You are on page 1of 36

Hydrometallurgy, 21 (1988) 155-190 155

Elsevier Science Publishers B.V., Amsterdam - - Printed in The Netherlands

T h e I n f l u e n c e of the E l e c t r o n i c S t r u c t u r e of Solids
on the A n o d i c D i s s o l u t i o n and L e a c h i n g of
Semiconducting Sulphide Minerals

F.K. C R U N D W E L L
Council for Mineral Technology (MINTEK), Private Bag X3015, Randburg2125 (South Africa)
(Received October 16, 1987; accepted in revised form March 12, 1988)

ABSTRACT

C rundwell, F.K., 1988. The influence of the electronic structure of solids on the anodic dissolution
and leaching of semiconducting sulphide minerals. Hydrometallurgy, 21: 155-190.

The energy level model of semiconductor electrochemistry is summarized, and is used to inter-
pret the literature of the leaching and the anodic electrochemistry of base-metal sulphides. It is
stressed that the electronic structure of the solid and the solution must be taken into account in
any fundamental interpretation of the anodic behaviour of semiconducting minerals. In semicon-
ductors with a band gap greater than i eV, anodic dissolution occurs almost entirely as a result of
hole injection into the valence band. In the case of chalcopyrite, which has a band gap of 0.6 eV,
both holes and electrons contribute to the dissolution. Fermi-level pinning is used to derive an
expression which is consistent with the observed dissolution kinetics of iron-containing sphaler-
ites, i.e., first order in iron content in the solid, and halt' order in oxidant concentration in the
solution. It is proposed that pyrite resists dissolution because its valence band is a crystal-field
band, which is non-bonding, so surface holes in the upper valence band cannot contribute to bond
breaking in the solid.

INTRODUCTION

The study of the electrochemical properties of base-metal sulphides has be-


come increasingly important in the development of extraction and beneficia-
tion processes of minerals. In particular, the stringent environmental
restrictions imposed on sulphide smelters have stimulated the development of
hydrometallurgical routes that avoid the production of sulphur dioxide. Such
routes typically involve the oxidation of sulphide ions to sulphur or sulphate
by means of an oxidant such as ferric ion or oxygen. This process, known as
oxidative dissolution, can be regarded as an electrochemical reaction involving
cathodic reduction of the oxidant and anodic oxidation of the sulphide. Many
dissolution reactions have been identified as electrochemical in nature, and

0304-386X/88/$03.50 © 1988 Elsevier Science Publishers B.V.


156

the mixed-potential model, which was developed to describe corrosion pro-


cesses, has recently gained acceptance in research on leaching and flotation.
The application of the mixed-potential model to leaching kinetics assumes
that the potential across the solid side of the interface (the space-charge re-
gion ) remains constant, so that the applied potential difference appears across
the Helmholtz layer. This is metal-like behaviour, and the dissolution reaction
is predicted to have a charge-transfer coefficient of 0.5 for an ideal single-
electron reaction. However, many sulphide minerals are semiconductors, and
the semiconductor model assumes that most of the applied potential drop oc-
curs across the space-charge region with a charge transfer coefficient of 1 for
the ideal case. Fundamentally different behaviour is therefore expected for the
dissolution kinetics described by the mixed-potential model and the semicon-
ductor model.
Solid-state properties have often been considered to influence the kinetics
of' dissolution and the flotation behaviour of semiconducting ore minerals.
Simkovich and Wagner [ 1 ] found that the dissolution kinetics of synthetically
grown stoichiometric, non-stoichiometric, and heavily Ag- or Bi-doped lead
sulphide varied with the concentration of point defects in the solid, and Ead-
ington and Prosser [2] reported a dependence of the oxidation rate on the
degree of non-stoichiometry of lead sulphide and, hence, on its electronic con-
duction type.
Richardson and O'Dell [3 ] used a combination of electrochemical, interfa-
cial-capacitance, and surface-photovoltage measurements in a study of oxygen
reduction and of xanthate adsorption onto lead sulphide. They concluded that
oxygen reduction is not dependent on the surface concentration of electrons,
while xanthate adsorption is dependent on the Fermi energy of the lead sul-
phide at the surface. Pawlek [4] proposed that the reduction of oxygen would
vary with the number of charged carriers, in support of which he found that
the irradiation of ZnS with UV light caused an appreciable increase in the rate
of dissolution of ZnS.
Springer [5] studied the anodic and cathodic polarization curves of n- and
p-type pyrite, n-type galena, and n-type chalcopyrite, and found no evidence
to suggest a rate-determining involvement of electronic charge carriers. He
concluded that the conduction properties of minerals do not normally play a
significant role in the dissolution reactions.
Attempts have been made to correlate the rates of dissolution with the solid
state properties of minerals. Pawlek [4] presented a table that listed the elec-
trical resistivity, the atomic structure and the rate of dissolution of a series of
minerals. He admitted that no correlation could be found, and added that little
was known of the atomic bonding structure of the minerals. Daiger and Ger-
lach [6] proposed that, since the oxidation of the sulphide occurs as a result
of the consumption of two holes, fast dissolution rates can be expected when
the concentration of holes is high, viz., in p-type semiconducting minerals.
157

Copper sulphides, as well as NiS, FeS and CuF%S3 belong to this group, whereas
sphalerite and chalcopyrite, which dissolve slowly, are n-type [6].
Peters [7] differs from Pawlek [4] and Daiger and Gerlach [5], proposing
instead that the differences in behaviour between sulphides can be accounted
for by changes in the molar volume of the solids when the metal dissolves and
solid sulphur forms. For example, pyrrhotite dissolves rapidly, while pyrite has
a slow rate of dissolution. Peters proposed that, since the molar volume of'
pyrite is 12 cm :~mol- 1 of contained sulphur and the molar volume of sulphur
is 16 cm :~mol- 1, there will be an increase in volume when the pyrite is oxidized,
and the sulphur layer will protect the pyrite from further dissolution. Peters
did not consider the differences in the structure and bonding of the minerals
in the formulation of his proposal.
Despite the lack of unanimity in the literature, a deeper understanding of
anodic-dissolution reactions can be gained from a consideration of the struc-
ture of the solid and the theory of semiconductor electrochemistry [8]. Most
research in semiconductor electrochemistry has been performed with high grade
synthetic crystals. Natural ores are often polycrystalline with a high density
of impurity levels. On the other hand, it is well known that ores of differing
origin may exhibit significant differences in leaching and flotation behaviour,
as a result of the differences in the solid-state properties of the minerals. It is
therefore of technological importance and academic interest to establish the
extent to which leaching and flotation behaviour is governed by the principles
of semiconductor electrochemistry.
This paper reviews the fundamental model of the electrochemistry of semi-
conductors, and then applies it to the dissolution kinetics of ore minerals. The
revised dissolution model that results takes into account the structure and
bonding of' solids as well as their electronic conduction properties. The revised
model is used to derive a mathematical expression for the leaching of iron-
containing sphalerites, and to interpret the anodic electrochemistry of other
binary sulphides. It is also shown that dissolution occurs mainly by a hole
mechanism if the band gap is greater than 1 eV. In the case of chalcopyrite,
which has a band gap of 0.6 eV, it is proposed that both holes and electrons
contribute to dissolution. Furthermore, it is proposed that pyrite resists dis-
solution because its valence band is of non-bonding character, and holes there-
fore do not contribute to bond breaking.

T HE ENERGY LEVEL MODEL OF SEMICONDUCTOR ELECTROCHEMISTRY

Garrett and Brattain [9] presented the first investigation of the semicon-
ductor-solution interface, which established the principles of semiconductor
electrochemistry. Three models have been proposed since for a semiconduc-
tor-solution interface: the idealized model [8], the recombinative model [10,
11 ], and the Fermi level pinning model [12]. Gerischer [13] and Green [14]
158

reviewed the semiconductor literature, and Morrison [ 15 ] and Memming [ 16 ]


presented the approach of semiconductor electrochemistry in great depth. Only
the more important aspects will be summarized here.

The band theory of semiconductors

The band structure of a semiconductor is characterized by the forbidden


band (the band gap) between the filled valence and empty conduction bands.
There are two types of charge carriers in a semiconductor: electrons in the
conduction band, and holes in the valence band. The probability of' any quan-
tum state of the bands being occupied by an electron or a hole is governed by
the Fermi-Dirac distribution law.
The band model for a semiconductor surface differs from that for a metal
primarily because of the excess charge density in the surface of the solid, called
the space-charge layer. The charge density at the surface increases as a result
of the termination of the lattice structure, resulting in Tamm or Shockley states
(intrinsic states ) at the surface. The excess charge density in the space-charge
region results in a "bending of the bands" at the interface. That is, the bands
in the bulk are lower than those at the surface if the applied potential differ-
ence is positive, and vice versa if the applied potential difference is negative.
The "one-electron" band model for an n-type semiconductor [17] is shown in
Fig. 1. The concentration of charge carriers at the surface is an exponential
/'unction of the potential difference between the surface and the bulk of the
semiconductor. The value of the applied potential difference when the bands
are flat is called the flat band potential.
The location of bands of compound semiconductors at the flatband potential
can be calculated using the method of Butler and Ginley [18], who assumed
that the Sanderson electronegativity, defined as the geometric mean of the
electronegativities of' the constituent atoms, was the midgap energy of the
semiconductor. In the present work, the band diagrams were calculated using
this method.
Most sulphides can occur as non-stoichiometric compounds, and this affects
the electronic properties of the solid. A metal-excess compound, such as chal-
copyrite, will display n-type semiconduction, while an anion rich compound
will display p-type semiconduction.

