sons, Inc.
Wollast R. and Chou L. (1985) Kinetic study of the dissolution of albite with a continuous flowthrough fluidized bed reactor. In: The chemistry of weathering, Drever J.I. (Ed), Nato ASI Ser C,
V.149, Rediel. Publishing Company, pp.75-96
Zhang Y., Charlet L. and Schindler R.W. (1992) Adsorption of protons, Fe(II) and Al(III) on
lepidocrocite (-FeOOH). Colloids and Surfaces, 63:259-268

58

9 Appendices
Appendix A. Pyrite oxidation rate expressions
Table APP-A1. Various empirical kinetic rate expressions for oxidation of pyrite by
oxygen from the literature (DO= dissolved oxygen)
Reference
Rate expression
pH
TC
Details
0.5( 0.04 )
rate (mol FeS2 m-2s-1)
Williamson
m
m (mol kg-1)
r = 10 8.19( 0.10) DO
and Rimstidt
2-10
25
( 0.01)
m H0.11
+

(1994)

Nicholson et
al. (1990)

McKibben
and Barnes
(1986)*
Nicholson et
al. (1988)

d [FeS2 ]
4 d [O2 ]
4
=
= k s SC n
7.6-6.8
dt
15 dt
15

rO2 = 10 6.77 [O2 ]

0.5

2-4

25

30

rate (mol FeS2 h-1)


ks (m h-1) = 3.07 x 10-6 46%
S surface area exposed to
FeS2(m2)
C O2 in gas phase (mol m-3)
n=1
rate (mol pyrite cm-2min-1)
[ ] molarity
rate (mol FeS2 s-1)
C O2 in gas phase (mol m-3)
k (best fit, n=0.5) = 4.27 x 10-8

r = kC n

Fit to data in their Figure 7, using


6.7-8.5 3-25
KH=1.3 x 10-3 mol l-1atm-1 and
A=222cm2 g-1
*The rate has been adjusted according to comment in Nicholson (1994), that units given in the
text do not fit with the data shown. The adjusted rate (by a factor of 103) also is in closer
agreement with other rates from the literature, which supports the comment by Nicholson.

Table APP-A2. Various empirical kinetic rate expressions for oxidation of pyrite by
Fe(III) from the literature
Reference
Rate expression
pH
TC
Details
rate (mol Fe3+ m-2s-1)
m (mol kg-1)
Wiersma
k (s-1) = 1.90 x 10-446%
dmFe3+
A
and
2
25
(k varies with [Fe3+]initial)
=k
mFe3+
Rimstidt
M
dt
A (m2)
(1984)
M mass water (kg)
A/M standard system=1m2 kg-1
McKibben
rate (mol FeS2 cm-2 min-1)
9.74
3+ 0.5
+ 0.5
and Barnes
1-2
25-30 m (mol kg-1)
rFe3+ = 10
Fe
H
(1986)
Rimstidt
rate (mol Fe3+ m-2s-1)
and
m (mol kg-1)
0.62 ( 0.10 )
rFe3+ = 3( 2 )x10 5 mFe
2
25
3+
Newcombe
(1993)
0.30 ( 0.02 )
rate (mol FeS2 m-2s-1)
Williamson
m
3+
m (mol kg-1)
r = 10 8.58( 0.15) 0.47 ( 0Fe
and
.03 ) 0.32 ( 0.04 ) 0.5-3
25
mFe 2+
mH +
Rimstidt
(1994)
(absence of dissolved oxygen)

] [ ]

59

Appendix B. Pyrite oxidation mechanism predominance.


In this appendix we show how the regions, in terms of pH and PO2, where different paths for
pyrite oxidation dominate were framed.
Speciation

To assess the pyrite oxidation rates we need to describe the speciation of ferric iron, which is
one of the two considered oxidants. We consider two separate case where a) the free
concentration of ferric iron is fixed by solubility equilibrium with an amorphous
ferric(hydr)oxide; or b) the total concentration of ferric iron in solution is known. We then
have the following set of equilibrium reactions with corresponding equilibrium expressions:

[Fe ]
[H ]
[FeOH ][H ]
K =
[Fe ]
[Fe(OH ) ][H ]
K =
[Fe ]
[Fe(OH ) (aq)][H ]
=
[Fe ]
[Fe(OH ) ][H ]
K =
[Fe ]

Fe(OH ) 3 (am) + 3H + Fe 3 + + 3H 2O

3+

Ks =

+ 3

2+

Fe 3+ + H 2O FeOH 2+ + H +
Fe 3+ + 2 H 2 O Fe(OH ) +2 + 2 H +

3+

H1

+
2
3+

H2

(B1)
(B2)

+ 2

(B3)

+ 3

Fe 3+ + 3H 2O Fe(OH )3 (aq ) + 3H +

KH3

Fe 3+ + 4 H 2O Fe(OH ) 4 + 4 H +

H4

3+

4
3+

(B4)

