Você está na página 1de 12

Hydrometallurgy 80 (2005) 1 12

www.elsevier.com/locate/hydromet

Short review

Kinetics and reaction mechanism of gold cyanidation:


Surface reaction model via Au(I)OHCN complexes
G. Senanayake *
A.J. Parker Cooperative Research Centre for Hydrometallurgy, Department of Mineral Science and Extractive Metallurgy,
Murdoch University, Perth, WA 6150, Australia
Received 25 February 2005; received in revised form 20 July 2005; accepted 3 August 2005
Available online 15 September 2005

Abstract
The current status of the mechanism of gold cyanidation based on diffusion and surface adsorptionreaction models are
reviewed. Published rate data based on chemical oxidation from flat gold surfaces in pure aerated cyanide solutions are
analysed to show a reaction order of 2.7 with respect to cyanide at low concentrations. At higher cyanide concentrations, the
reaction rate reaches a limiting value of R Au(lim) = 7.3  10 6 mol m 2 s 1, independent of the cyanide concentration and
stirring rate. This chemically controlled dissolution of gold in pure cyanide solutions is considered to be different from the
widely reported cyanide or oxygen diffusion controlled dissolution of gold, depending on their relative concentrations. The
proposed reaction mechanism to rationalise this behaviour involves the formation of a heterogeneous redox transition state
(Au.H2O)2.(CN)23.(O2) which produces the intermediate Au(I)(OH)(CN) on the gold surface. Oxygen is reduced to
hydrogen peroxide which may degrade in three ways: (i) oxidize gold to produce the same gold(I) intermediates on surface,
(ii) oxidize cyanide to cyanate (iii) disproportionate to water and oxygen. The surface adsorbed Au(I) intermediate reacts with
cyanide to produce more stable Au(CN)2 in solution. The proposed surface chemical model rationalises the reaction order of
c 3 at low cyanide concentrations and calculates an intrinsic rate constant of k Au = 8.6  10 6 mol m 2 s 1 for gold
cyanidation by oxygen. This value is in reasonable agreement with the value of k = 6.9  10 6 mol m 2 s 1 based on the
model proposed by Wadsworth et al. [Wadsworth, M.E., Zhu, X., Thompson, J.S., Pereira, C.J., 2000. Gold dissolution and
activation in cyanide solution: kinetics and mechanism. Hydrometallurgy, 57, 111.], which considered the mass transfer away
from the active crystalline gold surface site followed by fast charge transfer, combined with two-electron reduction of oxygen
on the gold surface.
D 2005 Elsevier B.V. All rights reserved.
Keywords: Gold cyanidation; Reaction mechanism; Surface transition state; Kinetics; Rate constants; Gold(I) speciation

* Tel.: +61 8 93602833; fax: +61 8 93606343.


E-mail address: G.Senanayake@murdoch.edu.au.
0304-386X/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.hydromet.2005.08.002

G. Senanayake / Hydrometallurgy 80 (2005) 112

Contents
1.
2.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Current status of reaction mechanism . . . . . . . . . . . . . . .
2.1. Limitations of diffusion model due to effect of impurities .
2.2. Formation of surface films . . . . . . . . . . . . . . . . .
2.3. Surface adsorptionreaction models . . . . . . . . . . . .
2.4. Need for further studies . . . . . . . . . . . . . . . . . .
3. Gold(I) speciation: justification and importance . . . . . . . . . .
4. Analysis of rate data. . . . . . . . . . . . . . . . . . . . . . . .
4.1. Levich equation . . . . . . . . . . . . . . . . . . . . . .
4.2. Rate data . . . . . . . . . . . . . . . . . . . . . . . . . .
5. Reaction mechanism. . . . . . . . . . . . . . . . . . . . . . . .
6. Summary and conclusions . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Gold cyanidation has been reported to involve the
chemical reactions shown in Eq. (1) (Bodlander, 1896)
and Eq. (3) (Elsner, 1846), where Eq. (3) can be
treated as the sum of the two partial reactions shown
in Eqs. (1) and (2). Hydrogen peroxide produced at
the interface by reduction of oxygen can react with
gold (Eq. [(2)), or with cyanide ion (Eq. (4)), or
disproportionate to H2O + 0.5O2; while cyanate produced in solution further degrades to other products.

2Au4CNO2 2H2 O2AuCN
2 2OH H2 O2

1

2Au 4CN H2 O2 2AuCN
2 2OH


4Au 8CN O2 2H2 O 4AuCN
2 4OH

3
CN H2 O2 CNO H2 O

Kudryk and Kellogg (1954) highlighted the importance of understanding the rate controlling factors of
gold cyanidation which would allow correct choice of
conditions such as agitation, temperature, and the
reagent concentrations. They showed the electrochemical nature of the gold cyanidation reaction and that
the rate is determined by the rate of diffusion of
cyanide or dissolved oxygen to the gold surface,
depending on their relative concentrations.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