The energy level model of ions in solution

The representation of ions in solution as electronic energy levels is more


complicated than the representation of energy levels found in or on solids. The
energy levels of an ion in solution are distributed as a result of the thermal
fluctuations of the dipoles in the solvation sheath surrounding the ion. Theo-
ries of electrostatic polarization [19, 20] indicate that the distribution of en-
159

Space-charge
labor (;ou:,
,~ tl l ['{t t.'0 ~, U.|[ C~ = 10 000 .%. layer

Helmhohy I d Oflci/tcd dipolC ~,


laye~ = 3 A ~] P
Ionilcd <,tlll[icc: ~[{llos

1 jp o,<,

Fig. 1. T h e charge d i s t r i b u t i o n a n d energy level d i a g r a m at a n n-type s e m i c o n d u c t o r - s o l u t i o n


interface [ 17 ].

ergy levels is parabolic (Gaussian distribution). Figure 2 illustrates this model


for the ferricyanide and ferrocyanide ions. The probability of electron transfer
between an electrode and an ion in solution depends on the energy level of the
ion at the instant of transfer, and so these fluctuations are important [19].
The energy needed for the ion to reach the activated state is the difference in
energy between the oxidizing agent, Eox and E*, and the probability, W,>x,that
the energy level of a species will be at E*, is given by

W,,x. = ( 4 n ; . k T ) - ~ / ' ~ exp [ - ( E , > x - E ) ~ / 4*x k"~T ] " (I)


where ;. is the re-organization energy, which includes the energy contributions
from the stretching of bonds in the inner co-ordination sphere and the rotation
of' solvent molecules in the outer sphere [13, see also 21], k is the Boltzmann
constant, and T is the temperature. These fluctuations in energy are not small,
and the standard deviation ( ) . k T ) 1 / 2 , can be several tenths of an electron volt
[22]. Frese [23] has determined )~ for a number of oxidants and reductants.
The location of the energy levels of ions in solution can be obtained from
the standard reduction potential of the redox couple. E'r'~d .... equal to
1/2 ( E r e d + E o x ) , is related directly to the reduction potential tabulated in
handbooks [22]. The levels of the ions on the vacuum energy scale are shown
in Fig. 2.
160

VtlCtlIHll NHE
scale (cV) scale (\')
l']CC[ I'OIIS ill
I'OS[illtl ()T 4.44
tICUtlnl

4.44- _ l i ''H2
~('tl 2 +(tl '
F | e(('N)i ~, /I-c(( N)~
~rnE~ s ,I-"]C ~ - l.e2 -
/
~4t, ~)7T E~ s ~ ( ) , H,()
6- ('c ~ • (e~ ,

7-

-3
PbS /nS CdS IeS,

[-Bcrgy Icxcls at t h e
scnliconduclol SlllIttcc
I
('uI'cS, ] Rcdox
s) S[ClIlS

l)islribulion of energy
levels itl lhc rcdox c [ c c t r o ] y l c
Fig. 2. A comparison of the energy levels at the semiconductor surface and the energy,levels of
ions in solution. The left hand scale shows both the vacuum and the hydrogenenergy scales. (All
potentials in this work are referred to the NHE unless otherwise stated.) E,.~ and E,~ are the
conduction- and valence-band edges, respectively, and the 4te and 2e level refer to the iron d-
orbital levels in ZnS. pEae,refers to the thermodynamicdecomposition potential.

The semiconductor-solution interface

The semiconductor-solution interface consists of a double layer structure in


which the charge in the solid balances the charge in the near-surface region of
the solution. Charge concentrations can be tbund in the space-charge region,
the Helmholtz plane, and the Gouy layer, and these can be treated to a good
approximation as capacitors in series. A typical interface is illustrated in Fig.
1.
A semiconductor and a solution are in equilibrium when the electrochemical
potential of the electron in the solid is equal to that in the solution, i.e.,
Ev = E'r'~d.x. Thus, the immersion of a solid in a redox electrolyte solution re-
distributes any pre-existing charge in the interfacial region, so that an excess
of charge density in the solution is balanced by an equal density of charge in
the solid (the space-charge layer) at equilibrium [8, 13, 24]. This re-distri-
bution of the charge results in a shift of the flatband potential compared with
its value in vacuo.
For most semiconductors it has been observed that the flatband potential is
strongly dependent on pH, a shift of 0.059 V per pH unit being recorded, due
161

to adsorption of O H - or H +. Ginley and Butler [25] found a similar depen-


dence of the flatband potential of CdS electrodes on pH, and concluded that
the H S - and H ÷ ions were the potential determining species.

Charge transfer between a semiconductor and the solution

Electron transfer can occur only between states of the same energy level
[26 ], i.e., without energy transfer with the surroundings or radiation. In terms
of the fluctuating energy model, an ion must have a significant probability of
fluctuating to the energy level of the band with which electron transfer occurs.
For electron transfer to occur, E~edox must be sufficiently close to the conduc-
tion- or valence-band edges, and a comparison of Evs, Ecs, and Er~dox will give
a qualitative picture of whether electron transfer is possible or not [ 15, 22 ]. A
redox couple may be able to exchange carriers with only one of the bands. This
has been determined for germanium. Redox potentials with an E°odox > 0.5 V
are valence-band processes, while those with an Er°edox< 0 V are conduction-
band processes [27 ].
The rate of electron transfer from the semiconductor to the solution is pro-
portional to the number of occupied states in the semiconductor,
D~c (E)f(E-Esc), and the number of unoccupied states in the redox electro-
lyte, D,,x (E), at the same energy, E. Dsc(E) is the density of states in the semi-
conductor, f(E-Es~) is the Fermi function that describes how many of these
states are occupied, and Do~ (E) is the density of states in the oxidizing agent
in solution. The total current from the semiconductor to the electrolyte is the
integral of this probability over all possible energy levels [8, 13].
r~

j-=q j v-(E)D~(E)f(E-E~)Dox(E)dE (2)

where v - (E) is the transmission coefficient, and D,,~ (E) = CoxWox, where C,,x
is the concentration of oxidant in the solution.
The current-voltage behaviour of the ideal semiconductor-solution inter-
face has often been described as a Schottky barrier [15]. For a p-type semi-
conductor, the substitution of eqn. (1) into eqn. (2), and on integration gives
Jv =Jr,,, [exp (qV~/kT) - 1] (3)
where Jv is the current density of the valence band and Jv,o is the exchange
current-density. The characteristic feature of a Schottky barrier is that, in the
conduction band, only the cathodic process is influenced by the applied poten-
tial and, in the valence band, only the anodic process depends on the voltage.
The anodic process in n-type semiconductors will thus be limited by the dif-
fusion of minority carriers (holes) to the surface, and will display a saturation
162

current [15]. Therefore, anodic dissolution of n-type minerals will be slower


than that of p-type minerals.
The current-voltage curve for a single-electron reaction at an ideal semicon-
ductor surface, eqn. (3), implies a charge-transfer coefficient of 1 and a Tafel
slope of 0.059 V. However, when the number of surface states exceeds 10 l:~
states per cm ~, the Fermi level in the semiconductor is constant, and the change
of electrode potential occurs across the Helmholtz layer. In this case, the Fermi
level is pinned [12, 14, 28], the transfer coefficient is 0.5, and the Tafel slope
is 0.118 V decade-1, which is expected for metal-electrolyte charge transfer
behaviour. A similar Tafel slope is expected if the semiconductor is degenerate.
Morrison [ 15, p. 113 ] has discussed the transition between metallic and semi-
conductor behaviour.
Turner [29 ], during the anodic dissolution of Ge, obtained saturation effects
tor n-Ge, but none on p-Ge. Boddy [30] obtained Tafel slopes on n-Ge and p-
Ge that were very similar, except, that above 4 A / m 2, a slight tendency towards
saturation was observed for n-Ge. The observed Tafel slope ranged between
0.07 and 0.082 V decade -1, which is significantly higher than the expected
0.059 V. This was attributed to an appreciable change in the Helmholtz double
layer. Boddy [30] confirmed this directly using measurements of the interfa-
cial capacitance and the photovoltage response. Therefore, even for pure ele-
mental semiconductors, a deviation from ideal behaviour is found.

Anodic corrosion of semiconductors

The bonding between neighbouring atoms in a crystal is due to the interac-


tions in bonding orbitals of the neighbouring atomic nuclei. The bonding elec-
tron states belong to the valence band of the solid, and the presence of a hole
in this band means that one of the bonding electrons has been removed, and
the bond weakened. Holes can be supplied by the electrode interior as a result
of an electrical current or, in the case of corrosion and leaching, they can be
injected into the valence band by an oxidizing agent.
In semiconductors with a band gap greater than 1 eV, anodic dissolution
occurs almost entirely as a result of hole injection into the valence band [31].
Table 1 summarizes the contribution of holes to the anodic decomposition
reaction.
The oxidative decomposition of a compound semiconductor, M'~+S '~-, can
occur in the following steps:

(M'~+ S '~- )~ + X ~ ( M X ) + + "S~ + e - (4a)


(M'~+ S '~- )~ + X + h +-~ ( M X ) + + "S~ (4b)
•S ~ + Y ~ ( S Y ) + + e - (4c)
163

TABLE 1

C o n t r i b u t i o n of holes to t h e a n o d i c d i s s o l u t i o n of s e m i c o n d u c t o r s [31 ]

E~ Oxidative decomposition C o n t r i b u t i o n of
(eV) products holes ( % )

Ge 0.66 Ge ~+ 55 70
Si 1.09 Si ~* (Si 1+ ) ca 100
GaAs 1.35 G a :~+, As :~ ca 99.8
CdTe 1.5 Cd'2 +, T e l * 9
CdSe 1.74 C d '+ , Se ca 96
CuO 1.95 Cu'-' +, 1 / 2 0 ~ 9
CdO 2.2 Cd "-'+, 1/20.2 9
GaP 2.35 Ga :~+, P or P:~ ~ ca 100
CdS 2.4 Cd ~+ , S ca 100
ZnO 3.3 Zn ~+ , 1/20._, ca 100

• S~ + Y + h + --, (SY) + (4d)


• S~ + - S ~ (S-S).~ (4e)
S,i-{-• Ss ---~S,,_l_ 1 (4f)
where X and Y represent ionic species in solution. It can be predicted that the
electropositive component, M'~+, i.e., the element associated with the conduc-
tion band, will go into solution while, in many cases, the more electronegative
component will undergo a recombination reaction, which will result in the el-
emental state for this component.
Bard and Wrighton [32] and Gerischer [33] have proposed a thermody-
namic model for the anodic dissolution of semiconductors. They defined a re-
dox potential for anodic dissolution, pEdec, for the dissolution reaction (e.g.
CdS + 2HCI-, CdC12 + S + H,), pEdoc = A G / n F = 0.4 V ) and compared this with
the level of the redox couple in solution, or with the Fermi level of the solid.
The conditions for stability are
,,Ea,,,: > E'r'~a,,x or EF stable
pEd,,¢, < E'r'ed,,x or E F unstable
,E,tec can be regarded as a thermodynamic energy level for dissolution. The
position of pEa~c to Ec and Ev for various sulphide semiconductors is shown in
Fig. 2. Oxidants with a potential greater than p E d e c a r e on the basis of ther-
modynamics able to oxidize the mineral.
The general requirement for steady state corrosion is that the anodic and
cathodic currents are equal: j + = j - . The electrical field at the surface of a
metal has an opposite influence on these two processes, and it adjusts to fulfil
the current equality condition. Corrosion can be described by the superposition
164

of the anodic and cathodic currents. In semiconductors, the equality of these


currents means that
J~ +J~ --Jb- + J v (5)
and the net currents in each band will have opposite signs
j~ -js- =Jv -J~ (6)
In order to understand the dissolution process, it is necessary to consider
the crystal structure, the semiconduction type, and the characteristics of the
semiconductor-solution interphase.