+ 4

(B5)

In the case of solubility equilibrium, the free concentration of ferric iron is directly
given by Equation B1

[Fe ] = K [H ]

+ 3

3+

(B1a)
whereas for the other case, it can be solved by combining Equations B2-B5 with the mass
balance condition, [Fe( III )]T = [Fe3+ ] + [FeOH 2+ ] + [Fe(OH ) 2+ ] + [Fe(OH )3 (aq )] + [Fe(OH ) 4 ] which
can be formulated in terms of the free concentration of Fe3+ as
s

[Fe( III )]T

K
K
K
K
= Fe3+ 1 + H+1 + H 22 + H 33 + H 44 . This yields the ferric iron concentration as

H
H+
H+
H+

[Fe( III )]T


(B6)
Fe3 + =

] [ ]
[ ] [ ] [ ]
[ ]

where is defined by
=1+

K H1

KH 2

KH3

KH 4

[H ] + [H ] + [H ] + [H ]
+

+ 2

+ 3

+ 4

(B7).

In this analysis, we have only considered mononuclear hydrolysis species. For a more detailed
analysis, a full speciation model is needed.

60

Comparing rates of different mechanisms

We compare three different paths for oxidation using the corresponding empirical rate laws

Oxygen path (compare Equation 2-10)

Ferric iron path (compare Equation 2-11)

(B8);

(B9);

[O2 (aq)]
1
2.5 pH
K o + [O2 (aq)] 1 + 10
+ 10 pH 4

(B10).

R fer = k fer Fe 3+

[ ]

Roxy = k oxy [O2 (aq)] H +


n

Microbial path (compare Equation 2-18)

Rbio = b ' k bio

Oxygen path dominates over ferric iron path

We start by assessing the abiotic reactions and by searching for the region where the
oxygen path dominates over the ferric iron path, i.e. Roxy>Rfer. We do this by
comparing Equation B8 and B9:
n
l
m
k fer [Fe3+ ] < kbio [O2 (aq)] [H + ]
(B11).
For the case where we have solubility control of the ferric iron concentration, this
equation in combination with Equation B1a yields
k fer n
l
1
3n
(B12).
log[O2 (aq )] > log
+ log K S
pH + pH

m k oxy
m
m
m
For the case where we do not have solubility control, we combine Equations B11 and
B6 to yield
log[O2 (aq)] >

k fer n
1
n
l
log
+ log[Fe( III )]T log + pH
k oxy
m
m
m
m

(B13).

Microbial path dominates over oxygen path


Next, we search for the region where the microbial path dominate over the oxygen
path, Roxy<Rbio, by comparing Equations B8 and B10
[O2 (aq)]
l
1
m
b ' k bio
> k oxy [O2 (aq)] H +
(B14).
pH 4
2.5 pH
K o + [O2 (aq)] 1 + 10
+ 10
For simplicity we define = 1 + 10 2.5 pH + 10 pH 4 and rearrange the equation to

[ ]

(K o + [O2 (aq)])
(B14a)
[O2 (aq)]
[ ]
which we first solved in the regime where [O2 (aq)] >> K o . Here Equation B14a
b ' k bio 1 1
k oxy H +

> [O2 (aq)]

approaches
b ' kbio 1 1
k oxy H +

[ ]

> [O2 (aq )]

(B14b)

or
log[O2 (aq)] <

b 'k 1
1
1 l
log bio + log + pH
k oxy
m
m
m

61

(B14c).

For the regime where [O2 (aq)] << K o , Equation B14a reduces to
b ' k bio 1 1
k oxy H +

[ ]

> [O2 (aq)]

m 1

(B14d)

Ko

which after rearrangement yields


log[O2 (aq )] <

b ' kbio
l
1
1
1
pH
log
+
log +
(m 1) koxy K o (m 1) (m 1)

(B14e).

Microbial path dominates over ferric iron path


Finally, we search the region where the microbial path dominates over the ferric iron
path, Rbio>Rfer, by comparing Equations B9 and B10
b ' kbio

[O2 (aq)] 1
ko + [O2 (aq)]

> k fer Fe3+

(B15).

We start by addressing the regime where [O2 (aq)] >> K o , in which Equation B15,

remembering that = 1 + 10 2.5 pH + 10 pH 4 = 1 + a[H + ] +

b
, where a=102.5 and b=10-4,
H+

[ ]

reduces to

[ ] [ ][

b ' k bio
b
> 1 + a H + + + Fe 3+
k fer
H

(B15b).

First we consider the case where ferric iron concentration is controlled by solubility
equilibrium, that is we combine Equations B1a and B15b to yield

[ ]

+a H+

[ ]

+b H+

b ' kbio
> H+
k fer K Sn

3n

[ ]

3 n+1

[ ]

(B15c).

b ' kbio
=0
k fer K Sn

(B15d)

+b H+

3 n1

We formulate this equation as

[H ]

+ 3n

+a H+

3 n +1

[ ]

3 n 1

and solve numerically for [H+] using the Newton-Raphson method.


