Vous êtes sur la page 1sur 11

Hydrometallurgy 81 (2006) 75 85

www.elsevier.com/locate/hydromet

The cyanidation of silver metal: Review of kinetics


and reaction mechanism
Gamini Senanayake
Parker Centre, Department of Mineral Science and Extractive Metallurgy, Murdoch University, Perth, WA 6150, Australia
Received 20 September 2005; received in revised form 2 November 2005; accepted 2 December 2005

Abstract
Published rate data are analysed for the chemical and electrochemical dissolution of silver metal from rotating discs in aerated/
oxygenated cyanide solutions at 25 C, pH 11 and different partial pressures of oxygen. The current status of the reaction
mechanism is also reviewed. Speciation analysis of 0.01 mM silver(I) in 1100 mM cyanide solutions shows that Ag(CN)2
is the predominant complex (50%) at cyanide concentrations b 20 mM. However, at higher cyanide concentrations, Ag(CN)2
3
(up to 60%) and Ag(CN)3
4 (up to 10%) can be formed. Thus, it is important to consider a silver(I) : cyanide ion ratio of 2 or
3 in the Levich equation to calculate the diffusion coefficient of cyanide ion. Likewise, it is important to consider a silver(I) :
oxygen ratio of 1 : 0.5 to calculate the diffusion coefficient of oxygen. This indicates the reduction of oxygen to hydrogen
peroxide in the surface reaction. Analysis of exchange current density data for silver oxidation as a function of cyanide
concentration shows the involvement of between 1 and 2 cyanide ions in the surface reaction. The limiting rate of silver
dissolution at high cyanide concentrations (2.5 10 5 mol m 2 s 1 at 21 kPa oxygen pressure) represents the maximum
surface coverage by cyanide. This value is in close agreement with the rate constant of the surface reaction 4 10 5 mol m 2
s 1 based on the pure kinetic current of the mixed charge transfer plus diffusion model proposed by Li and Wadsworth
[Li, J., Wadsworth, M.E., 1993. Electrochemical study of silver dissolution in cyanide solutions. J. Electrochem. Soc. 140,
19211927].
2005 Elsevier B.V. All rights reserved.
Keywords: Silver cyanidation; Surface adsorption; Diffusion; Kinetics; Reaction mechanism

1. Introduction

cyanate produced in solution (Eq. (4)) further degrades


to other products (Eq. (5), Arikado et al., 1976).

Silver cyanidation has been reported to involve the


chemical reactions shown in Eqs. (1)(3), that take the
same form as equations for gold cyanidation (Deitz and
Halpern, 1953; Bodlnder, 1896; Elsner, 1846). The

2Ag 4CN O2 2H2 O


2AgCN2 2OH H2 O2

2Ag 4CN H2 O2

2AgCN2

4Ag 8CN O2 2H2 O

1
2OH

4AgCN2

4OH

3
Tel.: +61 8 93602833; fax: +61 8 93606343.
E-mail address: G.Senanayake@murdoch.edu.au.
0304-386X/$ - see front matter 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.hydromet.2005.12.001

CN H2 O2 CNO H2 O

CNO 2H2 O

NH
4

CO2
3

4
5

76

G. Senanayake / Hydrometallurgy 81 (2006) 7585

The cyanidation process is widely used for gold


extraction from ores and concentrates. Silver co-exists in
most gold ores and is co-extracted during gold cyanidation. Scrap alloys from various sources such as
discarded/recycled electronic materials also make an
important contribution to the supply of gold, silver and
other metals. Thus, the co-extraction of silver during
gold cyanidation is an economical advantage. However,
silver can be beneficial or detrimental to the kinetics of
gold cyanidation depending on the mineralogy and the
mode of occurrence.
Hiskey and Sanchez (1990) reviewed and compared
the published data on relative rates of gold and silver
cyanidation, pre-dated to pioneering work by Kudryk
and Kellogg (1954). Previous results have shown that
silver dissolution lags behind that of gold (Hiskey and
Sanchez, 1990). Thus, increased attention has been paid
to the mutual effects of alloyed metals or aqueous metal
ions on the kinetics of cyanidation of gold, silver and
copper (Choi et al., 1991; Sun et al., 1996; Jeffrey and
Ritchie, 2000; Xue and Osseo-Asare, 2001; Wadsworth
and Zhu, 2003; Breuer et al., 2005).
Li et al. (1992) showed that there is excellent
agreement between the rates obtained from direct
leaching with the predicted values using the mixed
potential theory and polarisation curves of silver
oxidation and oxygen reduction. Despite this success
the mechanistic detail of the role of silver in gold
cyanidation is still under investigation (Xue and OsseoAsare, 2001; Wadsworth and Zhu, 2003). The aim of
this paper is to:
(i) review, compare, and contrast the rate data and
reaction mechanisms of silver cyanidation
reported by previous researchers,
(ii) highlight the applicability (and limitations) of the
diffusion models to rationalise the dissolution
kinetics and the number of cyanide ions involved
in the surface reaction, and
(iii) determine a rate constant for the surface reaction
that can be used to compare the kinetics of silver
cyanidation by the two methods: chemical or
electrochemical.
2. Silver(I) speciation
The formation of Ag(OH)0 and Ag(CN)0 on the
silver surface during the anodic oxidation of silver has
been reported by Xue and Osseo-Asare (2001). The
mixed complex Ag(I)(OH)(CN) has a stability constant
which is ten orders of magnitude larger than that of Ag
(OH)0 (Table 1). Yet, the two complexes are less stable

Table 1
Stability constants of silver(I) complexes
Complex

Ionic strength

log

Ag(CN)2

1(NaClO4)
1(NaClO4)
0
0 (or dil)

20.1
12.8
3.6 (4.2 at 18 C, 0.2 KNO3)
2.3, 3.9

Ag(OH)(CN)
Ag(OH)2
Ag(OH)0

At 25 C, Hogfeldt (1982), Sillen and Martell (1964); the two values


2.3 and 3.9 for Ag(OH)0 are based on solubility studies reported by
two research groups showing the difficulties of measuring these values
due to the equilibrium between Ag2O(s) and Ag(OH)(s).