2
3
3
3
4
5
5
6
6
7
8
11
11

The theoretical and practical aspects of gold cyanidation have been frequently reviewed (Cornejo and
Spottiswood, 1984; Nicol et al., 1987; Li et al., 1992;
Fleming, 1992), while the fundamental aspects of gold
cyanidation reaction and their relevance to leaching of
gold from ores have been advanced by Cathro (1963),
Cathro and Koch (1964), MacArthur (1972), Nicol
(1980), Kirk et al. (1980), Dorin and Woods (1991);
Osseo-Asare et al. (1984), Lorenzen and van Deventer (1992); Zheng et al. (1995), Guan and Han
(1994), Crundwell and Godorr (1997), Wadsworth
et al. (2000), Jeffrey and Ritchie (2000a,b, 2001)
and Xue and Osseo-Asare (2001). Significant progress has been made in recent years on mixed
electrochemical-transport (diffusion) models that
are capable of explaining the effect of oxygen
pressure, cyanide concentration and agitation (Li
et al., 1992; Wadsworth et al., 2000).
Despite long-term interest and industrial application, the reaction mechanism of gold cyanidation by
oxygen is still being debated and/or investigated
especially in the following three areas:
(i) stoichiometry and chemical or diffusion controlled nature of the reaction,
(ii) nature of the passivation layer on gold surface,
(iii) effect of host minerals and impurities in the
solid state or solution.
Although a number of different leaching models
have been presented, it is difficult to categorise them
due to the fact that each model considers a number

G. Senanayake / Hydrometallurgy 80 (2005) 112

of different factors such as diffusion, adsorption,


charge transfer, surface films etc. Nevertheless, it
is important to revise these models in order to
understand the current status of the reaction mechanism of pure gold in aerated cyanide solutions in the
absence of impurities in solid and aqueous phases.
This is useful to develop a surface chemical model
that would rationalise the role of metal ions in
aqueous phase, alloyed metals, and the effect of
host minerals.
This work describes a surface chemical model on
the basis of surface adsorbed species such as Au(OH)0
and Au(OH)(CN) to rationalise the results reported
by Jeffrey and Ritchie (2001) and to compare with the
adsorption-charge transfer model proposed by Wadsworth et al. (2000).

2. Current status of reaction mechanism


2.1. Limitations of diffusion model due to effect of
impurities
Based on electrochemical studies, Kirk and
Foulkes (1978) described gold dissolution as a reaction controlled by aqueous boundary layer diffusion
(mass transfer) of cyanide and oxygen to the gold
surface. Zheng et al. (1995) showed that the rates of
gold cyanidation based on a rotating disc electrode
quartz crystal microbalance (REQCM) were lower
than the predicted values based on the Levich equation. They related this to the effect of a boundary film
on the gold surface and the impurities in gold and/or
solution. Li et al. (1992) showed how the mixed
potential theory can be successfully used to model
the cyanidation kinetics reported by Kudryk and Kellogg (1954) by combining charge transfer with cyanide ion diffusion to the surface.
Jeffrey and Ritchie (2000b) found that the measured value for the rate of gold cyanidation by oxygen
in an air saturated solution of 20 mM cyanide on the
basis of oxygen transfer and stoichiometry in Eq. (3)
was smaller than the predicted value. Moreover, Jeffrey and Ritchie (2001) showed that pure gold has a
very low rate of dissolution in aerated ultra-pure
cyanide solutions, while the reaction is chemically
controlled rather than diffusion controlled. Thus,
they attributed some of the conflicting results on

gold cyanidation reported by previous researchers to


the different methods of stirring and the presence of
impurities in cyanide or other electrolytesas well as
impurities in the gold discs used in kinetic studies. For
example, alloyed silver and copper (Choi et al., 1991;
Sun et al., 1996; Xue and Osseo-Asare, 2001; Breuer
et al., 2005), and even the minor amounts of lead(II)
contaminated with analytical grade sodium cyanide
and sodium perchlorate (Jeffrey and Ritchie, 2001)
can affect the rate of gold cyanidation. Both silver(I)
and lead(II) cause a positive effect on gold dissolution; but excess dissolved lead(II) (Lorenzen and van
Deventer, 1992) and silver(I) (Wadsworth and Zhu,
2003) retards the leaching kinetics.
2.2. Formation of surface films
It has been widely reported that the anodic oxidation of gold in cyanide media is initiated by the
adsorption of cyanide onto the gold surface, while
passivation and the peaks in polarization curves
have been attributed to the formation of adsorbed
gold(I) and gold(III) species such as Au(I)(CN)0,
Au(I)(OH)0, Au(I)(OH)(CNx ), Au(III)(OH)(CN)3 and
Au(III)(OH)30 (Cathro and Koch, 1964; Kirk et al.,
1980; Nicol et al., 1987; Wadsworth et al., 2000; Xue
and Osseo-Asare, 2001). It is widely accepted that the
cyanidation of gold is a slow process due to passivation of the gold surface by a film of AuCN (Cathro
and Koch, 1964; MacArthur, 1972; Nicol, 1980;
Zheng et al., 1995; Jeffrey and Ritchie, 2001). The
anodic passivation of gold can also be a result of the
adsorbed hydroxide and the formation of AuOH on
the surface (Kirk et al., 1980). Supporting the views of
Nicol et al. (1987), Jeffrey and Ritchie (2001) proposed the formation of a chain-like film of AuCN to
account for the low rates of cyanidation of gold in the
absence of impurities. While the dissolution of AuCN
only occurs at the chain ends, the presence of impurities such as lead ions in the cyanide solutions accelerates this process.
Nicol et al. (1987) noted the difficulties in observing passivation of gold in aerated/oxygenated plant
liquors, despite electrochemical evidence for an
adsorbed layer of AuCN, because the liquors contain
trace heavy metal ions that disrupt the formation of
AuCN. Guan and Han (1994) failed to identify the
species that caused passivation, because it was an