For the case where we do not have solubility control of ferric iron we combine
Equations B6 and B15b to

[Fe( III )]T


b ' kbio
>
k fer
n

(B15e).

We insert the expressions for and and rearrange the equation to yield

[ ] [ ]

1+ a H

b ' k bio
b
+ +
n
H
k fer [Fe( III )]T

1 + K H 1 + K H 2 + K H 3 + K H 4
=0
+
2
3
4
+
+
+

H
H
H
H

[ ] [ ] [ ] [ ]

(B15f)

which we solve for [H+] by the Newton-Raphson method.

Now, we have to address the regime where [O2 (aq)] >> K o . Equation B15 then reduces
to

b ' kbio 1
[O2 (aq)] > k fer Fe3+
ko

(B15g).

Again, we start by addressing the case where we have solubility control on the ferric
iron concentration. We do this by inserting Equation B1a in B15g

( [ ])

b ' kbio 1
[O2 (aq)] > k fer K s H +
ko

62

3 n

(B15h)

which we can write as


k fer K o
log[O2 (aq )] > log '
+ n log K s 3npH
b kbio

(B15i).

Finally, we combine Equations B6 and B15g to address the case where we do not
have a solubility control on the ferric iron concentration. This gives

b ' k bio 1
[O2 (aq)] > k fer [Fe( III )]T
Ko

which, after rearrangement gives

(B15j)

k fer K o
+ n log[Fe( III )]T n log
log[O2 (aq )] > log '

b kbio

(B15k).

Figure APP-B1 and Figure 2-5 show the resulting predominance diagram.

Fraction diagram

For the abiotic reactions, we have also generated a mechanism fraction diagram. This
was done by inserting the equations for the free concentration of ferric iron (Equation
B1a and B6) in the following equations yielding the rate fractions Xfer and Xoxy for the
ferric iron and the oxygen paths, respectively. The results are shown in Figure 2-4.
X fer =

[ ]
[O (aq)] [H ] + k [Fe ]
k [O (aq )] [H ]
[O (aq)] [H ] + k [Fe ]
k fer Fe3+

k oxy

+ l

(B16)

+ l

oxy

k oxy

3+ n

fer

X oxy =

+ l

3+ n

fer

63

(B17).

a)

b)
Figure APP-B1. Mechanism predominance diagram for pyrite oxidation at T=1oC
and Ko=1 x 10-6 M. The full and dotted lines denote b=1 and b=0.01, respectively.
Line A is obtained from Equation B14c, line B from B12, Line C by numerically
solving B15d for [H+], Line D is obtained from Equation B14e, line E from B15k,
line F from B13, line G by numerically solving Equation B15f for [H+], and line H is
obtained from Equation B15i. a) Ferric iron concentration controlled by solubility
equilibrium with an amorphous Fe(III)-(hydr)oxide. b) [Fe(III)]T=1 x 10-6 M (only
hydrolysis species accounted for). Equilibrium and rate constants used in the model
are listed in Table APP-B1.

64

Table APP-B1. Constants used to yield model results shown in Figures 2-5 and APPB1.
Constant Value
Source
Comment
-8
kfer
8.1 x 10
Rimstidt and
Corrected to T=1oC (Ea=92 kJ mol-1;
Newcomb, 1993
Wiersma and Rimstidt, 1984)
n
0.62
-koxy
8.7 x 10-10 Williamsson and
Corrected to T=1oC (Ea=56.9 kJ mol-1;
Rimstidt, 1994
McKibben and Barnes, 1986)
l
-0.11
-m
0.5
-kbio
1.0 x 10-8
Section 2.2.2
Corrected to T=1oC (Ea=52.7 kJ mol-1;
Scharer et al., 1994)
a
320
Recalculated from Scharer et al., 1994
b
1.0 x 10-4
Recalculated from Scharer et al., 1994
-6
Ko
1.0 x 10
Compare Figure 2-3
In analogy with Jaynes et al., 1984
KH1
5.0 x 10-4
Nordstrm et al., 1990 Corrected to T=1oC, I= 0.27
KH2
3.7 x 10-8
Nordstrm et al., 1990 Corrected to T=1oC, I= 0.27
KH3
1.6 x 10-15 Nordstrm et al., 1990 Corrected to T=1oC, I= 0.27
KH4
8.1 x 10-25 Nordstrm et al., 1990 Corrected to T=1oC, I= 0.27
KS
1.1 x 106
Nordstrm et al., 1990 Corrected to T=1oC, I= 0.27