compared to Ag(CN)2. Thus, the published EHpH


diagrams by Xue and Osseo-Asare (1985) have shown
Ag(CN)2 as the predominant complex at low concentration ratios of CCN / CAg(I). The species distribution
diagram in Fig. 1a shows trends consistent with the EH
pH diagrams (Xue and Osseo-Asare, 1985, 2001),
indicating that the fraction of silver(I) which exists as
Ag(CN)32 and Ag(CN)43 increases with the increase in
ratio of CCN / CAg(I). Based on the Nernst equation and
measured Ag(I) / Ag electrode potentials, Li and Wadsworth (1993) reported a coordination number n 2 for
the complex Ag(CN)n(n1) at cyanide concentrations
b10 mM in solutions maintained at a wide range of
temperatures (060 C).
It is clear from Fig. 1a that in a solution which
contains 1 mg L1 Ag(I) (0.01 mM) at pH 11 the
percent composition of Ag(CN)2 decreases with
increasing CCN / CAg(I) ratio, while Ag(CN)32 becomes
the predominant complex at a total cyanide concentration greater than 20 mM. The composition of species
such as Ag(OH)0, Ag(OH)2 and Ag(OH)(CN) appears
to be negligibly small compared to the other three
complexes (Fig. 1a). Surprisingly, the concentration of
Ag(OH)0 is even less than that of Ag+ in alkaline
cyanide solution of pH 11 (Fig. 1b). The order of
stability shown in Fig. 1ab remains the same at pH 10.
The mixed complex Ag(OH)(CN) appears to be the
most predominant species in Fig. 1b, as expected from
the higher stability constant reported in Table 1.
3. Measurement of rates
Table 2 lists the conditions and experimental methods
used by previous researchers to study the rate (RAg, mol
m 2 s 1) of silver cyanidation. Deitz and Halpern (1953)
used a rotating silver disc and monitored the progress of
the reaction and rate by analysing the solutions for silver
(I) using the mixed-colour dithizone method. Choi et al.
(1991) used a rotating silver disc and monitored the
progress of dissolution by analyzing the solutions for
silver(I) using atomic absorption spectrophotometry.

G. Senanayake / Hydrometallurgy 81 (2006) 7585

Concentration of species (%)


Concentration of species /mM

6.E+03

100
CCN/ CAg(I)total

80

4.E+03

60

Ag(CN)32-

40

Ag(CN)2-

20
0

2.E+03
Ag(OH)(CN)Ag(OH)2-

3-

Ag(CN)4
0

10

20

40
30
CCN / mM

50

CCN / CAg(I)total

77

60

Ag(OH)0
0.E+00

1E-11

Ag(OH)(CN)-

1E-16

1E-21

1E-26

Ag+
Ag(OH)0
Ag(OH)2-

10

20

30
CCN / mM

40

50

60

Fig. 1. Distribution of silver(I) species in cyanide solutions at 25 C: (a) stable species, (b) unstable species.

Breuer et al. (2005) used a rotating electrode quartz


crystal microbalance which allowed the measurement of
anodic current as well as the rate of chemical dissolution
of freshly plated silver by monitoring the loss of mass
with time in air saturated solutions.
Hiskey and Sanchez (1990), Li et al. (1992) and Sun
et al. (1996) reported the anodic (silver) and cathodic
(oxygen on silver) polarisation curves of rotating silver

disc electrodes. This allowed the calculation of rates


using the limiting current density (ilim, A m 2) and/or
corrosion current density at mixed potentials (imix at
Emix) using the relationship RAg = i / zF (i = ilim or imix,
z = 1, F = Faraday constant = 96487 C mol 1). The term
imix represents the current density (A m 2) at the point
of intersection of the cathodic and anodic polarisation
curves. The anodic and cathodic branches during the

Table 2
Test conditions used to measure rate of silver cyanidation
Set

Method a

CNaCN (mM)

pH

T C

rpm

Reference

A
B
C
D
E
F
G
H
I
J

CD
icorr
CD
icorr
icorr
REQCM
AP
AP
AP
AP

265
0.510 c
0.410 c
0.540 d
40 d
20
120 e
10 e
0.510 f
0.540 d

25
24
24
25
25
25
23
23
24
25

895
100
500
450
4501100
300
500
20700
300
450

Deitz and Halpern, 1953


Li et al., 1992
Li et al., 1992
Sun et al., 1996
Sun et al., 1996
Breuer et al., 2005
Hiskey and Sanchez, 1990
Hiskey and Sanchez, 1990
Li and Wadsworth, 1993
Sun et al., 1996

11
11
11
11
10
10.210.9
10.210.9
11
11

See text for oxygen levels.


a
CD: chemical dissolution of rotating discs; icorr: corrosion currents of rotating disc electrodes; REQCM: rotating electrode quartz crystal
microbalance; AP: anodic polarisation of a rotating silver disc.
b
Not reported.
c
0.5 M KNO3.
d
0.5 M Na2SO4.
e
Ionic strength adjusted to 0.6 M Na2SO4.
f
0.5 M Na2SO4.

78

G. Senanayake / Hydrometallurgy 81 (2006) 7585

oxidation of a metal approach a plateau (ilim) at higher


potentials causing the diffusion of ligand (cyanide) or
oxidant (oxygen) to be the rate controlling step
depending on the relative concentrations of the two
reagents (Li et al., 1992). The following section
describes the use of such information to compare and
contrast literature data.
4. Interpretation of rates based on diffusion model

The free cyanide concentration denoted by CCN in


Eq. (8) depends on total cyanide, pH, pKa(HCN),
temperature and ionic strength. The calculated values of
CCN when plotted against total cyanide concentration
give a linear relationship of slope 0.98, indicating that
the fraction of HCN in these solutions was negligibly
small at pH 11. Thus, Eqs. (8) and (9) provide a means
of comparing the results reported by Deitz and Halpern
(1953), Hiskey and Sanchez (1990), Li and Wadsworth
(1993), and Sun et al. (1996).