G. Senanayake / Hydrometallurgy 80 (2005) 112

unstable intermediate. Yet, Crundwell and Godorr


(1997) reported the dominance of a passivating
layer of AuCN on the gold surface at latter stages
of batch leaching experiments. They presented a
kinetic model using batch leaching data based on
half order reaction rates with respect to cyanide and
oxygen.
2.3. Surface adsorptionreaction models
According to the widely accepted combined diffusionadsorptionoxidation model, the cyanidation of
gold follows the five steps described by Eqs. (5)(8),
which involve (i) the diffusion of cyanide from bulk
solution to interface denoted by b and i, (ii) surface
adsorption equilibrium, (iii) anodic oxidation, (iv)
stabilisation/desorption of surface products and (v)
diffusion of products into the bulk solution, leading
to the overall reaction given in Eq. (10).

CN
b Y CNi

Aus

CN
i

Au2s 2CN Au2 CN 2ads

13

Au2 CN 2ads CN Au2 CN 3ads

14


Au2 CN 3ads AuCN
ads AuCN 2ads

 rate determining step

15

AuCN
ads

0

AuCN
ads AuCNads e

AuCN0ads

Wadsworth et al. (2000) reviewed the early literature


in support of various adsorptionreaction paths of
gold dissolution. They proposed the involvement of
two (or more) gold atoms as active sites shown by
Au2(s) in Eqs. (13)(15), and bridging cyanide ions in
the rate-determining step, leading to higher reaction
orders up to 3 with respect to cyanide. Based on this
model, incorporating a two-electron reduction of oxygen on gold surface, they derived the rate equation
given by Eq. (17).


CN
i Y AuCN2 i


AuCN
2 i Y AuCN2 b


Aus 2CN
b AuCN2 b e


AuCN
ads CNads


AuCN
2 e fast charge transfer step

16
7
RAu k a Utot CN 3 = f1 KCN 3 g
8
9
10

A series of equations, similar to those in Eqs. (5)(9),


can be written for the cathodic reduction of oxygen
(Hiskey and Sanchez, 1990). Juttner (1984) reported
that the two-electron reduction of O2 to H2O2 predominates on Au substrate. Guan and Han (1994) considered the oxygen reduction as a two-stage process
described by Eqs. (11)(12). More recent studies by
Wadsworth et al. (2000) also considered a two-electron reduction process and showed that dissolved
cyanide depresses the rate of oxygen reduction on a
gold surface.

O2 H2 O 2e HO
2 OH

11



HO
2 H2 O 2e 3OH

12

17

where k a = rate constant for the anodic reaction (mol


m 2 s 1), U tot = total number of adsorption sites shown
by Au2(s) (bare), Au2(CN)2(s) and Au2(CN)3(s) in
Eqs. (13)(14), and K is the product of equilibrium
constants of Eqs. (13) and (14).
Xue and Osseo-Asare (2001) used the mass transfer law and the ButlerVolmer equation to show that
the reaction order with respect to cyanide concentration would depend on the rate-determining step. For
example, a slow rate controlling discharge step shown
by the forward reaction of Eq. (18) would give a
reaction order of 1, as observed by Xue and OsseoAsare (2001) and Guan and Han (1994). In contrast,
an equilibration of the discharge step in Eq. (18),
followed by the forward reaction of the adsorbed
species with cyanide shown in Eq. (8), will give rise
to an overall reaction order of 2 with respect to
cyanide concentration.
0

Aus CN
b AuCNads e

18

G. Senanayake / Hydrometallurgy 80 (2005) 112

2.4. Need for further studies

Table 1
Stability constants of gold(I) and silver(I) complexes

It is important to extend these models, which have


been largely originated from electrochemical studies,
to develop a surface chemical reaction model for the
dissolution of gold in oxygenated cyanide solutions.
Wadsworth et al. (2000) showed how the calculated
rates based on anodic and cathodic currents of gold
oxidation and oxygen reduction combined with the
adsorption-charge transfer model described in Eqs.
(13)(17) were in good agreement with the measured
data using rotating discs in aerated alkaline cyanide
solutions. A model that can be used to compare and
contrast the results reported by previous researchers,
and to rationalise the effect of solid state and solution
impurities, would also need to consider the following
issues, described in the present study.

Complex

Ionic strength

log b a

Ag(CN)
2
Ag(OH)(CN)
Ag(OH)
2
Ag(OH)0
Au(CN)
2
Au(OH)(CN)
Au(OH)
2
Au(OH)0
Au(CH3CN)+2
Au(CH3CN)(OH)0

1(NaClO4)
1(NaClO4)
0
0 (or dil)
0.025(KCN)

20.1
12.8
3.6 (4.2 at 18 8C, 0.2 KNO3)
2.3 (3.9)
36.6, 38.3
23.3b
22
10.2, 20.6
3.1
10.7

(i) intermediate gold(I) species involved in the surface reaction with oxygen and cyanide
(ii) a reaction mechanism involving the simultaneous reaction of gold and oxygen on gold
(iii) a rate constant for the intrinsic surface reaction.