D I S S O L U T I O N K I N E T I C S AND E L E C T R O C H E M I S T R Y OF S U L P H I D E S

The electrochemistry of metal sulphides has been reviewed by Koch [34]


and Hiskey and Wadsworth [35], while Wadsworth [36] and Dutrizac and
MacDonald [37] reviewed the leaching of metal sulphides. Burkin [38] has
reported on solid-state transformations during leaching, emphasizing the for-
mation of films that inhibit further reaction, and Peters [39] has discussed
the experimental techniques used in sulphide electrochemistry. Shuey [40]
reviewed the electronic properties of ore minerals, and Vaughan and Craig [41 ]
presented a volume on the mineral chemistry of sulphides.
Metal sulphides display a wide range of structural and electronic properties,
as shown in Table 2. The structural, electronic and bonding properties of each
mineral will now be discussed in detail, together with the relevant electrochem-
istry of each mineral, since the bonding properties of a mineral influence its
dissolution behaviour.

E L E C T R O C H E M I S T R Y OF C A D M I U M S U L P H I D E

The electrochemistry of CdS has been extensively studied because of its po-
tential use in liquid junction solar cells, but it has received very little attention
from hydrometallurgists.

TABLE2

Properties of selected base metal sulphides [35]

Mineral Formula Resistivity Semiconduction Rest potential


(~J m ) type (V)

Pyrite FeSe 0.1-3)< 10 e n, p 0.63


Chalcopyrite CuFeSe 0.2-9 × 10-:~ n 0.53
Chalcocite CueS 10-e-10 -~ p 0.44
Covellite CuS 10-4_10 6 metallic p 0.42
Galena PbS 10 -5 n, p 0.28
Sphalerite ZnS 10 ~- 107 - - 0.24
165

CdS is an n-type semiconductor with a band gap of 2.4 eV. The valence band
is associated with the sulphur anions and the conduction band with the cad-
mium cations. Defect-free single crystals of CdS display classic rectifying prop-
erties in which the cathodic current is exponentially dependent on the cathodic
potential, while the anodic current is limited by the supply of minority carriers
(holes). The anodic current is directly proportional to the intensity of illumi-
nation [42].
Tyagai and Kolbasov [43 ] and Van den Berghe et al. [44 ] have reported the
effects of surface states on CdS and CdSe. An ionic surface state associated
with the zero valent sulphur remains on the surface, available to be reduced.
Tyagai and Kolbasov observed that the electron transfer to ferric or ferricyan-
ide ions in solution is not linear in C,,x as expected from eqn. (2), but is given
by
j=k _(~'~7,,x,~P-°~r~dexp[ (1-o~)~le/kT] (7)
where 1 - c ~ = 0 . 7 . This was interpreted as the result of a surface state, the
charge of which changed with applied potential and the concentration of re-
ducing agent in solution. This represents a transition between ideal semicon-
ductor behaviour and metal-like behaviour. Van den Berghe et al. [44]
examined tunnelling through the surface barrier, and concluded that a surface
state energy level was present 0.12 eV below the conduction band [45].
Gerischer and Meyer [46 ] studied the corrosion of CdS crystals by tri-iodide
ions. The reduction of I :~- has a Tafel slope of 0.06-0.08 V decade -1, which is
consistent with the surface state results. The corrosion of CdS occurs at a slow
rate in neutral solutions, and the addition of Id- increases the rate of dissolu-
tion, although the I:~- does not overlap with the valence band. This is further
evidence of the surface-state mechanism.

T HE ELECTROCHEMISTRY AND DISSOLUTION OF GALENA

Galena is the principal ore mineral of lead, and its dissolution and electro-
chemical behaviour have been extensively studied. Brodie [46] studied the
anodic behaviour of galena electrodes, and Eadington and Prosser [ 2 ] reported
on the dissolution kinetics of natural and synthetic galena in acidic and alka-
line solutions. The influence of point defects on the rate of dissolution has
often been reported, and extensive research has shown that there is a close
correlation between the carrier concentration and the flotation behaviour of
galena [3, 48, 49].
Galena is a good conductor, displaying both n- and p-type semiconduction.
It has the halite structure, the internuclear distance being smaller than the
sum of the ionic radii, suggesting some covalency. The band structure associ-
ates the bottom of the conduction band with Pb 6p character, and the top of
the valence band with S 3p and Pb 6s character [40]. The band gap is about
166

0.37 eV [ 40 ]. Pure lead sulphide shows n- or p-type semiconduction, depending


on whether sulphur is deficient (n-type) or is in excess (p-type) compared to
the stoichiometric PbS. Common impurities in galena are Ag, Bi, Sb, and Cu.
Ag is an acceptor (p-type), while Bi and Sb are donors. Cu can act as a substi-
tutional acceptor or an interstitial donor.
Anodic electrochemical experiments [47] confirm that lead ions go into so-
lution, and it is suggested that sulphide ions form predominantly elemental
sulphur in acidic solutions, sulphate in neutral solutions, and thiosulphate in
alkaline solutions [2 ].
Eadington and Prosser [2], who investigated the influence of non-stoichi-
ometry on the dissolution kinetics of synthetic lead sulphide, reported that the
rate of dissolution after an initial inhibition period is much higher for p-type
samples than for n-type samples (Fig. 3), indicating a rate-determining in-
volvement of holes. They proposed that, during the induction period chemi-
sorption of oxygen, associated with a change in surface properties, takes place,
and the induction period ends with the nucleation of the first elemental sul-
phur. The dissolution mechanism proposed is the anodic oxidation of PbS and
the cathodic reduction of oxygen in a mixed-potential model. In subsequent
work, Eadington [50] observed that no inhibition period existed for the dis-
solution of natural and synthetic samples in nitric acid. Illumination had a
greater influence on the dissolution rate of the n-type sample than on the p-
type sample. The influence of point defects was studied by Simkovich and
Wagner [ 1 ], and their results support the work of Eadington and Prosser [ 2 ].
Brodie [47] investigated the anodic electrochemistry of museum grade nat-
ural galena. The anodic Tafel slope in acid perchlorate solution was 0.057 V
decade-~ - which is close to the 0.059 V decade- xexpected for a semiconductor
electrode - for one of his samples, and 0.076 V decade- 1 _ which is consistent

I I I
.E 1.2-

E 1.0-
©
~ 0.8-

>~ E 0.6-
"E

E Excess Pb / Excess S
" ~ , u~ 0.4-- -/ '-- p-type --1
" 0.2-

m 0 ! I I --
12 8 4 0 4
Non-stoichiometry, arbitrary units
Fig. 3. Effect of non-stoichiometry on the induction period and the steady rate of oxidation [2].
167

with the presence of surface states - for the other. Brodie reported that at 0
pH the production of sulphur occurs up to a potential of 0.9 V, above which the
production of sulphate is detected. The formation of a PbS04 film passivates
the electrode at these potentials.
Richardson and O'Dell [3] studied the semiconducting characteristics of
pure natural galena and synthetic lead sulphide, using electrochemical, inter-
facial-capacitance and surface-photovoltage (SPV) measurements to charac-
terise the electrodes. Variations in capacitance and SPV occurred over a
potential range greater than the band gap, indicating that changes in the
Helmholtz potential cannot be neglected. The anodic Tafel slopes ranged from
0.06 to 0.08 V decade-I, which is consistent with the surface-state semicon-
ductor model. Further changes in the electrode potential in the anodic direc-
tion presumably occur across the Helmholtz layer, and metal-like Tafel
behaviour is expected.
Paul et al. [51], using cyclic voltammetry, examined the anodic behaviour
of natural crystals, and reported that the electrode passivates at potential
greater than 0.6 V vs SCE as a result of the formation of a basic lead sulphate
on the galena surface. The Tafel slopes, which were dependent on the potential
and the concentration of lead in solution, ranged between 0.045 and 0.12 V
decade- 1. Paul et al. [51 ] concluded that these results were not consistent with
a conventional two-electron transfer mechanism, and proposed that dissolu-
tion occurred in two single-electron transfer steps in which an ' S - ' interme-
diate was formed. On the basis of this mechanism, they derived equations to
describe their results.
Although this mechanism is not incompatible with their results [51], it is
incompatible with the mechanism of Richardson and O'Dell [ 3 ], who proposed
that the change in Tafel slope at high potentials is a result of Fermi-level pin-
ning, and that a metal-like Tafel slope (0.12 V decade -~ ) is expected. A slope
of 0.084 V decade- 1 is the result of a significant potential drop over the Helm-
holtz layer increasing the Tafel slope from the expected 0.059 V decade-1.
Adsorption of potential-determining species on the surface would result in a
linear dependence of d VE/dlog (Pb 2+ ) with a slope of - 0.03 V decade- 1, which
is observed with Paul's data at low current density. Studies of the structure of
the double layer of galena as a function of lead concentration using the meth-
ods of Boddy [29], for example, might clarify this aspect. The adsorption of
copper ions which is important in flotation, should also be investigated using
semiconductor techniques.
Paul et al. [51] reported that illumination did increase the rate of anodic
dissolution by a factor of 3 to 7 if the electrode was held at a potential of be-
tween 0.2 and 0.35 V vs. SCE, which corresponds to a small peak before the
main anodic peak for the bulk oxidation of the electrode. Since no large photo-
effect was observed on the main anodic peak, they concluded that semicon-
ducting properties were not an important factor in dissolution kinetics. Their
168

results, interpreted in terms of the semiconductor model, suggest that, at po-


tentials more anodic than the small peak at 0.3 V vs. SCE, the Fermi level is
pinned at the valence-band edge, and if this is the case the SPV would be
saturated and no photo-effect would be observed [3].
In the presence of oxygen, the anodic Tafel slopes range between 0.18 and
0.22 V decade -1 [3]. Richardson and O'Dell attributed this to the formation
of a PbO film on the surface, such that the applied potential was distributed
between the space-charge layer, the PbO film, and the Helmholtz layer.

T H E DISSOLUTION OF IRON-CONTAINING S P H A L E R I T E S

The rate of oxidative dissolution of iron-containing sphalerites, (Zn,Fe)S,


is known to be higher than that of pure ZnS [52], and the rate of reaction has
been reported to be a linear function of the concentration of iron in the zinc
sulphide lattice [53]. Some data of Piao and Tozawa [53] are reproduced in
Fig. 4. Jin and Warren [54] observed a half-order dependence of the rate on
the concentration of Fe(III) in chloride solution, which is consistent with a
mixed-potential model and metal-like conduction. Analogous results have been
reported for sulphate solutions [55]. The results of Jin and Warren, on the
other hand, indicate that the anodic dissolution of sphalerite reaches a limiting
current, as expected for an n-type semiconductor (Fig. 5). Below we derive an
equation based on the semiconductor model that describes the effects of the
iron content in (Zn,Fe)S and oxidant concent ration on the rate of dissolution.