4.1. Levich equation


4.2. Anodic oxidation of silver
The Levich equation (Eq. (6)) has been widely used
for the interpretation of rate data based on the
electrochemical or chemical dissolution of metal from
a rotating disc under diffusion controlled conditions:
JX 0:62 DX x1=2 y1=6 CX
2=3

6
2

where JX = flux of reactant (mol m s ), = rotation


rate of the disc = rpm 2 / 60 (s 1 ), = kinematic
viscosity (m2 s 1), DX = diffusion coefficient of X (m2
s 1), and CX = concentration of X (mol m 3 or mM).
At low concentration of cyanide, the rate of diffusion
of cyanide is equal to the rate of surface chemical
reaction representing the steady state (Levenspiel,
1972). At higher concentrations of cyanide, representing
excess cyanide at the surface, the limiting rate is equal to
the rate of diffusion of oxygen. Thus Eq. (7) shows the
relationship between the limiting current and oxygen
flux (JO2). The values of p are 2 and 4, respectively, for
the 2e reduction of O2 to H2O2, and 4e reduction of O2
to H2O in Eqs. (1) and (3), respectively.
JO2 ilim =pF 0:62 y1=6 x1=2 DO2 2=3 CO2

RAg ilim =F JAgI JCN =n


0:62n1 y1=6 x1=2 DCN 2=3 CCN

RAg JAgI JO2 =m


0:62 m1 y1=6 x1=2 DO2 2=3 CO2

Eqs. (8) and (9) show the Levich equation which


takes into account the relevant stoichiometric ratios of
Ag : CN : O2 = 1 : 2 : 0.5 and 1 : 2 : 0.25, respectively, for
the overall reactions in Eqs. (1) and (3). According to
these two equations, in which Ag(CN)2 is the silver(I)
complex, the cyanide flux should be twice that of silver
(I) corresponding to a value of n = 2 in Eq. (8). In the
case of Ag(CN)32 and Ag(CN)43 the value of n should
be 3 and 4, respectively. Likewise the values of m for
oxygen in Eq. (9) should be 0.5 and 0.25 for the
stoichiometry shown in Eqs. (1) and (3), respectively.

In order to compare the effect of different rotation


rates and cyanide concentrations on the anodic oxidation
of silver, Fig. 2a shows a loglog plot of RAg vs {1 / 2
CCN} for data sets GJ, based on Eq. (10).
logRAg logfx1=2 CCN g
logf0:62n1 y1=6 DCN 2=3 g

10

The excellent linear relationship with slope 1 for


rates based on anodic limiting currents shows the
validity of Eq. (10). The y-intercepts of the linear
relationships correspond to the values of log{0.62 n 1
1 / 6 (DCN)2 / 3} listed in Table 3. The substitution of
the value of = 0.89 10 6 m2 s 1 (Jeffrey and Ritchie,
2000) and n = 1, 2, 3 or 4 gives different values of DCN.
The most acceptable reaction mechanism should consider the formation of both Ag(CN)2 and Ag(CN)32
according to Fig. 1a, depending on the free cyanide
concentration. Kudryk and Kellogg (1954) reported an
estimated value of DCN = 1.75 10 5 (0.25) cm2 s 1,
based on the similarity in properties between KCN and
potassium halides. Thus, the most acceptable value of
DCN of the order 10 5 cm2 s 1, comparable with the
literature data, was obtained through the substitution of
n = 2 or 3, respectively (Table 3).
4.3. Oxidation of silver by oxygen
Fig. 2b shows a loglog plot of the rate of silver
dissolution by oxygen RAg against 1 / 2 to examine the
validity of Eqs. (8) and (9). Line E, with slope 1
represents the effect of increasing rotation rates at
40 mM cyanide and 21.3 kPa oxygen pressure (Table 2)
reported by Sun et al. (1996) and confirms the diffusion
controlled nature of the dissolution. Kudryk and Kellogg
(1954) reported a value of DO2 = 2.2 10 5 cm2 s 1
based on measured diffusivity of oxygen in water at
20 C, and corrected to 27 C. Table 4 lists the values of

G. Senanayake / Hydrometallurgy 81 (2006) 7585

log{RAg(anodic) / mol m-2 s-1}

Table 4
Diffusion coefficient of cyanide or oxygen based on Fig. 2b and Eqs.
(8)(10)

-3.0

y = 0.99x - 5.40
R2 = 0.99
-3.5 y = 1.04x - 5.43
R2 = 1.00
-4.0 y = 1.06x - 5.47
R2 = 1.00
-4.5
-5.0

G
H

1.5

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

1
2
3
4

3.09 10 8
1.09 10 8
5.96 10 9
7.83 10 6

2.21 10
7.82 10 6

2.5

coefficient consistent with the literature data (Kudryk


and Kellogg, 1954) was DO2 = 2.2 10 5 cm2 s 1
(Table 4).
Fig. 2c shows a loglog plot of RAg (set BF) against
{1 / 2CCN} to examine the validity of Eqs. (8) and (9) at
different rotation rates and cyanide concentrations.
While curves D, and B are based on corrosion currents
from mixed potential theory, curves C and F are based
on chemical dissolution (Table 2). The slopes of curves
B and C are close to 1 at lower values of {1 / 2CCN},
indicating the validity of Eq. (10). This is especially in
the case of line B where the intercept of 5.4 is close to
the intercepts shown in Fig. 2a.

E
-4.3

y = 0.97x - 5.27
R2 = 1.00

-4.4

-4.5
0.8

0.9

1.0

1.1

log{( 1/2)}

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

DCN (cm2 s 1)

log {1/2CCN}

0.5
0.25

G (Hiskey and Sanchez, 1993)


H (Hiskey and Sanchez, 1990;
effect of rotation rate)
I (Li and Wadsworth, 1993)
J (Sun et al., 1996)

0.5

DO2 (cm2 s 1)

y = 1.12x - 5.44
R2 = 1.00

-5.5

79

-4.0
D

-4.5
F

4.4. Limiting rates due to oxygen diffusion

-5.0

D (Sun et al., 1996; effect of cyanide)


B (Li et al., 1992)
F (Breuer et al., 2005)
C (Li et al., 1992)

C
-5.5
Slope = 1
-6.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

log {1/2CCN}

Fig. 2. Loglog plot of rate of silver dissolution (RAg) vs {1 / 2CCN} or


{1 / 2}: (a) anodic dissolution rates based on limiting currents, (b), (c)
corrosion or chemical dissolution rates based on icorr, chemical
analysis, or nano-balance (see Table 2 and text).