65

Appendix C.
predominance

Ferrous

iron

oxidation

mechanism

In this appendix we show how the regions where different paths for ferrous iron oxidation
dominate were framed, in terms of pH and surface site concentration.
As shown in Section 2.3 , the abiotic rate of ferrous iron oxidation depends on the speciation.
If we consider hydrolysis and sulphate complexation as well as surface adsorption, the
following expressions form the mass balance criteria
[Fe( II )]T ,aq = [Fe 2+ ] + [FeOH + ] + [Fe(OH ) 2 (aq)] + [FeHSO4+ ] + [FeSO4 (aq)]
(C1)

{> FeOH }T

[SO ] = [SO ] + [HSO ] + [FeHSO ] + [FeSO (aq)]


= {> FeOH } + {> FeO }+ {> FeOH }+ {> FeOFe }+ {> FeOFeOH }
2
4 T

2
4

+
4

+
2

(C2)

(C3).
The species may all be formed from the components H+, Fe2+, SO42-, and >FeOH and their
concentrations are determined by the corresponding equilibrium expressions

[FeOH ][H ]
[Fe ]
+

Fe 2+ + H 2O FeOH + + H +

K1 =

Fe 2+ + 2 H 2O Fe(OH ) 2 (aq) + 2 H +

K=

Fe 2+ + SO42 + H + FeHSO4+

K K1 =

2+

Fe 2 + + SO 42 FeHSO

KK2

2+

+
4

[Fe(OH) 2 (aq)][H + ]2

[Fe ]
2+

[FeHSO ]
[Fe ][SO ][H ]
[FeSO (aq)]
=
[Fe ][SO ]
+
4

2
4

(C7)

2
4

[HSO ]
[H ][SO ]
{> FeO }[H ]
=

K sa =

> FeOH > FeO + H +

Kb

> FeOH + H + > FeOH 2+

Ka =

> FeOH + Fe 2+ > FeOFe + + H +

K s1

(C8)

2
4

(C5)
(C6)

H + + SO42 HSO4

> FeOH + Fe 2+ + H 2O > FeOFeOH + 2 H +

(C4)

2+

(C9)

{> FeOH }

{> FeOH }
{> FeOH }[H ]
{> FeOFe }[H ]
=
{> FeOH }[Fe ]
{> FeOFeOH }[H ]
=
{> FeOH }[Fe ]
+
2

(C10)

2+

(C11)

+ 2

Ks2

2+

(C12)

>FeOH denotes a surface site and curved brackets refer to a surface concentration.
We need to solve the concentrations of the components from the mass balance conditions
and the equilibrium expressions.

66

Solution speciation

By combining Equation C1 and C2 with C4-C8, respectively, we obtain

K1
K2
+
+ K K 1 H + SO42 + K K 2 SO42
+
2
+
H
H

(C1a)

[SO ] = [SO ](1 + K [H ] + K [Fe ] + K [H ][Fe ])

(C2a)

[Fe( II )]T ,aq = [Fe 2+ ]1 +

and

2
4
T

[ ][

[ ] [ ]

2
4

K2

2+

K1

By inserting Equation C2a in C1a we get

2+

SA

[Fe( II )]T ,aq = [Fe 2+ ]1 +

[ ][ ]
[ ]
[ ] [ ] [ ][

K K 1 H + SO42 T + K K 2 SO42 T
K1
K2
+
+
+
2
+
2
+
+
2
+
H
1 + K SA H + K K 2 Fe + K K 1 H Fe
H+

[ ] [ ] (

])

(C1b)
which can be solved for [Fe2+] as

[Fe ] =
2+

where
a = 1+

b
Fe 2+
ad

ab + c + d Fe 2+
+
2ad

[ ][

[ ]

K1
K2
+
; b = 1 + K SA H + ; c = K K 1 H + SO42
2
+
H
H+

[ ] [ ]
+ K [H ] .

d = KK2

ab + c + d Fe 2+
2ad

+ K K 2 SO42

(C13)

; and

K1

We then obtain the free sulphate concentration from Equation C2a.


Surface speciation

When the free concentration of Fe2+ has been determined we can combine Equations C9
C12 and C3 to derive the free concentration of >FeOH from the corresponding total
concentration

{> FeOH }T = {> FeOH }1 +

[ ]

[ ]

[ ]

and, hence,

{> FeOH } = {> FeOH }T where we define

Kb
K S1 Fe 2+
K S 2 Fe 2+
+
+
+
+
K
H
a
2
H+
H+
H+

[ ]

(C3a)

K
K Fe 2 +
K Fe 2 +
= 1 + b+ + K a H + + S1 +
+ S2 2
.

H
H
H+

[ ]

[ ]

[ ]

[ ]

Using these equations, we can have a brief look at fraction diagrams for the dissolved
ferrous iron and for the surface species (See Figures APP-C1 APP-C2).