76 - Anvil

i~ 5-
E
~ 4- eUchinota
x
~ 23 - /Kamioka

I-- " San Vicente

0 1 [ I I I I ! I ~ -~
l 2 3 4 5 6 7 8 9 l0
Fe in sphalerite (%)
Fig. 4. The effect of substitutional iron in sphalerite on the rate of dissolution using oxygen as an
oxidant [53 ].
169

2.1.-

ii
-2.3-

-2.5'-

-" -2.7-
.E
E
.at

vd~
2.9-

3.1-
: Part, le ,ze RT,on

3.3-

3.5 ¸
I I I I I
1.4 1.0 -0.6 0.2 0.2 0.6
log,{+ [Fe + + ] t o t a l
Fig. 5. T h e order of r e a c t i o n w i t h r e s p e c t to total ferric ion c o n c e n t r a t i o n for ferric chloride leach-
ing of a n a t u r a l Z n S c o n c e n t r a t e [54].

Structure of (Zn, Fe)S

Zinc sulphide is a semiconductor of intermediate ionic character, and has a


band gap in the range 3.6 to 3.9 eV [41]. Literature data indicate that the
bottom of the conduction band is derived from the Zn 4s orbital, while the top
of the valence band is derived primarily from the S 3p orbital [40]. This is in
qualitative agreement with an ionic model, which associates the valence band
with the anion, S, and the conduction band with the cation, Zn [56].
Zinc sulphide is such a poor electrical conductor that it is often classified as
an insulator. Resistivities have been reported in the range 107 to 109 ~ cm.
Telkes [57] obtained high positive thermopowers for four samples, indicating
p-type conduction. However, this result may have been due to the presence of
iron impurity, giving rise to conduction in a d-orbital band [58].
The substitution of transition metal impurities for Zn in ZnS has a pro-
nounced effect on the electrical properties of ZnS. Measurements of the elec-
trical conductivity of a natural sphalerite containing 12.4 atomic % Fe showed
behaviour typical of a semiconductor, with an activation energy for conduction
170

of' 0.49 eV, although the electrical resistivity remained high at 8 × 107 t2 cm
[ 58 ]. The absence of a measurable Hall effect and a positive thermopower [ 57 ]
suggested the hopping of holes in a localised d-orbital band as the conduction
mechanism [41, 58]. This model requires the holes to be generated as Fe :~+
ions, although only one in 800 Fe atoms need be Fe :~+ in order to account for
the conductivity [58]. Quantum calculations indicate that the 2e and 4re or-
bital levels are about 0.56 eV and 1.44 eV above the valence band respectively
[ 59 ]. The 2e and 4re orbitals can be represented as localized bands within the
ZnS band gap, and the removal of electrons from either of these orbitals will
create the holes required for the d-orbital conduction mechanism.

Charge transfer to a narrow band and the kinetics of dissolution of (Zn, Fe)S

The presence of an iron d-orbital within the band gap has two consequences
%r electron transfer between the solid and the solution: (i) it creates a narrow
localized band that is energetically more favourable for the transfer of elec-
trons than the valence band, and (ii) it "pins" the Fermi level at a level within
the d-orbital band [ 12, 20, 60, 61 ].
Holes may be injected into the d-orbital band by a suitable redox couple in
solution, and if the space charge region is sufficiently narrow ( ~ 2 nm), tun-
nelling will occur between the surface d-orbital and the valence band. The
capture of two holes by the valence band at the surface will then result in the
oxidation of S 2- to S O and the concomitant release of Zn 2+, associated with
the conduction band, into solution [34, 62].
Because of the narrowness of the d-orbital band, the density of states for the
carriers in the solid resembles a Dirac-delta thnction of population located at
the energy of the band [24]. The transmission coefficient v - (E) also has the
form of a Dirac-delta function and, since D,,x (E) = Cox W,,x (E), eqn. (2) can be
approximated as
j ~-qzN~lC,,xW,,x(E*)kT (8)
Another consequence of the iron impurity is that the Fermi level is pinned
at the energy level of the d-orbital. Hence, all the applied potential appears
across the Helmholtz layer, i.e., E* =EF, and E,,x-EF varies directly with ap-
plied voltage:
E,,x - E * = E,,x - Ev = E,,x - Er.,, + rlq (9)
Substitution of this expression into eqn. ( 1 ), and combination with eqn. (8),
gives the current as
j - =q z ( k T /47r2 ) 1/2 C,,xNd e x p [ - (Eox --EF.,, + ~lq)2/42kT] (10)
Since E . x - E r e d = 2J[ (see Fig. 2) and EF,,,= l/2(E,,x+Ere(i) +kTln(Cred/C,,x)
then
171

E,,x -E~.,, =,~-kTln(CreJC,,x) (11)


and
j =qz(kT/4~)L)I/2NaCoxexp[- (}~-kTln(CreJC,,x)+~lq)2/4,~kT] (12)
Applying the assumptions of Levich [20] and Marcus [19] that ).>>q~land
).>>k Tin (Cr~d/C,,~), the cross terms in q~l and kT in (C,.~t/C,,x) can be neglected.
The resulting equation describes the reduction of Fe :~+ at the sphalerite surface:
j = k 1C,,~Naexp( - ~lq/2kT) (13)
where k_ 1 is a rate constant. The anodic dissolution current is also dependent
on the number of point defects [ 1 ], and is given by
j + = k l N a exp(~lq/2kT) (14)
Since j + = j - , eqns. (13) and (14) can be equated, and an expression for ~/
obtained. Solving f o r j ÷ gives
j + = k l N a (Co×kl/kl) ~/'~ (15)
This equation, derived from fundamental principles, describes the results of
Figs. 4 and 5. The dissolution rate is first order in Na, the number of occupied
states in the d-orbital band, i.e., the concentration of Fe impurity in the
(Zn,Fe) S, and half order in the concentration of oxidant. From the energy level
model of semiconductor electrode processes, it is therefore expected that the
rate will be enhanced by the introduction of a stronger oxidizing agent, i.e., a
higher E .... or by ultraviolet irradiation, which will create free electrones and
holes in the valence band.
Figure 6 illustrates the results of Exner et al. [52], and shows that irradia-
tion does indeed increase the rate of dissolution of both natural and synthetic
ZnS. The rate of dissolution of the synthetic crystal is lower than that of the
natural crystal, suggesting that Al-doping, which is n-type, is more important
than Cu-doping.

THE DISSOLUTION OF PYRITE

Hiskey and Schlitt [63] recently reviewed the aqueous oxidation of pyrite.
Warren [64] reported the first detailed kinetic study, and observed an acti-
vation energy of 83.7 kJ mol-1 for the oxygen pressure leaching of pyrite be-
tween 130 and 210 ° C. The rate of reaction was proportional to the square root
of' the oxygen partial pressure, which is consistent with the mixed potential
model. The activation energy was reported to range between 58.6 and 51.1 kJ
tool-~ for pressure leaching below 130 ° C.
Anodic dissolution of pyrite at 1.05 V is thought to proceed according to the
fbllowing stoichiometry [65]:
172

20-
• Synthetic ZnS (Cu, AI) ) k
• Synthetic ZnS (Cu, AI) and UV li~h t/
"o Natural ZnS
15- a N a t u r a l ~ Y n < ~ , a ~ T V m ~ , ~ "

{
v 10-

?
5-

0 r I I i
0 30 60 90 120 150
Time (rain)
Fig. 6. The effect of crystal doping and UV radiation on the dissolution of natural and synthetic
ZnS [521.

FeS,2 + 8H,eO ~ Fe a+ + 2S0] - + 16H + + 15e- ( 16 )


with sulphate as the principal sulphur product; although Dutrizac and
MacDonald [66] reported that sulphur and sulphate were produced concur-
rently by separate but related processes. Nevertheless, the oxidation of sulphur
to sulphate, although thermodynamically possible, must be extremely slow.
Sulphur is a stable product in the oxidative dissolution mechanism [66].
An important finding of the work of Bailey and Peters [67] was that the
sulphate tormed from pyritic sulphur contains oxygen acquired from water
instead of from oxygen gas. Evidence in support of this hypothesis was ob-
tained from 1sO tracer experiments. In addition, detailed stoichiometric deter-
minations showed that the sulphate yield varied between 60 and 90%, and
increased with increasing oxidation potential.
Biegler and Swift [68] proposed a mechanism involving the electrochemical
transfer of the oxygen in the water molecule to the sulphur atoms in the pyrite
lattice. The mechanism they proposed is

FeS~ + H20=FeS~(OH)a,t~ + H + + e (17)


FeS2 (OH)~d~ -~FeS2 (O)ad~ + H + + e - (18)
where the second step is rate-limiting. They used the Temkin isotherm, and
derived an equation that predicted a Tafel slope of 0.059 V decade- and a
dependence of -0.03 V per pH unit for transfer coefficients of 0.5 for both
173

reactions. Biegler and Swift [68 ] measured a Tafel slope of 0.095 V and Meyer
[69] reported a dependence of - 0 . 0 3 V per pH unit.
Pyrite is generally regarded as displaying noble electrochemistry. It has a
high open-circuit potential and is galvanically protected from corrosion when
in contact with other minerals. Parker et al. [70] presented current-voltage
curves that were very similar to those of platinum.

Structure of FeS2

Pyrite may have both n- and p-type conductivity, with the p-type having the
higher resistivity ( ~ 1 × 10 -3 ~2- m for n-type and ~ 3 × 10 -2 t~-m for p-type )
[40 ]. Pyrite has a cubic structure, represented as Fe 2+ ($2) 2- by the ionic model.
The lower part of the conduction band is derived from the anti-bonding e~
doublet of the 3d orbitals on Fe, while the upper part is derived from the 4s
orbital. The valence band is derived from the t2~ triplet of the 3d orbitals on
Fe, and, most importantly, remains non-bonding [40, 41, 71 ]. The energy sep-
aration between the non-bonding t2~ and the anti-bonding e~ orbitals is there-
fore the band gap. Bither et al. [ 72 ] measured the optical reflectivity of pyrite,
and inferred a band gap of 0.9 + 0.1 eV. Since the sulphur atoms are in octa-
hedral co-ordination about the iron, the overlap between orbitals is slight, and
the valence band is narrow, about I eV. The band derived from the sulphur 3p
levels is about 7 eV wide, and is just below the t2~ valence band. The 'one-
electron' energy band diagram for pyrite is shown in Fig. 7.