a
At 25 8C, Hogfeldt (1982); Sillen and Martell (1964); Kissner
et al. (1997); Stefansson and Seward, 2003; Nicol et al., 1987.
b
See text.

example, the stability constant b{Ag(CN)(OH)} =


1012.8 corresponds to E8{Ag(CN)(OH)/Ag} = 0.044
V, based on E8{Ag + /Ag} = 0.799 V at 25 8C and
Eq. (20).
E8fMCNOH =MgE8fM=Mg

 0:059 log bfMCNOH g
20

3. Gold(I) speciation: justification and importance


The reported evidence for intermediate silver(I) and
gold(I) complexes such as Ag(OH)0, Ag(OH)2,
Ag(CN)(OH), Au(OH)0, Au(CH3CN)(OH)0, Au(CH3
CN)2+ and Au(OH)2 (Table 1) shows the possible formation of intermediates such as Au(OH)0 and
Au(CN)(OH) in addition to Au(CN)0 in surface
reactions during cyanide leaching of gold. Previous
studies (Finkelstein and Hancock, 1974; Senanayake
et al., 2003) highlighted problems associated with
measuring the stability constants of gold(I) complexes and showed the importance of the following
relationship between the standard reduction potentials of Au(I)/Au(0) and Ag(I)/Ag(0) redox couples:
E o fAuI=Au0g 1:79E8fAgI=Ag0g 0:236
19
This relationship can be used to predict the stability
constants of unstable gold(I) complexes by using the
published (Hogfeldt, 1982; Sillen and Martell, 1964)
values of stability constants of Ag(I) (Table 1). For

For gold, this corresponds to E8{Au(CN)(OH)/


Au} = 0.314 V and b{Au(CN)(OH)} = 1023.3 based
on Eqs. (19) and (20) and E8{Au + /Au} = 1.69 V at
25 8C. The value of 1023.3 for b{Au(CN)(OH)} is of
the same order as one of the values (1020.6) reported
for b{Au(OH)0}. However, both values are fifteen orders of magnitude smaller than 1038 for b{Au(CN)2}
(Table 1). Of the two values reported for b{Au(OH)0}
in the literature (1010.2, 1020.6), the lower value shows
a better fit to the linear relationship with stability
constants of other gold(I) complexes (Senanayake,
2004). Thus, Au(OH)(CN) would seem to be a
more plausible intermediate compared to Au(OH)0.
In the case of thiosulphate leaching of gold, the
intermediate complex Au(NH3)(S2O3) has a stability
constant of 1020 compared to 1024 for the stable complex Au(S2O3)23. Yet, the rate of anodic oxidation of
gold in ammoniacal thiosulphate can be modelled on
the basis of the formation of Au(NH3)(S2O3) as an
intermediate (Senanayake, 2005). Kissner et al. (1997)
noted that the high stability of Au(CN)2 is due to the
moderate basicity, minor hardness and k-acceptor capability of cyanide ligand, while hydroxide, which is a

G. Senanayake / Hydrometallurgy 80 (2005) 112

Au(OH)(CN) may be formed as an ionic intermediate


on the surface, it would not be an insoluble film, but
would readily desorb/dissolve as the more stable complex Au(CN)2. These results support the formation of
an insoluble film such as Au(OH), in addition to AuCN,
in dilute cyanide solution and highlights the need to
consider surface reactions on the basis of such species.

hard and strongly basic ligand, is a good donor for Au(I).


They also noted that OH and NH3 are basic and strong
j-bond donors which should coordinate in the same
way with Au(I). These views are supported by the
stability constants reported in Table 1. Thus, it is reasonable to analyse the rate of gold oxidation on the
basis of formation of reaction intermediates such as
Au(OH)0 and Au(CN)(OH) adsorbed onto the gold
surface. However, they are eventually converted to the
more stable Au(CN)2 with a higher stability constant
of 1038.
Support for such adsorbed species also come from
electrochemical/nanobalance studies. Jeffrey and
Ritchie (2001) measured the anodic polarization
curve for pure gold in a pure 20 mM cyanide solution
maintained at 25 8C. The measured current density of
0.03 A m 2 in the potential region for leaching ( 0.1
V) was independent of stirring rate. Assuming that the
measured current was due to the oxidation of gold to
AuCN, they calculated a mass increase of 650 ng in a
40 min scan. This was close to the measured mass
increase of 605 ng of the rotating gold disc. However,
Fig. 1 shows these two values as well as the predicted
mass increase on the basis of formation of other gold(I)
species shown in Table 1 and other intermediates such
as Au2O (Bard, 1973). It is clear that the measured
increase in mass is much closer to the predicted values
on the basis of solids such as Au(OH)0 + Au(CN)0 or
Au(OH)0 than that based on Au(CN)0 alone. Although

4. Analysis of rate data


4.1. Levich equation
The Levich (1962) equation (Eqs. (21)) has been
widely used for the interpretation of rate data based
on the electrochemical or chemical dissolution of
metal from a rotating disk, under diffusion controlled
conditions:
2=3

J X 0:62DX x1=2 t 1=6 X 

21

where, J X = flux of reactant (mol m 2 s 1); x = rotation


rate of the disc = rpm. 2k / 60 (s 1); t = kinematic viscosity (0.89.10 6 m2 s 1 for water); D X = diffusion coefficient of X(m2 s 1); [X] = concentration of X(mol m 3
or mM).
Thus, the pseudo first order dependence of the rate
of gold dissolution with respect to cyanide concentration and oxygen pressure could be the result of a

Measured increase in mass

Au2O

Au(OH)

Au(CN) +Au(OH)

Au(CN)

Au(CNO)

Au(OH)
+Au(OH)(CN)

700

Au(OH)(CN)

Predicted increase in mass / ng

800

600

500

Predicted Gold(I) species formed on anode


Fig. 1. Comparison between measured and predicted increase in mass of a gold anode due to passivation by different gold(I) species (see text).