67

2+

Fe

FeOH

0.75

0.5

0.25
Fe(OH)2 (aq)

0
2

10

pH

a)

FeSO4 (aq)

0.75

0.5

FeOH

0.25

2+

Fe

0
2

6
pH

10

b)

Figure APP-C1. Aqueous speciation of ferrous iron. a) In the absence of complexing


ligands. b) In presence of sulphate. [Fe2+]T=1 x 10-3 M and [SO42-]T=1 x 10-2 M.
Constants used in the model are listed in Table APP-C1.

68

>FeOH2

>FeOH
>FeO

0.75

0.5

0.25

0
2

10

pH

a)

>FeOH2

>FeOFeOH

0.75

0.5

0.25
>FeOH

0
2

6
pH

10

b)
FigureAPP-C2. The surface speciation of FeOH as simplistically modelled by
Equations C9-C12. a) In absence of ferrous iron. b) At [Fe2+]T=1 x 10-3 M and [SO42]T=0 M and {>FeOH}T=1 x 10-6 mol -2. Constants used in the model are listed in
Table APP-C1.

69

Comparing rates of different mechanisms

Having sorted out the speciation of the surface sites and of ferrous iron, we address
the rates of oxidation by the different paths:

Homogeneous oxidation (see Equation 2-24)


RHom = k o [Fe 2 + ] + k1 [FeOH + ] + k 2 [Fe(OH ) 2 (aq )] [O2 (aq )]

(C14);

Heterogeneous oxidation (compare with Equation 2-30)


A
R Surf =
k Surf ({> FeOFe + }+ {> FeOFeOH })[O 2 ( aq ) ]

(C15);

Microbial oxidation (see Equation 2-31)


Rbio = k bio Cbact . [Fe 2 + ][O2 (aq )][H + ]
at pH>2

(C16).

Note that Equation C15 has a different form than that originally suggested by Wehrli
(1990). This is accordance with the representation made by Strmberg and Banwart
(1994) and accounts for the additional surface species >FeOFeOH, which is reported
by Zhang et al. (1992) but not addressed by Wehrli (1990).
We begin by framing the regions where the heterogeneous reaction dominates
over the homogeneous. We do this by comparing Equations C14 and C15

] [

k o Fe 2 + + k1 FeOH + + k 2 [Fe(OH ) 2 (aq)] <

({

A
k Surf > FeOFe + + {> FeOFeOH }
V

(C17)

which can be expressed as


k o + k1

K1
K2
A {> FeOH }T
+ k2
<
+
2
+

H
V
H

[ ]

[ ]

k Surf K S 1 k Surf K S 2

+
2
H+
H+

[ ]

(C17a)

[ ]

which yields
k Surf K S1 k Surf K S 2

K
K 2
A

log {> FeOH }T > log + log ko + k1 1+ + k 2


log
+
2
2
+
H+

H
V
H
H+

[ ]

[ ]

[ ]

[ ]

(C17b).

Now, we want to frame regions where the biotic reaction dominates. We have two
possibilities namely,

Rbio>RHom in the region where RHom>RSurf;


and, RBio>RSurf in the region where RSurf>RHom.
We start by addressing the first case, RHom>RSurf. This is done by comparing
Equations C14 and C16

K
K2
kbio C Bact H + > k o + k1 1+ + k 2
2

H
H+

[ ]

We formulate this equation as

[ ]

[ ]

[ ]

[ ]

[ ]

2
ko
kK
k K
H+ + 1 1 H+ + 2 2 H+
kbioC Bact
kbio C Bact
kbio C Bact

=0

and solve for [H+] by the Newton-Raphson method.


To assess the second case, we compare Equations C15 and C16

70

(C18)

(C18a)

[ ]

kbio C Bact H + >

A {> FeOH }T

k Surf K S 1 k Surf K S 2

+
2
H+
H+

[ ]

(C19)

[ ]

which gives
k Surf K S 1 k Surf K S 2

A
log {> FeOH }T < log k bio C Bact H + + log log
+
2
H+
V
H+

[ ])

[ ]

(C19a).

[ ]

The resulting predominance diagram is shown in Figures APP-C3 and 2-7.


Concentration of bacteria [g/l]
10
Microbial
kinetics
dominates

-4

-2

10

10

Heterogeneous
kinetics
dominates

2+

[Fe ]tot [M]

-1
10

-6

10

-2

-2

-3

-4

10
-4
-5

-6

Microbial
kinetics
dominates

-6

10

Homogeneous
kinetics
dominates

-7
1.5

2.5

3.5

4.5

5.5

6.5

7.5

pH

Figure APP-C3. Mechanism predominance diagram for the oxidation of ferrous iron. Lines
A were obtained from Equation C17b. Lines B were obtained by numerically solving
Equation C18a for [H+]. Lines C were obtained by Equation C19a. Constants used in the
model are listed in Table APP-C1.