The influence of the structure of pyrite on the dissolution kinetics

An important difference between transition metal disulphides, such as py-


rite and molybdenite, and other polar semiconductors, such as CdS, ZnO, and
GaP, is that the valence band is narrow and of non-bonding character. Elec-
tron transfer with this band (hole injection) will therefore have no effect on
the structure or bonding of the solid, and corrosion will not occur. This feature
was cleverly exploited by Tributsch and co-workers [ 73-76 ] in their search for
a suitable material for the conversion of solar energy. Holes in the t2g band of
pyrite and the d~2 band of molybdenite are capable of oxidizing water, and
subsequently of producing oxygen as a fuel (splitting of water). Tributsch [73 ]
reported that the production of sulphate was the predominant long term photo-
anodic product of MoS2, according to
18hp

MoS2 + 8 H 2 0 + 18h + -~ Mo(VI) +2SO~4 - + 16H + (19)


where h u refers to the energy of the incident photon. He also proposed a mech-
anism for the oxygen-liberation reaction, and a scheme for the choice of suit-
174

t
4p ~ ~ -tO* Mo
i

1.75e~
H+/H2

Fe3-/Fe 2+

__ !i , Ce4*/Ce TM

~
~ s' p 3 dO

_ ["_/f_/4..
I',,\\\\\xl
Fe-'* ~ / S3p/
alomic
orbitals 2;b2
Pyrite, FeS: Molybdenite, MoS2
Fig. 7. The band diagram for pyrite and molybdenite. The pyrite diagram shows a simplified
molecular orbital diagram indicating the origin of the non-bonding valence band, the t.e~ band.
The crystal-field splitting of the Mo 4d orbitals in MoS~ is also shown. The 4d~2 band of MoS~ is
a non-bonding valence band.

able transition metal disulphides for photo-anodic devices [73]. There seems
to be little doubt that the non-bonding character of the valence band is the
reason for the slow dissolution behaviour and the reversible current-voltage
behaviour of both pyrite and molybdenite.
The formation of a hypothetical sulphur layer capable preventing further
oxidation is not consistent with the observed Tafel slopes, and Peters' [7]
proposal that dense sulphur layers are the reason for the slow rates of oxidation
of pyrite must be rejected in favour of an explanation based on the band struc-
ture of pyrite.
Recently Tributsch and co-workers [75, 76] examined the photo-electro-
chemistry of natural and synthetic pyrites. The electrochemical behaviour of
the natural samples was quasi-metallic, and a Schottky barrier was not ob-
served, indicating the high level of doping of the natural crystal. Anodic photo-
effects were observed for all samples, however. Jaegermann and Tributsch [ 75 ]
concluded that surface states within the band gap caused the applied potential
to appear mainly across the Helmholtz layer, and subsequently concentrated
their attention on RuS,,, which also has a similar band structure to pyrite, but
produces photo-anodic oxygen, in contradistinction to pyrite which corrodes
to sulphate and Fe(III).
A possible mechanism tbr the oxidation of FeS~ is shown in Fig. 8. Holes are
175

e~* Fe d
E<

h ~
H20/O2~ Ce4 + ~ ( ' e ~ +t3"
(Te ~ , / C o 4 + ~ E,
H20+h* ~H-+'OH
h+
t2~ Fc d

\ h +

D -'--"-OH~OH 4 I1+
.OH/OH S 3p

Fig. 8. A simplified scheme for the dissolution of pyrite by an oxidant, neglecting the effects of
surface states.

initially injected into the t2g valence band by the oxidizing agent. These holes
do not result in corrosion but are able to split water according to the reaction
H20+h+-~H+ +.OH (20)
where • OH is an intermediate hydroxide radical. The energy level of the - O H /
O H - couple has been estimated to be 2.7 V vs. SCE by Memming [77] and is
shown in Fig. 8. The • OH radical is a strong oxidizing agent and reacts directly
with the S valence band to form sulphate. Oxygen from the water molecule
becomes incorporated directly in the sulphate ion. The • OH radical can oxidize
the S valence band directly to produce sulphur. The iron enters the solution as
Fe :~+ and therefore consumes an extra hole.
Tributsch [ 73 ] presented a similar scheme in which O H - is chemisorbed by
the oxidized metal atom. The rate-determining steps in this model are the same
as those considered by Biegler and Swift [68]. The measured Tafel slope is
0.095 V decade-1 and the predicted slope is 0.059 V decade-1 for a variation
in the Helmholtz potential, indicating that the applied potential appears across
the space-charge layer and the Helmholtz layer. A Tafel slope of 0.095 V
decade- ~ is typical of the transition between a semiconductor Tafel slope of
0.059 V decade -1 and a metal-like Tafel slope of 0.118 V d e c a d e - L Similar
behaviour for germanium was shown by Boddy [78] to be a result of the sig-
nificant change in the potential drop across the Helmholtz layer in addition to
the potential drop across the space charge layer. Specific adsorption at the
inner Helmholtz plane by chloride ions, or an increase in dislocation defects,
would increase the Tafel slope to 0.118 V.
Biegler and Swift [68] suggested that the sulphur and sulphate pathways
have very similar activation energies or the same rate-limiting step. This is not
176

TABLE 3

Leaching of MoS2in various leaching systems [80]

Redox couple E~,.,~.,x Dissolution rate


(V vs. SCE)
Fe:~+ 0.77 no reaction
MnO~ 0.8 to 1.0 slow reaction
HNO:~ 0.95 to 1.0 slow reaction
KMn04 1.2 moderate reaction
Acid chlorate 1.2 moderate reaction
Hypochlorite 1.4 fast reaction

inconsistent with the above mechanism if the rate-limiting step is either the
injection of holes into the tzg valence band or the splitting of water. Biegler
[79] reported that there was no detectable correlation between the electro-
chemical reduction of oxygen at a pyrite electrode and the semiconducting
properties of various pyrite samples. This suggests that the splitting of water
is a feasible rate-determining step.
Although the open-circuit potential of pyrite is 0.62 V, significant corrosion
does not occur until the redox potential is about 1.0 V. Apparently, energy
levels below 1.0 V (Fig. 7 ) do not overlap sufficiently with the t2g band to inject
holes. Thus, the greater the electrode potential, the greater the rate of electron
transfer, and the greater the rate of dissolution.
The dissolution rate of molybdenite is slow with oxidizing agents having an
E'r~dox value of less than 1.0 V [80]. Table 3 summarizes the leaching rate in
various leaching systems. Krishnasway et al. [80 ] studied the electrochemistry
of this mineral and interpreted the results in terms of n-type conductivity.
The dissolution rate of arsenopyrite is also expected to be slow since it too
has the pyrite structure. Indeed, arsenopyrite does fall into the category of
minerals that are "difficult to dissolve".

THE DISSOLUTION OF CHALCOPYRITE

The anodic dissolution of chalcopyrite is thought to occur according to [81 ]:


CuFeS2 ~ Cu 2+ + Fe 3+ + 2S + 5 e - (21 )
and
CuFeS2 + 8 H 2 0 - ~ C u 2+ + F e 3+ + 2 S O ] - + 16H + + 17e- (22)
Biegler and Swift [81] found that 86% of the sulphide sulphur reports as
elemental sulphur, and 14% as sulphate. Ferric chloride leaching, which fol-
lows linear kinetics, is 4 to 5 times more rapid at 80-90 °C than ferric sulphate
leaching, which follows parabolic kinetics. Chloride solutions yield sulphur al-
177

most exclusively, whereas sulphate solutions produce between 14 and 25%


sulphate.
Dutrizac [82 ] showed that, in chloride solutions, dissolution followed linear
kinetics, and that the rate was proportional to [ Fe :~+ ]~):~4.The rate in sulphate
solutions was proportional to [Fe :~+ ]o.12,and followed parabolic kinetics. Jones
and Peters [83] showed that in sulphate solutions, corrosive attack was highly
selective, in as much as it only occurred in fissures or at grain boundaries,
whereas, in chloride solutions, attack occurred over the entire surface. The
addition of ferrous chloride had no effect on the rate of leaching, yet the ad-
dition of cupric chloride increased the rate. This was attributed to Cu ~+ being
the primary oxidant.
Many researchers have proposed that the oxidation products are resistive.
Sulphur [84], p- or n-type semiconductors such as p-type CuS, and metal-
deficient polysulphides [81, 85 ] have all been proposed as hindering the trans-
fer of electrons or ions at the solution interface.

Structure of CuFeS~

Chalcopyrite is a good conductor with a resistivity of about 10- :~t2.m [40].


It is an n-type semiconductor, due to metal excess, with a band gap of about
0.6 eV. The ionic bonding model of chalcopyrite is represented as
Cu+Fe :~+ (S ~- )~. Apparently, there is not extensive direct covalence among
the anions, but there is considerable covalence between the cations and anions
[40]. Unfortunately, precise computation of the energy bands in chalcopyrite
is difficult because of its antiferromagnetism. However, studies of other I - I I I -
VI compounds, in particular Fe-doped CuGaS2, suggest that the lower part of
the conduction band of chalcopyrite is Fe 3d in character, while the upper part
of the valence band is Cu 3d and S 3p in character [86, 87]. The relevant "one-
electron" band diagram is shown in Fig. 9.

Application of the energy level model to the corrosion of CuFeS~

Parker et al. [70] conducted extensive studies of the electrochemistry of


chalcopyrite in chloride solutions in the temperature range 70 to 90 ° C and the
potential range 0.2 to 0.6 V (SCE), which is appropriate for chalcopyrite leach-
ing. They reported that the rates of reduction of oxidants were in the order
Bre, CuC12, I:7 > Fe ( CN ) ~- > FeCI:~ > F% ( SO~ ):~, CI,,
The location of the bands at the flatband potential and the location of the
redox potentials of these oxidants are shown in Fig. 9. In order for electron
capture or hole injection to occur, the energy level of the oxidant must suffi-
ciently overlap the conduction-band edge, or the valence-band edge, respec-
178

Conduction Cu 2 ~/Cu
band
Fe 3d
E(
Fe(CN)~'/Fe(CN)~
If/1

0.6 eV
Fe 3 • / F e 2 +

Ev NOS/NO
Br2/Br

Cu 3d
S 3p

Upper 02/H20
valence
band

Fig. 9. The energy level diagram for CuFeS.2, indicating the Cu 3d contribution to the valence
band, and the Fe 3d contribution to the conduction band.

tively. It is apparent from this diagram that the rate of reduction of CuC12 and
I:~ is high because the energy levels of these ions overlap the conduction-band
edge, and the rate of reduction of Br2 is high because the bromine energy level
overlaps the valence-band edge. In contrast, the iron chloride and iron sulphate
energy levels are well within the band gap, and the rate of reduction is low.
Fe ( C N ) ~ - is reported by Parker et al. [70] as having a lower rate of reduction
than I:7, which has a higher Er°dox, but this is because they compared these
rates at different temperatures. The rate of reduction of Fe ( C N ) ~ - was deter-
mined at 25 ° C, while the others were all determined at 80 ° C. The level of CI~
is well below the valence-band edge.
Parker et al. [ 70 ] also reported anodic saturation currents, which are typical
for n-type semiconductors where the supply of holes is migration-limited, for
the oxidation of Fe 2+ in solution. Similarly, Beckstead et al. [88] presented
results showing that the rate of dissolution increased with Fe 3÷ and then
reached a steady value, indicative of saturation currents. Saturation currents
are not found for I2 or CuCI~, however, since the energy levels of these ions
overlap the conduction band edge. From these arguments, it may be concluded
that the interaction of I:T/I-, CuC12/CuCI~- and F e ( C N ) ~ - / F e ( C N ) ~ - are
179

conduction band processes, while Fe 3+/Fe 2+ and B r 2 / B r - are valence band


processes.
Zevgolis and Cooke [89] presented results suggesting an anodic limiting-
current due to migration of holes (minorities) in the n-type chalcopyrite. They
illuminated the rear of the sample and observed that the current density was
proportional to the light intensity.
Subsequently, they have been severely critisised for illuminating the rear of
their sample, and for the fact that the photo-effect occurred only at high ( > 3
V) anodic potentials [ 79 ].