G. Senanayake / Hydrometallurgy 80 (2005) 112

chemically or diffusion controlled surface reaction


(Sun et al., 1996; Guan and Han, 1994; Choi et al.,
1991). For example, at steady state, the rate of diffusion is equal to the rate of surface reaction (Levenspiel, 1972). Thus, the Levich equation given in the
form of Eqs. (22) can be used to compare the rate data
obtained under diffusion controlled conditions, where
J Au(I) represents the flux in each case with m being the
relevant stoichiometric factor that should satisfy the
overall mass balance. For example, the relevant values
in the case of Eqs. (1) are m = 2 for CN and 0.5 for
O2 whereas for Eqs. (3), the relevant values are m = 2
for CN and 0.25 for O2.

tion of cyanide was maintained constant at 20 mM.


The rate of gold dissolution was determined by measuring the loss of mass with time. Wadsworth et al.
(2000) used a rotating gold disc in cyanide solutions
maintained at pH = 10.5 and 300 rpm. The progress of
reaction was measured by determining dissolved
Au(I) in solution using ICP analysis. The results
were in excellent agreement with the results reported
by Jeffrey and Ritchie (2001). However, since the pH
used by the two groups were different, the concentration of free cyanide [CN]free was calculated in the
present study using [CN]total and pK a (HCN). A plot
of [CN]free vs. [CN]total gave a slope of 0.9 confirming that only a small fraction of cyanide was in
the form of HCN.
Fig. 2 shows a loglog plot of R Au vs. [CN]free.
The slope at low cyanide concentration is 2.7. Therefore, it is clear that rates at low cyanide concentrations
do not follow Eq. (22). This indicates that the rate is
controlled by a surface chemical reaction that involves
two to three cyanide ions. This behaviour of pure gold
in pure dilute cyanide solutions is different from
results reported previously which indicate that gold
cyanidation was generally first order with respect to
cyanide concentration, oxygen partial pressure and
square root of disc rotating speed with a low activation energy E a = 2229 kJ mol 1 (Sun et al., 1996;
Guan and Han, 1994; Choi et al., 1991). By contrast,

logfRAuI g logfJ AuI g


logf0:62mt 1=6 DX 2=3 x1=2g logX 
22
4.2. Rate data
Jeffrey and Ritchie (2001) used a rotating electrode
quartz crystal microbalance to determine the progress
of cyanidation of freshly plated gold at 25 8C, in
aerated solutions of different cyanide concentrations
(pH c 10) at a disc rotation speed of 300 rpm (x 1/2 =
5.61 s 1/2). They have also reported results obtained
using 4 different rotation rates, while the concentra-

-4.5

{log RAu /mol m-2 s-1}

Jeffrey and Ritchie, 2001


Wadsworth et al., 2000
-5.5

-6.5

y = 2.74x - 7.37
R2 = 1.00

-7.5
0

0.4

0.8
-

1.2

1.6

-3

log {[CN ]free / mol m }


Fig. 2. Loglog plots of gold oxidation rate vs. [CN]free at 25 8C. Data from Jeffrey and Ritchie, 2001 (pH = 10) and Wadsworth et al., 2000
(pH 10.5).

G. Senanayake / Hydrometallurgy 80 (2005) 112

the higher activation energy of E a = 4755 kJ mol 1


(Thurgood et al., 1981) for oxidation of a pure gold
anode in cyanide solutions, or E a = 47 F 2 kJ mol 1
for pure gold cyanidation by oxygen (Jeffrey and
Ritchie, 2001), as in the present example, supports a
chemically controlled reaction.
The value of R Au reaches a limiting value R Au(lim)
at higher concentrations of [CN] (Fig. 2). According
to the mixed potential theory, the limiting rate at
higher cyanide concentrations is a result of the reaction being controlled by oxygen diffusion to the interface, so that the anodic dissolution rate of gold is
matched by the cathodic reduction rate of oxygen.
Thus, the oxygen diffusion, which is enhanced at
higher rotation rates, should result in an increase in
R Au with increasing x 1/2. This is not observed, as
noted by Jeffrey and Ritchie (2001), and the limiting
rate remains independent of rotation rate. Therefore, it

II

III

OH2
Au

IV

Au

5. Reaction mechanism
It is possible to consider the chemical dissolution of
gold in oxygenated cyanide solutions as a reaction that
involves the simultaneous reduction of oxygen and
oxidation of gold as shown in Fig. 3. The three cases
shown in Fig. 3 consider the involvement of 2 gold
atoms per oxygen molecule and (a) reaction without
the involvement of cyanide producing Au(OH)0, (b)
reaction with only one cyanide ion producing
Au(OH)0 + Au(OH)(CN), and (c) reaction with two
cyanide ions producing Au(OH)(CN). Eqs. (23)(26)
consider only the case described in Fig 3b because it
is representative of the other two extreme cases.