DCN or DO2 calculated from the intercept of line E which


should represent log{0.62 n 1 1 / 6 (DCN)2 / 3}for
cyanide diffusion or log{0.62 m 1 1 / 6 (DO2)2 / 3}for
oxygen diffusion. Clearly, the latter seems to be the
case because the only acceptable value for the diffusion
Table 3
Diffusion coefficient of cyanide ion based on Fig. 2a and Eq. (10)
Set

Intercept

105 DCN (cm2 s 1)


n=1

n=2

n=3

n=4

G
H
I
J

5.40
5.43
5.47
5.44

0.49
0.45
0.39
0.44

1.41
1.27
1.11
1.23

2.64
2.37
2.07
2.29

3.99
3.60
3.14
3.48

The data points in sets B, C and F in Fig. 2c based on


the results reported by Li et al. (1992), and Breuer et al.
(2005), obtained using different techniques and different
rotation speeds, merge at high values of {1 / 2CCN}.
This indicates that the rate of silver cyanidation is
controlled by the rate of oxygen diffusion to the surface.
The limiting rates RAg(lim) of silver cyanidation based on
data sets B, C, D, and F listed in Table 5 correspond to the
right-hand side of Eq. (9) representing oxygen diffusion.
The value of RAg(lim) is slightly higher in data set D than
that in the other three data sets. The corresponding values
of DO2 calculated assuming CO2 = 0.25 mol m 3 are also
listed in Table 5. The value of DO2 = 2.2 10 5 cm2 s 1
(Kudryk and Kellogg, 1954) closely agrees with the

Table 5
Diffusion coefficients of oxygen based on limiting rates in Fig. 2b and
Eq. (9)
Set

RPM

log{RAg(lim)}

105 DO2 cm2 s 1


m = 0.5

m = 0.25

B
C
D
F

100
500
450
300

4.67
4.61
4.48
4.65

3.02
1.11
1.88
1.42

1.07
0.39
0.67
0.50

G. Senanayake / Hydrometallurgy 81 (2006) 7585

4.5. Effect of oxygen pressure


Previous researchers have reported the effect of
oxygen partial pressure on the rate of dissolution of a
rotating silver disc (Deitz and Halpern, 1953), or a
rotating silver electrode (Hiskey and Sanchez, 1990;
Sun et al., 1996) in alkaline cyanide solutions. Fig. 3a
shows a loglog plot of RAg vs {1 / 2CCN} based on
the chemical dissolution of a rotating disc at three
different oxygen partial pressures 21.3 kPa (Li et al.,
1992) and 344 and 758 kPa (Deitz and Halpern, 1953)
at 2425 C. In all cases the rates reach limiting
values at higher values of {1 / 2CCN}, corresponding
to the limiting rates RAg = 2.6 10 5 mol m 2 s 1
(21.3 kPa), 1.8 10 4 mol m 2 s 1 (344 kPa), and
3.5 10 4 mol m 2 s 1 (758 kPa). The initial slope of
line C is close to unity at low cyanide concentrations.
According to these results the limiting rate is
approximately doubled with the doubling of PO2
from 344 to 758 kPa (curves A1 and A2). However,
the initial slope of lines A1 and A2 is close to 1.5
rather than 1 observed in curve C, Fig. 3a. This is
discussed in more detail in Section 7.

a
log(RAg / mol m-2 s-1)

values of DO2 in the range 1.11 10 51.88 cm2 s 1 for


sets C, D and F (m = 0.5) listed in Table 5. In the case of
set B, a low rotation rate of 100 rpm resulted in a higher
DO2 of 3.0 10 5 cm2 s 1 (m = 0.5); or a lower DO2 of
1.1 10 5 cm2 s 1 (m = 0.25).
While oxygen is essential for the chemical oxidation
of silver, the differences in limiting rates between line D
and lines B, C and F in Fig. 2b may be largely related to
differences in the nature/concentration of electrolyte
solutions. For example, the value of CO2 depends on
temperature, ionic strength, electrolyte composition, and
the partial pressure of oxygen, which in turn depends on
the elevation. La Brooy et al. (1991) discussed the
decrease in CO2 with the increase in salinity and its
effect on gold cyanidation. The effect of oxygen
pressure is discussed in more detail in Section 4.5.
Although the value of CO2 is generally taken as
0.25 mM in air saturated solutions at 25 C, Wadsworth et al. (2000) used a lower value of 0.19 mM
after correcting for elevation. Another factor, which can
contribute to the different limiting rates, is the presence
of other (impurity) metals in the solid state or metal ions
in solution. For example, Jeffrey and Ritchie (2001)
demonstrated that trace impurities in cyanide salts or
background salts such as sodium perchlorate can affect
the rate of gold cyanidation. The surface roughness
factor, which affects the actual surface area, may also be
responsible for different limiting rates.

-3.0

A1, 344 kPa oxygen pressure, 895 rpm


A2, 758 kPa oxygen presure, 895 rpm
C, 21.3 kPa oxygen pressure, 500 rpm

-3.5
-4.0

slope 1.5
slope 1.5

-4.5
-5.0 slope 1
-5.5
0

b
log(RAg / mol m-2 s-1)

80

0.5

1.5

2.5

log{1/2CCN}
-3.0

y = 1.01x - 6.68
R2 = 1.00

-4.0

-5.0
y = 0.75x - 6.71
R2 = 0.99
-6.0
1.5

2.0

2.5

y = 1.00x - 7.26
R2 = 0.99
A, 55 mM cyanide, 895 rpm
H1, 500 rpm
H2, 700 rpm
D1, 40 mM cyanide, 450 rpm
D2, 10 mM cyanide, 450 rpm
D3, 5 mM cyanide, 450 rpm
D4, 2.5 mM cyanide, 450 rpm
3.0

3.5

4.0

log{ 1/2PO2 / kPa s-0.5}

Fig. 3. Loglog plot of rate of silver dissolution (RAg) vs {1 / 2CCN}


(a), {1 / 2PO2}(b) based on corrosion or chemical dissolution (see
Table 2).