Mechanism of dissolution of chalcopyrite

An interesting observation of Parker et al. [70] and others [83] is that the
addition of CuC12 to a FeCl:~ solution results in enhanced current densities at
a chalcopyrite electrode.
Table 4 presents a selection of data from Parker et al. [ 70 ], which indicates
that the anodic dissolution of chalcopyrite occurs with a contribution of holes
in the valence band and electrons in the conduction band. This is the case for
many semiconductors, such as Ge, with a band gap of less than 1 eV [28]. Since
the dissolution current for the contribution of holes from oxidant Fe 3+ is greater
than for the contribution of electrons to Cu 2+ in solution, it is assumed that
the initial step in the dissolution of chalcopyrite is the consumption of a hole.
Chalcopyritic Fe 3+ is associated with the conduction band and chalcopyritic
Cu + with the valence band; the electropositive Fe 3+ will therefore be released
into solution, and the Cu + will remain after the initial step. The initial hole
transfer is followed by electron transfer, according to
CuFeS2 + 3h + --, Fe '~+ + - CuS2 (23)
followed by hole transfer
• CuS2 + 2h + -~ Cu 2+ + 2S (24)
or electron transfer
• CuS2-~Cu 2+ + 2S + 2 e - (25)

TABLE4

C o m p a r i s o n of t h e corrosion c u r r e n t s of chalcopyrite [70 ]

Ox Red E ...... j ...... Source


(V vs. S C E ) (kA/m'-')

Fe :*+ Fe 2+ 0.39 55 Wards


Fe :~+ / C u ~+ Fe 2+ 0.48 110 Wards
Cu ~~ Cu + 0.28 2.5 Wards
180

The first step illustrates the breaking of the Fe-S bond and the second the
breaking of the Cu-S bond. "CuS2 is an unstable radical intermediate. This
mechanism implies that copper remains in the lattice because it is associated
with the valence band, which is consistent with the observation that the Fe/
Cu ratio in solution is greater than 1 in the initial stages of leaching.
The radical intermediate, -CuS> may decompose to form a stable CuS in-
termediate according to
•CuS., ~ C u S + S (26)
Biegler and Swift [81 ] presented cyclic voltammogrammes for pure natural
chalcopyrite. A large anodic response occurs at potentials greater than 1.0 V.
Prior to this response, a smaller "prewave" response occurs at 0.8 V, which
they interpreted as the formation of a copper sulphide film.

The formation of surface films on CuFeS2

The application of a positive potential pulse to the chalcopyrite electrode


produces an anodic current response due to oxidation of the electrode. This
current subsequently decays with time to a steady state value. Reactivation
occurs if the electrode is kept on open circuit and, when the potential is re-
applied, large currents flow that again decay with time. Parker et al. [70] in-
terpreted this as the formation of a metal-deficient polysulphide film that slows
the transport of Cu ÷ and Fe 2+ ions to the solution. This film is supposed to
break down thermally (reactivation of the electrode) to produce sulphur.
Munoz et al. [84] concluded that the kinetics in sulphate solutions were
controlled by diffusion through a sulphur product-layer. Thus, they attributed
the high activation energy to the rate-determining transport of electrons
through the sulphur product layer, and used Wagner's theory to describe the
kinetics. However, the sulphur layer is porous and is expected to have a high
concentration of impurities. Parker et al. [ 70 ] removed the sulphur with CS2
and failed to observe any effect on the rate. Therefore it is not clear that the
sulphur forms a protective layer.
McMillan et al. [90] reported anodic currents that decayed with time. A
series of potential pulses in the range 0.2-0.6 V vs. SCE revealed a blue product
layer, suggesting a copper-rich sulphide phase. McMillan et al. [90] inter-
preted their results as showing the formation of a protective layer in which the
diffusion of ions was rate-limiting. McMillan et al. [90] compared the electro-
chemistry of n- and p-type samples. However, their conclusion that the semi-
conduction type had no influence on the electrode behaviour seems to be
unjustified. The main reason is that they used natural crystals, and their p-
type specimen had significant pyrite inclusions, rendering their conclusions
doubtful because of the possibility of galvanic interactions.
The growth of a film should lead to Tafel slopes greater than 0.118 V decade- 1
181

[91 ]. McMillan et al. [90] reported slopes of 0.2 V per decade due to the for-
mation of a film of CuS or intermediate copper sulphide; indeed, a blue surface
film has been frequently observed [70, 83, 90], suggesting covellite. If such a
film formed on the surface of the electrode, the current would decrease until a
stable copper sulphide formed. Moreover, reactivation would occur on open
circuit as the original chalcocite-covellite intermediate formed, since both the
solid-solid and the solid-solution reactions are reversible [92].
Biegler and Horne [93] analysed the anodic "prewave" and presented evi-
dence for the formation of a CuS film about 3 nm thick. The ratio of iron to
copper during leaching, the dissolution of copper during the prewave, and an
analysis of the charge passed in reducing the prewave product suggest that
chalcopyrite initially produces a CuS film, sulphur, Cu 2+ and Fe 2+.
The copper sulphide film, or whatever layer is formed, does not inhibit the
rate of' leaching in chloride solutions after the layer thickness has reached a
steady state, since linear kinetics and a strong dependence on ferric chloride
concentration are observed (the one-half dependence on Fe :~+ suggests metal-
like corrosion control due to Fermi-level pinning). In sulphate solutions, how-
ever, a layer does apparently inhibit leaching, since "parabolic" kinetics are
observed. This difference in behaviour might be due to the ability of' chloride
ions to be specifically adsorbed at the surface, whereas sulphate ions are not.
This would result in surface states within the film band gap that would aid the
passage of current, and hence would catalyze the electrochemical dissolution
of the film.
The production of sulphate as the final chalcopyritic sulphur product is en-
visaged as occurring by the same mechanism as that discussed for pyrite. Holes
that have been injected into the Cu 3d contribution to the valence band are
capable of oxidizing water, and the • OH radicals directly oxidize the valence-
band sulphur to sulphate. The oxidation of elemental sulphur is very slow, and
is not considered to contribute to the production of sulphate.

THE DISSOLUTION AND ELECTROCHEMISTRY OF COPPER SULPHIDES

Chalcocite

Sullivan [94] studied the dissolution of chalcocite ores containing 5 to 10%


Fe at temperatures in the range 23 to 95 ° C. Cu2S dissolves according to the
reaction
Cu~ S + 2Fe :~+ ~ Cu ~+ + 2Fe :~+ + CuS ( 27 )
CuS + 2Fe :~+ ~ Cu ~+ + 2Fe :~+ + S (28)
where the CuS intermediate is not the mineral covellite. Sullivan noted that
the rate is independent of pH.
182

Marcantonio [95] identified two stages in the ferric sulphate leaching of


natural chalcocite in which the product sequence Cu2S ~ CUl.8S ~ Cul.2S ~ CuS
occurred. The first stage included the rapid formation of Cul.8S and the slower
formation of blue-remaining covellite, CUl.2S. The second stage is the disso-
lution of Cu~.0S to form Cu u+ in solution and elemental sulphur. The kinetics
of' the first stage was first order in Fe :~+ concentration and the activation en-
ergy was low (11.5 kJ mol-~); the leaching results were modelled as solution
difthsion control in conjunction with spherical shrinking core geometry. The
second-stage reaction was slower than the first stage, and the kinetics were
half order in Fe :~+ with a high activation energy (75.5 kJ mol -l ); the leaching
results for this latter stage were modelled as an electrochemical reaction with
cylindrical particle geometry.
Copper sulphide minerals have a solid solution range Cu0S-CuS, the ther-
modynamics of which are not clearly understood [96]. Chalcocite is consis-
tently p-type, and has a band gap of about 1.8 eV. An indirect adsorption edge
at 1.1 eV fbr light perpendicular to the c-axis and at 1.4 eV fbr light parallel to
the c-axis, has been attributed to interband transitions. The conduction band
is derived from the Cu 4s orbital, and the valence band from the S 3p level
[97 ]. An empirical energy level diagram constructed from sulphur X-ray emis-
sion spectra of' Cu~S is in agreement with the quantum mechanical calculations
of the triangular coordinated CuS~- cluster. An ionic model for chalcocite would
be (Cu + },_,So and, for a copper-deficient chalcocite, Cu0 xS, it would be
(Cu + )0_0x(Cu u+ )~S °-. In the band model, a hole introduced for each missing
Cu + (acceptor defect) is delocalized, with a mobility of 2-10 cm ~ V - ' s - '
Chalcocite exhibits high ionic mobility in the range 5-10 cm ° V -I s ', due to
the high mobility of Cu + ions.
Koch and MacIntyre [92] prepared thin film electrodes of synthetic Cu0S
by vacuum evaporation, and measured the changes in the reflectance and ab-
sorbance of' the electrode with anodic dissolution (removal of copper ions).
They observed that, when the anodic current was interrupted, the original
spectrum returned. In addition, CuS produced by anodic treatment could be
converted to the original material by reversal of the current, evidence of the
reversibility of both the solid-state and solid-solution reactions. Figure 10 plots
the open-circuit potentials of the electrode as a function of copper removal.
The plateaux correspond to two-phase regions, and the sharp rises to compo-
sitional changes. The phases identified are: Cu2S; Cul.9~,S to CU1,1iS; Cul.s~;S to
Cul.sS; Cuj.~sS to CUl.658; Co1.4S to CUl.368; CoS.
The wavelength of the maximum absorbance decreased from 2300 to 1200
nm for the Cu~.9,~ ~.9~Sto 610 nm for CuS. Wavelengths less than 1250 nm were
interpreted as band to band transitions. The shift to lower wavelengths re-
fleeted the progressively lower Fermi level in the valence band as the copper
content decreased. The extraction of copper ions from Cu0S was accompanied
183

320-

3o0 ! ,..'.:
280-

260-

,,~ 240-

220 -

"~ 20{}-

180-

160-

140-

120-

I0O I 1 I I I
Cu~S 20 40 60 80 CuS
Copper removed (%)
Fig. 10. R e s t p o t e n t i a l s o f a t h i n f i l m c h a l c o c i t e e l e c t r o d e as a f u n c t i o n o f t h e c o p p e r c o n t e n t i n
t h e r a n g e Cu~S C u S [92].

by the removal of electrons from the top of the valence band. As a result, the
adsorption band shifted to higher energies since the difference in energy be-
tween the highest filled levels of the valence band and the conduction band
increased [98].
Hillrichs and co-workers [ 99, 100 ] investigated the electrochemistry of syn-
thetic Cu2S using cyclic voltammetry. They reported a sharp anodic peak at
0.4 V vs. 3.5 M calomel electrode (CE) in which the diffusion of copper in the
solid was claimed to be rate-determining. This peak reached a maximum due
to the formation of CuS which apparently limit further diffusion of copper
ions. Results at potentials greater than 1 V vs. CE were interpreted in terms
of CuO formation and dissolution. Figure 11 illustrates their interpretation of
the cyclic voltammogram of Cu2S and CuS.
Biegler and Swift [ 101 ] used natural and synthetic chalcocite and synthetic
digenite anodes. In their galvanostatic experiments the potential rose steeply
after a certain period, which led the authors to concur with Ettienne and Peters
184

CuxS ~ C u S
0.6-

.,"" 0.4-
r- CuS~S
<
E
0.2

0--
Cu2S~CuS
! I I
(a) 0 1 2 3
Potential (V vs 3.5 M CE)

cuo /

(b)
ot 0
,
1
,
2
,
3
Potential (V ~s 3.5 M CE)
Fig. 11. Cyclic v o l t a m m o g r a m of (a) Cu,2S a n d (b) C u S in 1M H,2SO4 [100].