Au(OH)0

OH2 O
OH2 O

is important to consider a reaction mechanism that can


describe this behaviour of gold.

OH

OH

(a) cyanide not involved

OH

OH

(b) cyanide involved in one site

OH

OH

(c) cyanide involved in both sites

Au(OH)0

OH2

CNAu(OH)(CN)-

Au
OH2 O
OH2 O
Au

Au(OH)
OH2

CNAu(OH)(CN)-

Au
OH2 O
OH2 O
Au

Au(OH)(CN)-

CNFig. 3. Formation of Au(I) intermediates in surface reaction with O2 and (a) 0; (b) 1; (c) 2 CN ions (I) gold surface, (II) adsorbed water and/or
cyanide, (III) oxygen, (IV) adsorbed Au(I) species after reaction, (V) hydrogen peroxide.

G. Senanayake / Hydrometallurgy 80 (2005) 112

Surface equilibrationredox reaction


2pAus 2H2 O CN O2


2pAuH2 O:CN :O2 ads

23

2pAuH2 O:CN :O2 ads

h K ads O2 CN n =1 K ads O2 CN n

30

RAu k Au 

31

RAu k Au K ads O2 CN n =f1 K ads O2 CN n g


from Eqs:30; 31:

32

pAuOHads pAuOHCN ads H2 O2


24

At lower values of [O2][CN]n


f1 K ads O2 CN n g c 1

Desorption/stabilisation

pAuOHads CN pAuOHCN
ads=aq OH

25
pAuOHCN ads=aq CN p AuCN
2
OH :

26

The equilibration shown in Eq. (23) can be considered as the formation of a surface adsorbed transition state followed by the redox reaction which
produces the adsorbed gold(I) species and hydrogen
peroxide. The two unstable species Au(OH)0 and
Au(OH)(CN) formed on the surface react with cyanide to produce more stable Au(CN)2 in solution as
shown in Eqs. (25)(26), while hydrogen peroxide
can react in the manner described before. The reaction
model in Fig. 3 shows a maximum of 2CN ions
involved in the surface reaction mechanism. It is
important to rationalise the reaction order of 23
demonstrated by the slope of linear relationship in
Fig. 2 at low cyanide concentrations. Thus, a general
form of the surface reaction is shown in Eqs. (27)
(28), which involves nCN ions. The equilibrium
constant for Eqs. (23), surface coverage (h), and the
rate expression are given by Eqs. (29)(31).
General surface reaction
2Au 2H2 O nCN O2
Au:H2 O2 :CN n :O2

27

Au:H2 O2 :CN n :O2 Y Products

28

K ads h=f1  hO2 CN n g for Eq: 27

29

33

Then,
RAu k Au K ads O2 CN n from Eq: 32:

34

General rate equation


CN n k Au K ads O2 RAu  1  K ads O2 
from Eq: 32:

35

In the case of solutions of low cyanide concentrations, Eq. (32) simplifies to Eq. (34). This explains the
value of n between 2 and 3 obtained as the reaction
order with respect to cyanide at low concentrations
(Fig. 2). Moreover, Eq. (32) can be rearranged to Eq.
(35) that can be used to plot [CN] n vs. (R Au) 1 to
obtain values for K ads and k Au. Fig. 4a shows the
relevant plots for n = 1 and 2 while Fig. 4b and c
represent n = 3 and n = 4 respectively. In all cases,
the dissolved oxygen concentration can be considered
as c0.25 mM, because the experiments were carried
out in air saturated solutions. The best linear relationship is given by n = 3 (Fig. 4b) with slope of 3  10 8
and intercept in the range  3.3  10 3 to  3.7 
10 3 for the two data sets reported by Jeffrey and
Ritchie (2001) and Wadsworth et al. (2000). The value
of k Au = slope/(-intercept) = 8.6  10 6 mol m 2 s 1
based on the average intercept is close to the limiting
rate of R Au(lim) = 7.3  10 6 mol m 2 s 1 calculated
from the plateau in Fig. 2 described in Section 4.
Thus, the limiting rate of gold cyanidation in pure
impurity-free solutions at high cyanide concentrations is approximately equal to the intrinsic rate
constant. The slope showing a reaction order of 2.7
indicates the predominant reactions represented by
Eqs. (36)(38), which involve 2 or 3 cyanide ions in

10

G. Senanayake / Hydrometallurgy 80 (2005) 112

(a)

0.6

{[CN-]free}-n

Solid lines: Jeffrey and Ritchie, 2001


Dashed lines: Wadsworth et al., 2000
n=1

0.4

0.2

n=2
0
0.E+00

2.E+06

4.E+06

6.E+06

(RAu)-1

(b)
0.2

{[CN-]free}-n

Solid line: Jeffrey and Ritchie, 2001


Dashed line: Wadsworth et al., 2000

y = 3E-08x - 0.0033
R2 = 0.9999

n=3
y = 3E-08x - 0.0037
R2 = 0.9998

0
0.E+00

2.E+06

4.E+06

6.E+06

4.E+06

6.E+06

(RAu)-1

(c)
0.05

{[CN-]free}-n

0.04

Solid line: Jeffrey and Ritchie, 2001


Dashed line: Wadsworth et al., 2000

0.03

n=4
0.02
0.01
0
0.E+00

2.E+06
(RAu)-1

Fig. 4. Plot of {[CN]free} n vs. (R Au) 1 using data from Fig. 2 to examine the validity of Eq. (35). (a) n = 1 or 2, (b) n = 3, (c) n = 4.