The effect of oxygen pressure is further exemplified


in Fig. 3b that shows a loglog plot of RAg vs {1 / 2
PO2}. Results reported by Deitz and Halpern (1953) for
tests carried out at 55 mM NaCN, PO2 = 136780 kPa,
and a rotation rate of 895 rpm show a good linear
relationship of slope 1 (line A). Results reported by Sun
et al. (1996), based on the point of intersection of the
anodic and cathodic polarisation curves of pure silver in
solutions of 40 mM cyanide and PO2 = 10.13101.3 kPa,
also show a good linear relationship (line D1). The
higher rates reported by Sun et al. (1996) for pure silver
may be a result of different pH, ionic strength, trace
impurities, surface roughness, or the effect of different
oxygen solubility.
The effect of increasing oxygen pressure on reaction
rate has to be quantified by considering the actual
concentration of dissolved oxygen. According to the
data reported by Livingston et al. (1930), the concentration of dissolved oxygen in water at different oxygen
partial pressures in the range 23101 kPa at 25 C
follows a good linear relationship given by CO2(mol/
m3) = 0.0123 PO2 (kPa) + 0.0042 with R2 N 0.99. This
relationship was used to convert PO2 in Fig. 3b to CO2.
For example, the calculated values of CO2 (mol/m3) at

G. Senanayake / Hydrometallurgy 81 (2006) 7585

Table 6
Pressure dependence of oxygen solubility in water at 25 C
PO2 (atm.)

PO2 (kPa)

CO2 (mol/m3)

CO2 (mol/t)

0.21
3.40
7.48

21.3
344
758

0.266
4.24
9.32

0.269
4.35
9.57

Based on data or equation reported by Livingston et al. (1930) and


Tromans (1998).

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

-3.5
y = 1.03x - 4.76
R2 = 1.00
-4.0

D1, 40 mM cyanide
D2, 10 mM cyanide
D3, 5 mM cyanide
D4, 2.5 mM cyanide

-4.5

-5.0
-0.2

log{RAg(lim) / mol m-2 s-1}

the three oxygen pressures related to Fig. 3a are listed in


Table 6. Tromans (1998) described a thermodynamic
analysis of oxygen solubility data in a range of oxygen
pressures and temperatures, which simplifies to the
concentration / pressure ratio given by CO2(mol/t) / PO2
(atm.) = 1.279 at 25 C. The calculated values of CO2
(mol/t) based on this relationship are also listed in Table
6, which show that the two sets of oxygen solubilities
are in close agreement.
Fig. 4a shows a loglog plot of RAg vs {1 / 2CO2}
for selected results from Fig. 3b. The intercept of the
two lines D1 and D2, representing the high concentrations of cyanide (40 and 10 mM, respectively), is
4.76 and this corresponds to the diffusion coefficient
DO2 = 1.61 10 5 cm2 s 1 for m = 0.5 based on Eq. (9).
The substitution m = 0.25 gives an unacceptable value
of DO2 = 0.6 10 5 cm2 s 1.
Moreover, as the cyanide concentration decreases
from 40 to 2.5 mM, the slope of some curves in Figs. 3b
and 4a decreases. The rate reaches a limiting value at the
concentration of 2.5 mM (line D4). It is of interest to
note that the cyanide diffusion coefficient calculated
using the limiting rate log{RAg} = 4.2 of line D4 and
n = 3 in Eq. (10) is DCN = 1.5 10 5 cm2 s 1. This value
is consistent with the literature value noted in Section
4.2. Thus, the four lines/curves D1D4 in Fig. 3b shows
how the rate determining step changes from oxygen
diffusion at high cyanide (D1, 40 mM cyanide) to
cyanide diffusion at low cyanide (D4, 2.5 mM cyanide).
This is similar to the situation in Fig. 2c based on
chemical dissolution of silver by oxygen as noted by Li
et al. (1992). In this case, the limiting rates at high values
of {1 / 2CCN} show that the rate is controlled by oxygen
diffusion.
Results reported by Hiskey and Sanchez (1990) at a
higher rotation speed of 700 rpm show a linear
relationship, but a lower slope of 0.75 (line H2 in Fig.
3b). The limiting rates of silver cyanidation by oxygen
at three different oxygen pressures at higher values of
1 / 2CCN (represented by plateaus of curves C, A1 and
A2 of Fig. 3a), can be used to examine the effect of
oxygen diffusion. Thus, Fig. 4b shows a loglog plot of
limiting rates RAg(lim) vs {1 / 2CO2}. Again, the slope of

81

0.0

0.2
0.4
0.6
log{ 1/2CO2 /mol m-3 s-0.5}

0.8

1.0

-3.0

-3.5

-4.0
y = 0.74x - 4.92
R2 = 1.00

-4.5

-5.0
0.0

0.4

0.8

1.2

1.6

2.0

2.4

log{ 1/2CO2}

Fig. 4. Loglog plot of rate of silver dissolution (RAg) vs {1 / 2CO2}


based on (a) corrosion or (b) chemical dissolution (limiting rates from
Fig. 3a, CO2 from Table 6 or from equations described in text).