[ 102 ] that the diffusion of copper was impeded by the precipitation of a copper
salt in the pores of the solid. Price [ 103 ] later re-analysed the published data
and proposed a model that combined solid-state diffusion and electrochemical
reaction in the first stage of reaction (product CuS).

CoveUite

Thomas and Ingraham [104] leached rotating discs of synthetic CuS in 0.1
M H2SO4and 0.25 M F% (SO4) 3. The rate of reaction was determined from the
concentration of the copper ions in solution. They reported a transition from
interfacial control at 25 °C to solution diffusion control at 80 ° C. Dutrizac and
185

MacDonald [105] reported that the rates of dissolution of natural and syn-
thetic CuS were slow, but increased during the initial stages of dissolution
owing to pit formation. The reaction
CuS + 2Fe :~+ - . C u 2+ + 2Fe 2+ + S (29)
was observed to be directly dependent on Fe :~+ concentration below 0.005 M,
but independent of Fe :~+ concentration above 0.005 M. The rate-controlling
step was tound to be a reaction at the surface. The sulphur product did not
tbrm a protective layer.
The crystal structure of covellite is more complex than that of sphalerite,
galena, pyrite or chalcopyrite [40]. The copper occurs in both triangular and
approximately tetrahedral co-ordination, and two thirds of the sulphur atoms
are S~- anions [106]. The proposed ionic model for covellite is
CubIT (Cu~),,$2-S~ - in which Cu 2÷ is in a triangular site, and Cu ÷ is in a
tetrahedral site. The calculated electronic structure for CuS indicates that the
top of the partially-filled valence band is predominantly Cu 3d in character,
and the conduction band is Cu 4s in character. CuS is a p-type metallic con-
ductor because the valence band is partially filled. The energy difference be-
tween the Fermi energy and the bottom of the conduction band is about 2 eV.
The passivation behaviour of eovellite was investigated by Hillrichs and
Bertram [ 107 ] using cyclic voltammetry. They reported a dependence of 0.4 V
per pH unit, and showed that the open-circuit potentials were compatible with
the tbrmation of a layer of copper oxide or copper hydroxide. Dissolution oc-
curred only at relatively high potentials ( > 1.3 V vs. 3.5 M CE) owing to the
formation and dissolution of the CuO film. This may also have been due to the
fact that the upper valence band is mainly a crystal field band, similar to the
valence band of pyrite. Peters [7, 39] and Ghali et al. [108] proposed that the
sulphur product-layer hindered the rate of reaction.

GALVANIC INTERACTIONS DURING DISSOLUTION

When two minerals of different open-circuit potentials are in contact, the


mineral with the more positive potential is reduced, while the mineral with the
more negative potential is oxidized. The difference in their open-circuit poten-
tials is the driving force of galvanic corrosion.
Although the study of galvanic interactions has been established experimen-
tally [109, 110], most of the results have been interpreted only in a semi-quan-
titative manner. Nowak et al. [111] studied galvanic couples using natural
mineral electrodes, and found that the difference in open-circuit potentials was
in the order: galena-pyrite>chalcocite--pyrite>>chaleocite-chalcopyr-
ite > galena-chaleopyrite > chaleopyrite-pyrite.
Hepel et al. [ 112 ] considered the influence of the solid junction between two
semiconducting minerals of p- and n-type on the galvanic corrosion current.
186

The polarization of a mineral p-n junction is, of course, dependent on the


open-circuit potential of each mineral. (The open-circuit potential of a min-
eral is a measure of the Fermi level, or the electrochemical potential of the
electrons in the solid, with respect to the reference half cell. The open-circuit
potentials of n- and p-type specimens of the same mineral differ by the differ-
ence in their Fermi levels, the maximum being the energy of the band gap. This
has been claimed for pyrite [5 ] and galena [ 113 ]. ) If the open-circuit potential
of the p-type mineral is higher than that of the n-type mineral, the potential
difference across the junction is small, and high currents can flow; if the p-type
mineral has a more negative open-circuit potential than the n-type mineral,
the potential difference across the junction is high, and may be high enough to
prevent any current across the junction. Hepel et al. [112] derived a general
relation with which to relate the electrochemical parameters of the system and
an anodic acceleration coefficient. Their experimental results confirm the rec-
tifying properties of galena-galena, galena-chalcopyrite, and chalcocite-pyrite
electrodes, although these were not of the ideal Schottky barrier type. The
fbrmation of surface states and Fermi level pinning was observed. Hepel et al.
concluded that, despite the high level of impurities in natural minerals, the
ohmic and semiconducting properties of minerals forming solid junctions should
always be taken into account when galvanic interactions are being considered.

SUMMARY AND CONCLUSIONS

The importance of semiconductor and other solid-state properties in the


anodic electrochemistry of metal sulphides is now widely accepted. In general,
sulphide electrodes do not display classical semiconductor-solution behaviour
owing to the formation of surface states that aid the transfer of charge. The
presence of surface states, especially those due to zero valent sulphur, may give
rise to behaviour that is intermediate between that of a metal and a semicon-
ductor [43, 44].
Generally speaking, the binary structure of the mineral sulphides is such
that sulphur anions provide an immobile lattice for more mobile metallic cat-
ions [62]. Hence, the equilibrium between the metal ions in solution and the
surface is important, and strongly affects electrode behaviour [51]. The dis-
solution of sulphide semiconductors occurs mainly by a hole mechanism, which
is corroborated by the enhanced rate of dissolution of most sulphides when the
dissolving material is irradiated by light having energy greater than the band
gap energy.
The bonding structure of the solid is particularly important with transition
metal sulphides. In sphalerite, the substitutional iron content provides d-band
energy levels within the band gap; in pyrite, crystal-field splitting results in a
187

valence band that has a non-bonding character, and is therefore less easily
oxidized.
The non-stoichiometry of galena influences the semiconduction type and
hence the kinetics of dissolution. A Tafel slope consistent with the semicon-
ductor model is observed, although this is influenced by the equilibrium with
lead ions in solution, and the applied potential is distributed across both the
space-charge layer and the Helmholtz layer.
The kinetics of chalcopyrite dissolution is complicated by the formation of
a surface film, probably a film of copper sulphide [93]. The influence of redox
couples is in accordance with the semiconductor model, and a dissolution
mechanism involving both electrons and holes is proposed.
Hepel et al. [112] have discussed the importance of the rectifying properties
of semiconductor junctions in galvanic interactions. They, too, reported the
presence of surface states that give rise to non-ideal solid junction behaviour.
Ore sulphides contain larger concentrations of impurities than synthetic sul-
phides, and this makes the direct application of semiconductor principles to
leaching and flotation more difficult. However, any study of the electrochem-
istry of leaching and flotation must take into account the influence of the elec-
tronic structure of the mineral.

ACKNOWLEDGEMENT

This paper is published by permission of the Council for Mineral Technology


(MINTEK). The suggestions of Dr R.L. Paul are gratefully acknowledged.

REFERENCES

1 Simkovich, G. and Wagner, J.B., J. Electrochem. Soc., 110 (1960) 513.


2 Eadington, P. and Prosser, A.P., Trans. Inst. Min. Metall., 78 (1969) C74.
3 Richardson, P.E. and O'Dell, C.S. (1984). In: Proc. Int. Symp. on Electrochem. in Mineral
and Metal Processing, P.E. Richardson, S. Srinvasan and R. Woods (Eds.), Electrochem.
Soc., Pennington.
4 Pawlek, F., J.S. Afr. Inst. Min. Metall., 69 (1969} 632.
5 Springer, G., Trans. Inst. Min. Metall., 79 (1970) C l l .
6 Daiger, K. and Gerlach, J., Erzmetall., 35 (1982) 609.
7 Peters, E., (1984). In: Proc. Int. Symp. on Electrochem. on Mineral and Metal Processing,
P.E. Richardson, S. Srinvasan, R. Woods (Eds.), Electrochem. Soc., Pennington.
8 Gerischer, H., (1970). Semiconductor electrochemistry. In: H. Eyring, D. Henderson and
W. Jost (Eds.), Physical Chemistry - an advanced treatise, Vol. IXA, Academic Press, New
York.
9 Garrett, C.G.B. and Brattain, W.H., Phys. Rev., 99 (1955) 376.
10 Frank, S.N. and Bard, A.J., J. Am. Chem. Soc., 97 (1975) 7427.
11 Peter, L.M., Li, J. and Peat, R , J. Electroanal. Chem., 165 {1984) 29.
12 Bard, A.J., J. Am. Chem. ~9c., 102 (1980) 3671.
188