G. Senanayake / Hydrometallurgy 80 (2005) 112

the transition state. The intermediate Au(OH)(CN)


formed at the interface will rapidly react with cyanide to produce the most stable Au(CN)2.

11

! At higher cyanide concentrations, the reaction rate


approaches a limiting value of R Au(lim) = 7.2  10 6
mol m 2 s 1 which is close to the intrinsic rate
constant k Au = 8.6  10 6 mol m 2 s 1.

Au:H2 O2 :CN 2 :O2 2AuOHCN H2 O2


36
Au:H2 O2 :CN 3 :O2

AuCN
2

AuOHCN H2 O2 OH
37

Au:H2 O2 :CN 3 :O2 2AuOHCN



CNO H2 O

38

Wadsworth et al. (2000) reported the two values for


the terms k aU tot = 6.9  10 6 and K = 5.3  10 3 in
Eq. (17), based on an electrochemical model which
involved the anodic oxidation of gold and cathodic
reaction of oxygen via reactions in Eqs. (11)(16).
The slight differences in rate constants determined in
this work (8.6  10 6 mol m 2 s 1) and from their
model (6.9  10 6 mol m 2 s 1) may be partly attributed to the use of free cyanide concentration in the
present analysis. Nevertheless, the good agreement in
rate constants supports the validity of both models and
highlights the possibility of extending the surface
chemical model to rationalise the cyanidation kinetics
of silver and gold alloys, which will be presented in
future communications.

6. Summary and conclusions


! The rate of chemical dissolution of gold in pure
cyanide solutions is controlled by the surface chemical reaction between gold, cyanide and oxygen
via a transition state (Au.H2O)2.(CN)23.(O2).
! The simultaneous reaction produces the surface
intermediate Au(OH)(CN) while oxygen is reduced to hydrogen peroxide.
! The intermediate Au(OH)(CN) reacts with cyanide
ions to produce more stable Au(CN)2 in solution.
Hydrogen peroxide may react with gold or cyanide,
or disproportionate to oxygen and water.
! The reaction order for cyanidation at low cyanide
concentrations is close to 3.

References
Bard, A.J., 1973. Encyclopedia of Electrochemistry of the Elements,
vol. IV. Marcel Dekker, New York.
Bodlander, G., 1896. Die chemie des cyanidverfahrens. Z. Angew.
Chem. 9, 583 587.
Breuer, P.L., Dai, X., Jeffrey, M.I., 2005. Leaching of gold and
copper minerals in cyanide deficient copper solutions. Hydrometallurgy 78, 156 165.
Cathro, K.J., 1963. The effect of oxygen in the cyanide process for
gold recovery. Proc. Aus. I.M.M., vol. 207, pp. 181 205.
Cathro, K.J., Koch, D.F.A., 1964. The dissolution of gold in cyanide
solutions. Proc. Aus. I.M.M., pp. 111 126.
Choi, Y.U., Lee, E.C., Han, K.N., 1991. The dissolution behaviour
of metals from Ag/Cu and Ag/Au alloys in acidic and cyanide
solutions. Metall. Trans. 22B, 755 764.
Cornejo, L.M., Spottiswood, D.J., 1984. Fundamental aspects of the
gold cyanidation process: a review. Miner. Energy Resour. 27,
1 18.
Crundwell, F.K., Godorr, S.A., 1997. A mathematical model of the
leaching of gold in cyanide solutions. Hydrometallurgy 44,
147 162.
Dorin, R., Woods, R., 1991. Determination of leaching rates of
precious metals by electrochemical techniques. J. Appl. Electrochem. 21, 419 424.
ber das verhalten verschiedener metalle in
Elsner, L., 1846. U
einer wassrigen losung von zyankalium. J. Prakt. Chem. 37,
441 446.
Finkelstein, N.P., Hancock, R.D., 1974. A new approach to the
chemistry of gold. Gold Bull. 7, 72 77.
Fleming, C., 1992. Hydrometallurgy of precious metals recovery.
Hydrometallurgy 30, 127 162.
Guan, Y.C., Han, K.N., 1994. An electrochemical study on the
dissolution of gold and copper from gold copper alloys. Metall.
Mater. Trans. 25B, 817 827.
Hiskey, J.B., Sanchez, V.M., 1990. Mechanistic and kinetic aspects
of silver dissolution in cyanide solutions. J. Appl. Electrochem.
20, 479 487.
Hogfeldt, E., 1982. Stability Constants of MetalIon Complexes,
2nd Supplement, IUPAC Chemical Data Series No. 2, Part A,
Inorganic Ligands. Pergamon, Oxford.
Jeffrey, M.I., Ritchie, I.M., 2000a. The leaching of gold in cyanide
solutions in the presence of impurities: I. The effect of lead.
J. Electrochem. Soc. 147, 3257 3262.
Jeffrey, M.I., Ritchie, I.M., 2000b. The leaching of gold in cyanide
solutions in the presence of impurities: II. The effect of silver.
J. Electrochem. Soc. 147, 3272 3276.
Jeffrey, M.I., Ritchie, I.M., 2001. The leaching and electrochemistry
of gold in high purity cyanide solutions. J. Electrochem. Soc.
148, D29 D36.