the linear relationship is 0.74, showing that oxygen


diffusion alone is not responsible for these limiting rates.
This is unlike line E in Fig. 3b, or lines D1 and D2 in
Fig. 4a, in which the slope is unity.
The results based on chemical dissolution reported
by Li et al. (1992) agree with the diffusion model
expressed by the Levich equation (Eq. (9)). However,
the electrochemical data based on polarisation curves for
the oxidation of silver and reduction of oxygen obey the
Levich model only at higher cyanide concentrations and
low oxygen pressures (lines D1 vs D4 in Fig. 3b). The
slopes less than unity shown by some sets of data in
Figs. 3b, 4a and b suggest mixed chemically diffusion
controlled kinetics of silver cyanidation which leads to
the mixed charge-mass transfer model described in
Sections 5 and 6.
5. Current status of reaction mechanism
Based on rotating silver disc studies, Deitz and
Halpern (1953) concluded that:
(i) the cyanidation reaction takes place according to
Eq. (1), because the transition point at which the
rate determining factor changes from cyanide

82

G. Senanayake / Hydrometallurgy 81 (2006) 7585

transport to oxygen transport corresponds to a


concentration ratio of CNaCN / CO2 = 4.4, a value
close to 4.0 based on the stoichiometry in Eq. (1);
(ii) the transport of reactants to the silver surface
predominantly determines the rate, because the
concentration of hydrogen peroxide increased
with the amount of silver dissolved, while the
addition of small amounts of hydrogen peroxide
(up to 0.35 mM) during cyanidation with 57 mM
NaCN at an oxygen partial pressure of 486 kPa
made no significant difference to silver cyanidation kinetics;
(iii) the difference between the theoretical ratio of
CNaCN / CO2 = 4 (Eq. (1)) compared to the experimental value of 4.4 is due to incomplete
ionisation of cyanide and/or a slight difference
between the two diffusion coefficients of oxygen
and cyanide.
A series of reactions, as given by Eqs. (11)(15),
can be written for the anodic oxidation of silver,
while a similar set of reactions can be written for the
cathodic reduction of oxygen (Hiskey and Sanchez,
1990; Xue and Osseo-Asare, 2001). The subscripts
(b), (i) and (ads) denote bulk, interface, and adsorbed
species, respectively.
CNb CNi

11

Ags CNi AgCNads

12

AgCNads AgCN0ads e

13

AgCN0ads CNi AgCN2 i

14

AgCN2 i AgCN2 b

15

Ags 2CNb AgCN2 b e

16

Baltruschat and Vielstich (1983) also noted that


neither the anodic dissolution of silver in cyanide
solutions, nor its reverse process, are simple chargetransfer reactions. The rate determining step contains an
adsorption process, which causes the reaction orders to
be non-integers. Thus, they presented a slightly different
mechanism for silver cyanidation in aqueous KCNKCl
electrolytes that involved a partial charge transfer to
explain the non-integer reaction orders found in their
experiments:
AgCNads p1CN AgCNpq
qe
p

17

npCN
AgCNpq
p

AgCNn1
1qe
n

18

Nevertheless, the values of n = 2 or 3 in Eq. (18) for,


respectively, low or high cyanide concentrations, as
reported by Baltruschat and Vielstich (1983), are
consistent with the most acceptable values of n = 2 or
3 noted in Table 3. They also noted evidence based on
surface enhanced Raman scattering (SERS) of cyanide
ion on silver which indicated that the adsorbed silver
complex during anodic oxidation has a coordination
number 2.
Following the work of Kudryk and Kellogg (1954),
Hiskey and Sanchez (1990) and Li et al. (1992) used the
mixed potential theory to model the kinetics of silver
cyanidation by oxygen. They also showed the importance of incorporating the two effects due to diffusion/
adsorption and charge transfer in modeling the kinetics
of silver cyanidation.
More recent work by Xue and Osseo-Asare (2001)
showed that silver is more active than gold in a sodium
hydroxide solution in the absence of cyanide, which was
attributed to the formation of Ag(OH)0 on the surface.
However, the rate of anodic dissolution was low due to
the passivation or precipitation of Ag2O and AgO. Here,
the formation of the two oxides was supported by the
EHpH diagram. There was no thermodynamic evidence (EHpH) for the formation of silver oxides in the
presence of cyanide, except for Ag2O3 at EH above 1.5
V. Xue and Osseo-Asare (2001) used the mass transfer
law and ButlerVolmer equation to show that the
reaction order with respect to cyanide concentration
would depend on the rate-determining step. For
example, equilibration of the discharge step in Eq.
(19), followed by the forward reaction of the adsorbed
species with cyanide (Eq. (14)), will give rise to an
overall reaction order of 2. In contrast, a slow rate
controlling discharge step shown by the forward
reaction of Eq. (19) would give a reaction order of 1.
An observed reaction order of 1 with respect to cyanide
(Xue and Osseo-Asare, 2001; Sun et al., 1996) supports
the latter.
Ags CNb AgCN0ads e

19

6. Rate constant based on mixed charge-mass


transfer model
Li and Wadsworth (1993) treated the anodic dissolution of silver according to the reaction mechanism

G. Senanayake / Hydrometallurgy 81 (2006) 7585

83

given in Eq. (19) followed by Eq. (14) leading to the


formation of Ag(CN)2 . Neglecting the back reaction of
Eq. (19), this gave the overall anodic reaction given by
Eq. (16). They derived a rate equation, which contained
terms for both diffusion of cyanide and intrinsic kinetics
of the electrochemical reaction:

diffusion effect. This value is of the same order as the


limiting rates of silver cyanidation (2.1 10 5
3.3 10 5 mol m 2 s 1) at higher values of {1 / 2
CCN} (Fig. 2b, Table 5).

ia 1 f1:61sm1=6 =zFDCN 2=3 CCN gx1=2


fi0 expaa zFg=RT g1

Li and Wadsworth (1993) also reported the exchange


current density of silver cyanidation as a function of
cyanide concentration at 24 C for the concentration
range of 0.510 mM. The relevant rates (RAg)0 = i0 / zF
(z = 1) are of the order 2.9 10 612 10 6 mol m 2 s 1
(Table 7)) and generally an order of magnitude lower
than the rate based on pure kinetic current reported by Li
and Wadsworth (1993). A plot of ln{i0} vs ln{CCN}
corresponds to a slope 0.51. This is in agreement with the
slope 0.49 reported by previous researchers (Vetter,
1967) for the variation in exchange current density of the
Ag/Ag(CN)32 electrode in solutions of cyanide concentration N 200 mM. Vetter (1967) presented the
following equation derived in Appendix 1. It shows
the dependence of the exchange current density on
cyanide ion concentration, which in turn is useful to
determine the reaction order of the anodic electrochemical reaction ():