71

Table APP-C1. Constants used to yield model results shown in Figures 2-7 and APPC3.
Constant
ko
k1
k2
ksurf
kbio

K1
K2
KK1
KK2
Ksa
Ka
Kb
KS1
KS2

Value
7.9 x 10-6
25
7.9 x 106
5.0
2.9 x 104

Source
Wehrli, 1990
Wehrli, 1990
Wehrli, 1990
Wehrli, 1990
Pesic et al., 1989

3.16 x 10-10
2.7 x 10-21
1.2 x 103
1.8 x 102
95
2.8 x 106
5.4 x 10-9
1.0 x 10-2
4.1 x 10-9

Nordstrm, 1990
Nordstrm, 1990
Nordstrm, 1990
Nordstrm, 1990
Nordstrm, 1990
Zhang et al., 1992
Zhang et al., 1992
Zhang et al., 1992
Zhang et al., 1992

72

Comment

Alternative formulation, see


Section 2.3.3
Corrected to I=0.27
---------

Appendix D. Geochemical database: Slow processes


Slow reaction stoichiometry, rate expressions, activation energies and rate constants and
parameters used in the model of Impoundment 1, Kristineberg (see Section 3).
Table APP-D1 The stoichiometry of some weathering and oxidation reactions.
Process
Reaction stoichiometry

(a)
(b)
(c)
(d)

(e)
(f)

Pyrite oxidation
(oxygen path)
Pyrite oxidation
(ferric iron path)
Ferrous iron
oxidation
Chalcopyrite
oxidation
(oxygen path)
Chalcopyrite
oxidation
(ferric iron path)
Sphalerite oxidation
(oxygen path)

(g)

Sphalerite oxidation
(ferric iron path)

(h)

Chlorite weathering
(chlinochlore)

(i)

Muscovite
weathering

7
FeS 2 ( s ) + H 2 O + O2 (aq ) Fe 2 + + 2SO42 + 2 H +
2
FeS 2 ( s ) + 14 Fe(OH ) 3 ( s ) + 26 H + 15 Fe 2 + + 2 SO42 + 34 H 2O
Fe 2 + +

5
1
H 2O + O2 (aq ) Fe(OH )3 ( s ) + 2 H +
2
4

CuFeS 2 ( s ) + 4O2 (aq ) Fe 2 + + Cu 2 + + 2SO42

CuFeS 2 ( s ) + 16 Fe(OH ) 3( s ) + 32 H +
17 Fe 2+ + Cu 2+ + 2SO42 + 40 H 2 O
ZnS ( s ) + 2O2 (aq ) Zn 2 + + SO42

ZnS ( s ) + 8Fe(OH ) 3( s ) + 16 H +
Zn 2+ + SO42 + 8Fe 2+ + 20 H 2 O
(Mg 4.5 Fe0II.2 Fe0III.2 Al )AlSi3O10 (OH )8 (s) + 775 H +
4.5Mg 2+ + 0.2 Fe 2+ + 0.2 Fe(OH ) 3 ( s ) + 2 Al 3+ + 3SiO2 ( s )
57
+
H 2O
5
KAl 3 Si3O10 (OH ) 2 ( s ) + 10 H +
K + + 3 Al + + 3SiO2 ( s ) + 6 H 2 O

73

Table APP-D2. Rate expressions for equations in Table D1, used in the model of
Impoundment 1, Kristineberg (see Section 3).
Process
Rate expression
0.11
Pyrite oxidation
(a)
[O2 (aq)]0.5
rate = hA py k pyo H +
(oxygen path)
Pyrite oxidation
3+ 0.62
(b)
rate
=
hA
k
Fe
py
pyf
(ferric iron path)

[ ]
[ ]

(c)

Ferrous iron
oxidation

+ k fe3 > FeOFe + + k fe 4 [Fe(OH ) 2 ])

Chalcopyrite
oxidation
(oxygen path)
Chalcopyrite
oxidation
(ferric iron path)
Sphalerite oxidation
(oxygen path)

rate = hAcp k cpo [O2 (aq)]

(g)

Sphalerite oxidation
(ferric iron path)

rate = hAsp k spf Fe 3+

(h)

Muscovite
weathering

rate = hAmu (k mu1 H +

(i)

Chlorite weathering
(chlinochlore)

(d)

(e)
(f)

rate = h [O2 (aq)](k fe1 Fe 2+ + k fe 2 FeOH +

0.5

rate = hAcp k cpf Fe 3+

0.43

rate = hAsp k spo [O2 (aq)]

0.5

rate = hAch (k ch1

0.58

[ ]
[H ]

0.40

+ 0.50

+ k mu 2 )