12

G. Senanayake / Hydrometallurgy 80 (2005) 112

Juttner, K., 1984. Oxygen reduction electrocatalysis by underpotential deposited metal atoms at different single crystal
faces of gold and silver. Electrochim. Acta 29, 1597 1604.
Kirk, D.W., Foulkes, F.R., 1978. A study of anodic dissolution of
gold in aqueous alkaline cyanide solutions. J. Electrochem. Soc.
125, 1436 1443.
Kirk, D.W., Foulkes, F.R., Graydon, W.F., 1980. Gold passivation in aqueous alkaline cyanide. J. Electrochem. Soc. 127,
1962 1969.
Kissner, R., Welti, G., Geier, G., 1997. The hydrolysis of gold(I) in
aqueous acetonitrile solutions. J. Chem. Soc. Dalton Trans.,
1773 1777.
Kudryk, V., Kellogg, H.H., 1954. Mechanism and rate controlling
factors in the dissolution of gold in cyanide solutions. J. Met. 6,
541 548.
Levenspiel, O., 1972. Chemical Reactions Engineering. Wiley,
New York.
Levich, V.G., 1962. Physico-Chemical Hydrodynamics. Prentice
Hall, Englewood Cliffs, NJ.
Li, J., Zhong, T., Wadsworth, M.E., 1992. Application of
mixed potential theory in hydrometallurgy. Hydrometallurgy
29, 47 60.
Lorenzen, L., van Deventer, J.S.J., 1992. Electrochemical interactions between gold and its associated minerals during cyanidation. Hydrometallurgy 30, 177 194.
MacArthur, D.M., 1972. A study of gold reduction and oxidation in
aqueous media. J. Electrochem. Soc. 119, 672 677.
Nicol, M.J., 1980. The anodic behaviour of gold. Gold Bull. 13,
105 111.
Nicol, M.J., Fleming, C.A., Paul, R.L., 1987. The chemistry of the
extraction of gold. In: Stanley, G.G. (Ed.), The Extractive
Metallurgy of Gold, vol. 2. S. African Inst. Min. Metall.,
Johannesburg, pp. 831 905.
Osseo-Asare, K., Xue, T., Ciminelli, V.S.T., 1984. Solution chemistry of cyanide leaching systems. In: Kudryk, V., Corrigan,
D.A., Liang, W. (Eds.), Precious Metals: Mining, Extraction
and Processing. Metall. Soc. AIME, Warrendale, pp. 173 197.
Senanayake, G., 2004. Fundamentals and applications of metal
ligand complexes of gold(I/III) in non-cyanide gold pro-

cesses. Green Processing 2004. Aus. I.M.M., Melbourne,


pp. 113 122.
Senanayake, G., 2005. Catalytic role of ammonia in the anodic
oxidation of gold in copper-free thiosulfate solutions. Hydrometallurgy 77, 287 293.
Senanayake, G., Perera, W.N., Nicol, M.J., 2003. Thermodynamic
studies of the gold(III)/(I)/(0) redox system in ammonia-thiosulphate solutions at 25 8C. In: Young, C.A., Alfantazi, A.M.,
Anderson, C.G., Dreisinger, D.B., Harris, B., James, A.
(Eds.), Hydrometallurgy 2003, Leaching and Solution Purification, vol. 1. TMS, Warrendale, pp. 155 168.
Sillen, L.G., Martell, E., 1964. Stability constants of metalion
complexes. Special Publication, vol. 17 and 26. Chemical
Society, London.
Stefansson, A., Seward, T.M., 2003. The hydrolysis of gold(I) in
aqueous solutions to 600 8C and 1500 bar. Geochim. Cosmochim. Acta 67, 1677 1688 (and references therein).
Sun, X., Guan, Y.C., Han, K.N., 1996. Electrochemical behaviour
of the dissolution of goldsilver alloys in cyanide solutions.
Metall. Mater. Trans. 3B, 355 361.
Thurgood, C.P., Kirk, D.W., Foulkes, F.R., Graydon, W.F., 1981.
Activation energies of anodic gold reactions in aqueous alkaline
cyanide. J. Electrochem. Soc. 128, 1680 1685.
Wadsworth, M.E., Zhu, X., 2003. Kinetics of enhanced gold dissolution: activation by dissolved silver. Int. J. Miner. Process.
72, 301 310.
Wadsworth, M.E., Zhu, X., Thompson, J.S., Pereira, C.J., 2000.
Gold dissolution and activation in cyanide solution: kinetics and
mechanism. Hydrometallurgy 57, 1 11.
Xue, T., Osseo-Asare, K., 2001. Anodic behaviour of gold, silver,
and goldsilver alloys in aqueous cyanide solutions. In:
Young, C.A., Twidwell, L.G., Anderson, C.G. (Eds.), Cyanide:
Social, Industrial and Economic Aspects. TMS, Warrendale,
pp. 563 576.
Zheng, J., Ritchie, I.M., La Brooy, S.R., Singh, P., 1995. Study of
gold leaching in oxygenated solutions containing cyanidecopperammoniausing a rotating quartz crystal microbalance.
Hydrometallurgy 39, 277 292.

Você também pode gostar