20

where ia = anodic current density, i0 = exchange current


density, s = stoichiometric number = 2 for Eq. (16), z = 1
for Eq. (16), a = transfer coefficient for anodic reaction,
= overpotential, R = gas constant, and T = absolute
temperature, and the other terms have usual meanings
described in previous sections. A series of experiments
were carried out in 2.5 mM cyanide solutions at a fixed
over-potential of 0.125 V and 24 C at different rotation
rates. A plot of (ia) 1 against 1 / 2 gave a linear
relationship with the values for the slope and intercept
given by Eqs. (21) and (22) thus confirming the
validity of Eq. (20).
Slope : 1:61sm1=6 =zFDCN 2=3 CCN
0:146 cm2 s1=2 lA1

21

dfln io g=dflnCCN g ka n

Intercept (at 1 / 2 0):


fi0 expaa zFg=RT g1 2:5  103 cm2 lA1

7. Reaction order for cyanidation of silver

22

These two equations are important for the comparison of results from the two methods of dissolution:
chemical and electrochemical, as the slope represented
(ilim) 1 and intercept gave information on the intrinsic
rate constant of the surface reaction. Li and Wadsworth
(1993) reported the value of a = 0.48 at 24 C. The
substitution of relevant values for and F in Eq. (21)
gives DCN = 1.50 10 5 cm 2 s 1 . This value is
consistent with the value of DCN = 1.29 10 5 cm2
s 1 (Sun et al., 1996) based on limiting currents and
the Levich equation. Hiskey and Sanchez (1990) used
a relationship similar to Eq. (20) based on a mixed
kinetic model involving coupled diffusion/adsorption
and charge transfer for the anodic oxidation of silver
and cathodic reduction of oxygen. They calculated the
values of DCN = 1.77 10 5 cm2 s 1 and DO2 = 1.60
10 5 cm2 s 1.
The reciprocal of the intercept, shown in Eq. (22),
corresponds to a current density of 400 A cm 2, which
in turn translates to a rate of 4 10 5 mol m 2 s 1 based
on rate = i / zF for z = 1. Li and Wadsworth (1993)
referred to this value as pure kinetic current without

23

The substitution of slope = 0.51 based on the results


listed in Table 7, and = 0.48 (Li and Wadsworth, 1993),
gives the values of = 2 (for n = 3), = 1.5 (for n = 2), or
= 1 (for n = 1). While the substitution of n = 2 or 3 gives
acceptable values for DCN in Table 3, the values of = 1
and 1.5 are consistent with the initial slopes shown in
Fig. 3a for silver leaching. These results highlight the
need to consider the silver(I) species formed at the
surface, as well as in solution, in developing surface
chemical models.
8. Activation energy and rate constant
Deitz and Halpern (1953) reported an activation
energy Ea = 10 kJ mol 1 for the chemical dissolution of
Table 7
The values of exchange current densities and relevant rates at 24 C
CKCN (mM)

io (A cm 2)

106 (RAg)0 (mol m 2 s 1)

0.5
1.0
2.5
5.0
10

25.9
35.9
58.4
85.1
118

2.86
3.72
6.05
8.82
12.2

Li and Wadsworth (1993).

84

G. Senanayake / Hydrometallurgy 81 (2006) 7585

silver in oxygenated cyanide solutions at 37 mM NaCN


and an oxygen partial pressure of 344 kPa in the
temperature range 24110 C, and confirmed the
diffusion controlled nature of the reaction. Li and
Wadsworth (1993) reported a value of Ea = 13.7 kJ
mol 1 based on the effect of temperature (060 C) on
the diffusion coefficient of cyanide ion. However, the
value of Ea based on the effect of temperature on the
exchange current density was slightly higher (21.1 kJ
mol 1, Li and Wadsworth, 1993). This supports the
mixed charge-mass transfer model described in Section
6, and indicates the possible analysis of rate data in set C
for chemical dissolution of silver reported by Li et al.
(1992) on the basis of mixed surface adsorption
reactiondiffusion model.
The product of oxygen reduction in Eq. (24) is
written as H2O2 in accordance with the findings of Deitz
and Halpern (1953). It is also supported by the fact that
the substitution of m = 0.5 in Eq. (9) (rather than
m = 0.25) gives the most acceptable values for DO2 in
Tables 4 and 5.

for the cyanidation of a silver electrode, which is based


on the effect of cyanide concentration on the exchange
current density reported by Li and Wadsworth (1993).
Results based on chemical dissolution of silver by
oxygen at 21 kPa also obey the Levich equation with a
stoichiometric ratio of Ag : O2 = 1 : 0.5.
The limiting rate of silver cyanidation at an oxygen
partial pressure of 21 kPa and 25 C at high cyanide
concentration is 2.5 10 5 mol m 2 s 1 . This
average value is of the same order as the rate
constant 4 10 5 mol m 2 s 1 based on pure kinetic
current independent of the diffusion effect in the
mixed charge-mass transfer model for anodic
oxidation of silver reported by Li and Wadsworth
(1993). Moreover, at high oxygen pressures of 344
and 758 kPa representing high dissolved oxygen the
reaction order with respect to cyanide vary from 1.5
to 0 showing deviations from Levich model at both
low and high cyanide concentrations, and supporting
the mixed chemically diffusion controlled model.