+ k ch 2 )

rate = qin cin


(j)
inflow
h height of the modelled box
Ai volumetric fraction of the ith mineral (i) x total specific surface area (As,tot) x specific
density
k rate constant
water content
q infiltration rate
References:
a) Williamson and Rimstidt (1994)
b) Rimstidt and Newcombe (1993)
c) Wehrli (1990)
d) modified from Scharer et al. (1994)
e) Rimstidt et al. (1994)
f) modified from Scharer et al. (1994)
g) Rimstidt et al. (1994)
h) Knauss and Wolery (1989)
i) Malmstrm et al. (1995)

74

Table APP-D3. Activation energies for reactions given in Table APP-D1, used for
temperature corrections of rate constants with Equation 2-5. Used in the model of
Impoundment 1, Kristineberg (see Section 3).
Process
Activation energy
Reference
-1
(kJ mol )
Pyrite oxidation
56.9
McKibben and Barnes, 1986
(a)
(oxygen path)
Pyrite oxidation
92
Wiersma and Rimstidt, 1984
(b)
(ferric iron path)
Average of values in Lowson,
(c)
Ferrous iron oxidation
70
1982
Chalcopyrite oxidation
20
Scharer et al., 1994
(d)
(oxygen path)
Chalcopyrite oxidation
63
Rimstidt et al., 1994
(e)
(ferric iron path)
Sphalerite oxidation
21
Scharer et al., 1994
(f)
(oxygen path)
Sphalerite oxidation
27
Rimstidt et al., 1994
(g)
(ferric iron path)
kmu1
22
Nagy, 1995
(h)
Muscovite weathering
kmu2
22
Analogy with Nagy, 1995
Analogy with biotite,
Chlorite weathering
kch1
30
White & Brantley, 1995
(chlinochlore)
(i)
kch2
30

75

Table APP-D4. Parameters in rate expressions in Table APP-D2, used in the model of
Impoundment 1, Kristineberg (see Section 3). Rate constants corrected for temperature of
1C. Rate given in units of mol dm-2 s-1; concentration in mol l-1 (=mol dm-3).
Parameter
Value
Units
-8
qin
7.7 x 10
dm s-1
cin
1 x 10-11
M
h
1
m
3 -3
0.15
mm

kpyo
8.7 x 10-12
mol0.61 dm-0.83 s-1
kpyf
8.1 x 10-10
mol0.38 dm-0.14 s-1
-6
kfe1
6.8 x 10
mol-1 dm4 m s-1
kfe2
2.11 x 101
mol-1 dm4 m s-1
kfe3
4.2
mol-1 dm4 m s-1
6
kfe4
6.8 x 10
mol-1 dm4 m s-1
-13
kcpo
4.0 x 10
mol0.5 dm-0.5 s-1
kcpf
1.2 x 10-11
mol0.57 dm-0.71 s-1
-14
kspo
5.4 x 10
mol0.5 dm-0.5 s-1
kspf
3.9 x 10-10
mol0.42 dm-0.26 s-1
-14
kmu1
2.9 x 10
mol0.6 dm-0.8 s-1
-16
kmu2
3.6 x 10
mol dm-2 s-1
kch1
2.6 x 10-13
mol0.5 dm-0.5 s-1
-15
kch2
4.6 x 10
mol dm-2 s-1
5
A*
2.48 x 10
m2 m-3
0.0384
m3m-3
py
0.0017
m3m-3
cp
0.0013
m3m-3
sp
0.144
m3m-3
mu
0.48
m3m-3
ch

*For the ith mineral, Ai= A x i, where i is the volumetric fraction of the mineral and
A=As,tot x specific density. Area measured by BET, As,tot = 0.1m2g-1;
specific density = 2.48 x 106 g m-3 (see Section 3.2.3)