13 Gerischer, H., (1960). In: C.W. Tobias and P. Delahay (Eds.), Advances in Electrochem.
and Electrochem. Eng., Vol. 1, Interscience.
14 Green, M., (1959). In: J. O'M. Bockris (Ed.), Modern aspects of electrochemistry, Vol. 2,
Butterworths, London.
15 Morrison, S.R., (1980). Electrochemistry of semiconductor and oxidized metal electrodes.
Plenum, New York.
16 Memming, R., (1960). In: B.E. Conway (Ed,), Comprehensive Treatise on Electrochem-
istry, vol. 7, Plenum, New York.
17 Dewald, J.F., (1980). In: H. Gatos (Ed.), Surface Chemistry of Metals and Semiconduc-
tors, John Wiley and Sons, New York.
18 Butler, M.A. and Ginley, D.S., J. Electrochem. Soc., 125 (1978) 228.
19 Marcus, R.A., Ann. Rev. Phys. Chem., 15 (1964) 155.
20 Levich, V.G., (1970). In: H. Eyring (Ed.), Physical Chemistry an advanced treatise, vol.
IXB, Academic Press, New York.
21 Bockris, J.O'M. and Khan, S.U., J. Electrochem. Soc., 132 (1985) 2648.
22 Morrison, S.R., J. Vac. Sci. Technol., 15 (1978) 1417.
23 Frese, K.W.,J. Phys. Chem.,85 (1981)3911.
24 Mehl, W. and Hale, J.M., (1966). In: C.W. Tobias and P. Delahay (Eds.), Advances in
Electrochem. and Electrochem. Eng., Vol. 6, Interscience.
25 Ginley, D.S. and Butler, M.A., J. Electrochem. Soc., 125 (1978) 1968.
26 Gurney, R.N.,Proc. Roy. Soc. London, A134 (1931) 137andA136 (1932) 378.
27 Gerischer, H., (1960). In: H. Gatos (Ed.), Surface Chemistry of Metals and Semiconduc-
tors, John Wiley and Sons, New York.
28 Khan, S.U., (1983). In: R. White (Ed.), Aspects of Modern Electrochemistry, Vol. 15,
Plenum Press, New York.
29 Turner, D.R., J. Electrochem. Soc., 103 (1956) 252.
30 Boddy, P.J., J. Electrochem. Soc., 111 (1964) 1136.
31 Gerischer, H. and Mindt, W., Electrochimica Acta, 13 (1968) 1329.
32 Bard, A.J. and Wrighton, M.S., J. Electrochem. Soc., 124 (1977) 1706.
33 Gerischer, H., J. Electroanal. Chem., 82 (1977) 133.
34 Koch, D.F.A., (1975). In: J.O'M. Bockris and B.E. Conway (Eds.), Modern Aspects of
Electrochemistry, Vol. 10, Plenum Press, New York.
35 Hiskey, J.B. and Wadsworth, M.E., (1981). In: M. Kuhn (Ed.), Process and fundamental
considerations in selected hydrometallurgical systems, SME-AIME, New York, p. 303.
36 VJadsworth, M.E., Mineral Sci. Eng., 4(4) (1972) 36.
37 Dutrizac, J.E. and MacDonald, R.J.C., Mineral Sci. Eng., 6(2) (1974) 59.
38 Burkin, A.R., Mineral Sci. Eng., 1 (1) (1969) 4.
39 Peters, E., (1977). In: J.O'M. Bockris (Ed.), Trends in Electrochemistry, Plenum, New
York.
40 Shuey, R.T., (1975). Semiconducting Ore Minerals, Elsevier, Amsterdam.
41 Vaughan, D.J. and Craig, J.R., (1978). Mineral Chemistry of Metal Sulphides, Cambridge
University Press, London.
42 Gerischer, H., J. Electrochem. Soc., 116 (1966) 1174.
43 Tyagai, V.A. and Kolbasov, G.Y., Surf. Sci., 28 (1971) 423.
44 Van den Berghe, R.A.L., Ber Bunsenges Phys. Chem., 78 (1974) 331.
45 Morrison, S.R., (1977). The Chemical Physics of Surfaces, Plenum, New York.
46 Gerischer, H. and Meyer, E., Z. Phys. Chem. N.F., 74 (1971) 302.
47 Brodie, J.B., (1968). M.Sc. Thesis, University of British Columbia.
48 Plaskin, I.N. and Shafeev, R.S., Trans. Inst. Min. Met., 72 (1963) 715.
189

49 Holberg, H. and Schneider, F.U., (1975). In: Proc. of l l t h Int. Min. Processing Congr.,
Cagliari.
50 Eadington, P., Trans. Inst. Min. Metall., 82 (1973) C158.
51 Paul, R.L., Electrochimica Acta, 23 (1978) 625.
52 Exner, F., Gerlach, J. and Pawlek, F., Erzmetall., 22 (1969) 219.
53 Piao, S.Y. and Tozawa, K., J. Min. Metall. Inst. Jpn., 101 (1985) 795.
54 Jin, Z., ( 1985 ). In: K. Tozawa (Ed.), Proc. Int. Symp. on Extractive Metall. of Zinc, Tokyo.
55 Crundwell, F.K., Report M298, MINTEK, Randburg, 1987.
56 Mott, N.F. and Gurney, R.W., (1948). Electronic Processes in Ionic Crystals, 2nd edn.,
Oxford University Press, New York.
57 Telkes, M., American Mineral, 35 (1950) 536.
58 Keys, J.D., Can. Mineral., 9 (1968) 453.
59 Majewski, J.A., Phys. Stat. Sol, B, 102 (2) (1981) 663.
60 Jellinek, F., (1972). In: D.W.A. Sharpe (Ed.), MTP Int. Review of Science, Transition
Metals, Part 1, Inorganic Chemistry Series 1, Vol. 5. Butterworth, London.
61 Horowitz, G., J. Electrochem. Soc., 131 (1984) 2563.
62 Williams, R., J. Chem. Phys., 32 (1960) 1505.
63 Hiskey, J.B. and Schlitt, W.J., (1981). In: Proc. 2nd SME SPE Int. Solution Mining Syrup.,
Denver, AIME, p. 55.
64 Warren, I.H., Austr. J. Appl. Sci., 7 (1956) 346.
65 Majima, H., Can. Metall. Q., 8 (1969) 269.
66 Dutrizac, J.E. and MacDonald, R.J.C., Mineral Sci. Eng., 6 (1974) 59.
67 Bailey, L.K. and Peters, E.. Can. Metall. Q., 15 (1976) 333.
68 Biegler, T. and Swift, D.A., Electrochimica Acta, 24 (1979) 415.
69 Meyer, R.E., J. Electroanal. Chem., 101 (1979) 59.
70 Parker, A.J., J. Electroanal. Chem., 118 (1981) 305 and Austr. J. Chem., 34 (1981) 13.
71 Goodenough, J.B., J. Solid State Chem., 5 (1972) 144.
72 Bither, T.A., Inorganic Chem., 7 (1968) 2208.
73 Tributsch, H., Z. Naturforsch., 32a (1977) 972.
74 Tributsch, H. and Bennet, J.C., J. Electroanal. Chem., 81 (1977) 97.
75 Jaegerman, W. and Tributsch, H., J. Appl. Electrochem., 13 (1983) 743.
76 Ennaoui, A., J. Electrochem. Soc., 133 (1986) 97.
77 Memming, R.J., J. Electrochem. Soc., 116 (1969) 785.
78 Boddy, P.J., J. Electrochem. Soc., 111 (1964) 1136. and 110 (1963) 570.
79 Biegler, T., J. Electroanal. Chem., 70 (1975) 265.
80 Krishnasway, P. and Fuerstenau, D.W., (1982). In: K. Osseo-Assare and J.D. Miller (Eds.),
Hydrometallurgy, Research, Development and Plant Practice, TMS-AIME, Warrendale,
Pennsylvania.
81 Biegler, T. and Swift, D.A., J. Appl. Electrochem., 9 (1979) 545.
82 Dutrizac, J.E., Metall. Trans. B, 12B (1981) 371.
8'3 Jones, D.L. and Peters, E., (1976). In: J.C. Yannopoulos and J.C. Agarwal (Eds.), Extrac-
tive Metallurgy of Copper, Vol. 2, p. 633, AIME, New York.
84 Munoz, P.B., Miller, J.D. and Wadsworth, M.E., Metall. Trans., 10B (1979) 149.
85 Linge, H.G., Hydrometallurgy, 2 (1976) 51.
86 Shay, J.L. and Tell, B., Surf. Sci., 37 (1973) 748.
87 Jaffe, J.E. and Zunger, A., Phys. Rev. B, 28 (1983) 5822.
88 Beckstead, L.W., (1976). In: J.C. Yannopoulos and J.C. Agarwal (Eds.), Extractive Me-
tallurgy of Copper, Vol. 2, AIME, New York.
89 Zevgolis, E.N. and Cooke, S.R.B., (1975). In: Proc. 11th Int. Mineral Processing Congr.,
Cagliari.
190

90 McMillan, R.S., MacKinnon, D.J. and Dutrizac, J.E., J. Appl. Electrochem., 12 (1982) 743.
91 Vijh, A.K., (1973). Electrochemistry of metals and semiconductors, Marcel Dekker, New
York.
92 Koch, D.F.A. and McIntyre, R.J., J. Electroanal. Chem., 71 (1976) 285.
93 Biegler, T. and Horne, B., (1984). In: P.E. Richardson (Ed.), Proc. Int. Symp. on Electro-
chem. in Mineral and Metal Processing, The Electrochem. Soc., Pennington.
94 Sullivan, J.D., Trans. Am. Inst. Min. Metall., 106 (1933) 515.
95 Mercantonio, P., (1976). Ph.D. Thesis, University of Utah.
96 Potter, R.W., Econ. Geol., 72 (1977) 1524.
97 Tossell, J.A. and Vaughan, D.J., Inorg. Chem., 20 (1981) 3333.
98 Mulder, B.J., Phys. Status Solidi A, 13 (1972) 79.
99 Hillrichs, E. and Bertram, R., Hydrometallurgy, 11 (1983) 181.
100 Hillrichs, E., (1982). In: K. Osseo-Assare and J.D. Miller (Eds.), Hydrometallurgy, Re-
search Development and Plant Practice, TME-AIME, Warrendale, Pennsylvania.
101 Biegler, T. and Swift, D.A., Hydrometallurgy, 2 (1977) 335.
102 Ettienne, A. and Peters, E., Trans. Inst. Min. Me,all., 8l (1972) C176.
103 Price, D.C., Metall. Trans. B, 12B (1981) 231.
104 Thomas, G. and Ingraham, T.R., Can. Metall. Q., 6 (1967) 153.
105 Dutrizac, J.E. and MacDonald, R.J.C., Can. Metall. Q., 13 (1974) 423.
106 Vaughan, D.J. and Tossell, J.A., Can. Mineral., 18 (1980) 157.
107 Hillrichs, E. and Bertram, R., Hydrometallurgy, 11 (1983) 195.
108 Ghali, E., J. Appl. Electrochem., 12 (1982) 369.
109 Berry, V.K., Hydrometallurgy, 3 (1978) 309.
110 Mehta, A.P. and Murr, L.E., Hydrometallurgy, 9 (1983) 235.
111 Nowak, P., Hydrometallurgy, 12 (1984) 95.
112 Hepel, T., (1984). In: P.E. Richardson (Ed.), Proc. Int. Symp. on Electrochem. in Mineral
and Metal Processing, Electrochem. Soc., Pennington.
113 Gutierrez, C. and Lopez, M.C., J. Electrochem. Soc., 125 (1978) 678.

You might also like