Ags H2 O nCN 0:5O2


AgCNn1
0:5H2 O2 OH
n

Appendix A
24

According to the surface adsorptionreaction model,


the limiting rate at maximum surface coverage 1
corresponds to the rate constant of the surface reaction
(RAg = kAg). Thus, it is comforting to see that the
limiting rate of silver cyanidation by oxygen based on
data set C 2.6 10 5 mol m 2 s 1 (Fig. 3a, line C) is in
good agreement with kAg = 4 10 5 mol m 2 s 1 based
on Eq. (22). A more detailed analysis of the rate data of
chemical dissolution based on mixed surface adsorptionreactiondiffusion model will be presented in a
future communication.
9. Summary and conclusions
The literature data on the rate of chemical or
electrochemical dissolution of silver in alkaline
cyanide solutions of a range of concentrations and
oxygen partial pressures have been reviewed to
compare and contrast the kinetic models for
dissolution.
Results based on corrosion currents show excellent
agreement with the Levich equation, especially at
high cyanide concentration and low oxygen partial
pressures, although the involvement of 2 or 3 cyanide
ions in the reaction has to be taken into account in
order to calculate the diffusion coefficient DCN. This
is consistent with the anodic reaction order of 1.52

Exchange current density io for the anodic reaction in


Eq. (26) is given by Eq. (27), where Eeq = equilibrium
potential and ka = electrochemical rate constant for the
anodic reaction and = reaction order. Differentiating
Eq. (28) with respect to ln [CN] gives Eq. (29).
k1

Ags kCN AgCNk


adsorption equilibrium
k1
AgCNk

25

0:5O2 H2 O nkCN Y Products


26

io ka CN k expazFEeq =RT

27

ln io ln ka kln CN  azFEeq =RT

28

dfln io g=dflnCN g
k azF=RT dfEeq g=dflnCN g

29

If the stable Ag(I) species in solution is Ag(CN)n(n1),


the redox reaction at equilibrium and the expression for
Eeq are given by Eqs. (30) and (31). Differentiating
Eq. (31) with respect to ln[CN] at a given concentration of Ag(CN)n(n1) gives Eq. (32).
AgCNn1
e Ags nCN
n

30


Eeq E o Ag =Ag RT =zFlnAgCNn1
n
n
RT =zFlnCN 
31
dEeq =dlnCN  nRT =zF

32

G. Senanayake / Hydrometallurgy 81 (2006) 7585

Substitution of Eq. (32) in Eq. (29) gives the


following relationship shown in Eq. (23) (Vetter,
1967), (d{ln i0} / d{ln[CN]}) = n.
References
Arikado, T., Iwakura, C., Yoneyama, H., Tamura, H., 1976. Anodic
oxidation of potassium cyanide on the graphite electrode.
Electrochim. Acta 21, 10211027.
Baltruschat, H., Vielstich, W., 1983. On the mechanism of silver
dissolution and deposition in aqueous KCNKCl electrolytes.
J. Electroanal. Chem. 154, 141153.
Bodlnder, G., 1896. Die chemie des cyanidverfahrens. Z. Angew.
Chem. 9, 583587.
Breuer, P.L., Dai, X., Jeffrey, 2005. Leaching of gold and copper
minerals in cyanide deficient copper solutions. Hydrometallurgy
78, 156165.
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, 755764.
Deitz, G.A., Halpern, J., 1953. Reaction of silver with aqueous
solutions of cyanide and oxygen. J. Met. 11091116.
Elsner, L., 1846. ber das verhalten verschiedener metalle in einer
wssrigen lsung con zyankalium. J. Prakt. Chem. 37, 441446.
Hiskey, J.B., Sanchez, V.M., 1990. Mechanistic and kinetic aspects of
silver dissolution in cyanide solutions. J. Appl. Electrochem. 20,
479487.
Hogfeldt, E., 1982. Stability Constants of Metal-Ion Complexes, 2nd
Supplement, IUPAC Chemical data series No. 2, Part A, Inorganic
ligands. Pergamon, Oxford.
Jeffrey, M.I., Ritchie, I.M., 2000. The leaching of gold in cyanide
solutions in the presence of impurities II. The effect of silver.
J. Electrochem. Soc. 147, 32723276.
Jeffrey, M.I., Ritchie, I.M., 2001. The leaching and electrochemistry of
gold in high purity cyanide solutions. J. Electrochem. Soc. 148,
D29D36.
Kudryk, V., Kellogg, H.H., 1954. Mechanism and rate-controlling
factors in the dissolution of gold in cyanide solutions. J. Met.
541548.

85

Levenspiel, O., 1972. Chemical Reactions Engineering. Wiley, New


York.
La Brooy, S.R., Muir, D.M., Komosa, T., 1991. Oxygen requirements
and monitoring for gold ore processing. World Gold'91. Aus.I.M.
M., Melbourne, pp. 165171.
Li, J., Wadsworth, M.E., 1993. Electrochemical study of silver
dissolution in cyanide solutions. J. Electrochem. Soc. 140,
19211927.
Li, J., Zhong, T., Wadsworth, M.E., 1992. Application of mixed
potential theory in hydrometallurgy. Hydrometallurgy 29, 4760.
Livingston, J., Morgan, R., Richardson, A.H., 1930. Solubility of
oxygen in water. J. Phys. Chem. 34, 23562366.
Sillen, L.G., Martell, E., 1964. Stability Constants of Metal-Ion
Complexes, Special Publication Nos. 17 and 26. Chemical Society,
London.
Sun, X., Guan, Y.C., Han, K.N., 1996. Electrochemical behaviour of
the dissolution of goldsilver alloys in cyanide solutions. Metall.
Mater. Trans. 3B, 355361.
Tromans, D., 1998. Temperature and pressure dependent solubility of
oxygen in water: a thermodynamic analysis. Hydrometallurgy 48,
327342.
Vetter, K.J., 1967. Electrochemical Kinetics. Academic Press, London.
Wadsworth, M.E., Zhu, X., 2003. Kinetics of enhanced gold
dissolution: activation by dissolved silver. Int. J. Miner. Process.
72, 301310.
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.
Xue, T., Osseo-Asare, K., 1985. Heterogeneous Equilibria in the Au
CNH2O and AgCNH2O systems. Metall. Trans. 16B,
455463.
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.), In Cyanide: Social,
Industrial and Economic Aspects. TMS, Warrendale, pp. 563576.

Vous aimerez peut-être aussi