76

Appendix E. Geochemical database: Equilibrium processes


Equilibrium processes and constants used in the model of Impoundment 1 (see Section 3).
Table APP-E1. Equilibrium processes and constants for the aqueous phase. Constants are
corrected for 1C, using the Vant Hoff equation and H values from the same reference as
given; the are also corrected for ionic strength = 0.27 with the Davies equation (see Section
2.1.1).
Aqueous phase
log K
Reference*
H2O - H+ = OH-14.6
N
H+ + SO42- = HSO41.16
N-W
Na+ + SO42- = NaSO40.05
N-W-M
K+ + SO42- = KSO40.13
N-W-M
Mg2+ + H2O- H+ = MgOH+
-13.1
M
2+
2Mg + SO4 = MgSO4
0.92
N-W
Ca2+ +H2O- H+ = CaOH+
-13.8
M
Ca2+ + SO42- = CaSO4
1.04
N-W
Ca2+ + H+ + SO42- = CaHSO4+
0.45
W
2+
+
+
Cu + H2O- H = CuOH
-8.3
W-M
Cu2+ + 2H2O- 2H+ = Cu(OH)2
-14.0
W-M
Cu2+ + 3H2O- 3H+ = Cu(OH)3
-26.9
W-M
2+
2Cu + SO4 = CuSO4
-1.16
W-M
Zn2+ +H2O- H+ = ZnOH+
-10.1
W-M
Zn2+ +2H2O- 2H+ = Zn(OH)2
-17.2
W-M
Zn2+ + SO42- = ZnSO4
1.13
W-M
Zn2+ + 2SO42- = Zn(SO4)222.01
W-M
Fe2+ + H2O- H+ = FeOH+
-10.6
N-W
Fe2+ + 2H2O- 2H+ = Fe(OH)2
-22.7
W-M
Fe2+ + SO42- = FeSO4
0.89
W-M
Fe2+ + H+ + SO42- = FeHSO4+
0.45
N-W
Fe3+ + H2O- H+ = FeOH2+
-3.45
N-W
Fe3+ + 2H2O- 2H+ = Fe(OH)2+
-7.64
N-W-M
3+
+
Fe + 3H2O- 3H = Fe(OH)3
-15.0
N-W
Fe3+ + 4H2O- 4H+ = Fe(OH)4-24.2
N-W-M
Fe3+ + SO42- = FeSO4+
2.05
N-W
3+
+
22+
Fe + H + SO4 = FeHSO4
1.53
N-W
Fe3+ + 2SO42- = Fe(SO4)22.77
N-W
Al3+ + H2O- H+ = AlOH2+
-6.31
N-W
Al3+ + 2H2O- 2H+ = Al(OH)2+
-12.7
N-W-M
3+
+
Al + 3H2O- 3H = Al(OH)3
-20.3
N-W
Al3+ + 4H2O- 4H+ = Al(OH)4-26.0
N-W
Al3+ + SO42- = AlSO4+
1.15
N-M
3+
+
22+
Al + H + SO4 = AlHSO4
-0.49
N-W
Al3+ + 2SO42- = Al(SO4)22.42
N-M
2H+ + CO32- = H2CO3
15.76
M
H+ + CO32- = HCO39.85
N-W
Ca2+ + H+ + CO32- = CaHCO3+
10.07
M
Mg2+ + H+ + CO32- = MgHCO3+
10.16
N-W
Fe2+ + H+ + CO32- = FeHCO3+
11.13
N-W
*References: N- Nordstrom et al. (1990)
W- (WATEQ) Ball and Nordstrom (1991)
M- (MINTEQ) Allison et al. (1991)
More than one reference indicates that the value was the same in all references named.

77

Table APP-E2. Solid species, gas phase and surface complexes used in the model of
Impoundment 1, Kristineberg (see Section 3). Notation > denotes a surface specie.
Reaction
log K
Reference*
Fe(OH)3(s) = Fe3+ + 3H2O- 3H+
5.76
W-M
0
SiO2(s) = H4SiO2 - 2H2O
-2.92
N-W
O2(g) = O2(aq)**
-3.69
W
CO2(aq)= 2H+ + CO32-#
-20.38
N-M
>FeOH+ + H+ = >FeOH2+
6.45
Z
>FeOH+ - H+ = >FeO-8.27
Z
>FeOH+ - H+ + Fe2+ = >FeOFe+
-1.99
Z
>FeOH+ - + H2O - 2H+ + Fe2+ = >FeOFeOH
-8.39
Z
*References: N- Nordstrom et al. (1990)
W- (WATEQ) Ball and Nordstrom (1991)
M- (MINTEQ) Allison et al. (1991)
Z- Zhang et al. (1992)
More than one reference indicates that the value was the same in all references named.

**[O2(aq)] = Po2 x KH where the partial pressure of oxygen at equilibrium with the aqueous
phase (Po2) was 0.1 atm in the base case, and the Henrys law constant, KH= 10-2.69 at 1C
(WATEQ - Ball and Nordstrom, 1991).
#

Similarly for CO2(aq) as for oxygen, where Pco2 was 10-3.5 atm (atmospheric pressure) in the
base case.

Table APP-E3. Processes tested in the model but not present in the base case; equilibrium
constants corrected for T=1C, I=0.27, rate constants corrected for T=1C.
Process
Constants
Reference
-6.91
Equilibrium with calcite
K=10
Nordstrom et al., 1990
CaCO3(s) Ca2+ + CO32Equilibrium with gibbsite
Allison et al., 1991
K=1011.1
+
3+
Al(OH)3(c) + 3H Al + 3H2O
K=10-3.41
Equilibrium with gypsum
Nordstrom et al., 1990
2+
2.
CaSO4 2H2O(s) Ca + SO4 + 2H2O
Equilibrium with siderite
K=10-9.57
Nordstrom et al., 1990
FeCO3(s) Fe2+ + CO32Calcite dissolution
k1= 6.6 x 10-1 Chou et al., 1989
k2= 1.1 x 10-4
CaCO3(s) Ca2+ + CO32k3= 5.2 x 10-7
rate = k1[H+] + k2[H2CO3] + k3

78

You might also like