You are on page 1of 17

Minerals Engineering 17 (2004) 785801

This article is also available online at:


www.elsevier.com/locate/mineng

Gold leaching in non-cyanide lixiviant systems: critical issues


on fundamentals and applications
G. Senanayake

Department of Extractive Metallurgy and Mineral Science, A.J. Parker Cooperative Research Centre for Hydrometallurgy,
Murdoch University, Perth, WA 6150, Australia
Received 24 November 2003; accepted 20 January 2004

Abstract
Thermodynamics and speciation of the Au(0/I/III) system are reviewed on the basis of published data on disproportionation of
Au(I) and solubility of gold metal and Au(I/III) salts. Kinetics of gold oxidation from rotating gold discs, gold colloid, goldsilver
alloy and gold ores reported in the literature are reviewed to compare and contrast dierent lixiviant systems. Linear correlations of
stability constants of Au(I) complexes with respect to relevant complexes of Ag(I) and Cu(I) show the stability order of Au(I)



2


complexes with dierent ligands: CN > HS > S2 O2
3 > SC(NH2 )2 > OH > I > SCN > SO3 > NH3 > Br > Cl > CH3 CN. In the






case of Au(III) the stability order of complexes is CN > OH > SCN > Br > Cl . The EhpH and Ehlog[Cl ] diagrams of Au(0/I/
III)OH NH3 and Fe(II/III)Cu(0/I/II)Au(0/I/III)Cl systems are revised to incorporate complex species such as AuOH0 ,

Au(OH)
2 and AuCl2 . Solubility of gold salts follows the order NaAuCl4 > KAuCl4 > AuCl3  Au(NH3 )4 (NO3 )2 > Au2 S/
H2 S > Au(OH)3 /NaOH > Au(OH)3 . Solubility of gold metal in the presence of oxidants decreases in the order MgS/H2 S  Fe(III)/
Cl > Cu(II)/Cl > Cu(II)/NH3  NaOH/H2 O at a given temperature, but increases with increasing temperature. Results from
rotating disc studies in dierent lixiviant systems show that the rate of gold oxidation depends on the temperature and concentration of
oxidant and ligand. Silver dissolves faster than gold and thus Ag(I) catalyses the gold dissolution by redox-displacement. The surface
chemical reaction mechanism for gold dissolution is rationalized on the basis of electrochemical and adsorption theory. Rate of gold
leaching in dierent lixiviant systems is represented by a shrinking sphere/core model with an apparent rate constant of 105 s1 .
 2004 Elsevier Ltd. All rights reserved.
Keywords: Hydrometallurgy; Gold ores; Leaching; Reaction kinetics; Redox reactions

1. Introduction
Gold can be dissolved using a range of oxidants in the
presence of a range of ligands (L) to produce soluble
species Au(I)L2 and Au(III)L4 (charge ignored, Eqs. (1)
(6)). The cyanidation process (Eq. (1)) has been used to
leach gold over 100 years since it was patented in 1888
by MacArthur and Forrest brothers (MacArthur et al.,
1888). Despite the long-term use of cyanidation there is
a growing interest in non-cyanide gold technology based
on the lixiviants noted in Eqs. (2)(6) and Table 1,
mainly due to the failure of cyanidation to extract gold
from the so-called dicult-to-treat ores and the environmental and safety issues (Nicol et al., 1987; Hiskey
and Atluri, 1988; La Brooy et al., 1994; Sparrow and
Woodcock, 1995; Ritchie et al., 2001).
*

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


E-mail address: g.senanayake@murdoch.edu.au (G. Senanayake).

0892-6875/$ - see front matter  2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2004.01.008

4Au O2 2H2 O 8NaCN


4NaAuCN2 4NaOH
4Au O2 2H2 O 8L 4AuIL2 4OH

1


2Au H2 O2 4L

2AuIL2 2OH L Cl ; S2 O2


3 ; SCNH2 2
3

2Au L2 2L or L
3 L

2AuIL2 L Cl; Br; I; SCN; SCNH2 2


Au 1:5L2 L AuIIIL4 L Cl; Br; I

4
5

Au CuII or FeIII 2L
AuIL2 CuI or FeII


L Cl ; S2 O2
3 ; SCN ; SCNH2 2 ; NH2 CH2 COO ;

NH2 CHCH3 COO

786

G. Senanayake / Minerals Engineering 17 (2004) 785801

Table 1
Oxidants and ligands for non-cyanide gold processing (selected from
Sparrow and Woodcock, 1995)
Oxidants

Ligands

NH
4
H2 S
H2 O or OH
H2 SO3
O2 or H2 O2
S or S2
x

X2 , X
3 or OX
Fe(III), Cu(II)

NH3
HS , S2
OH
SO2
3

Cl , OH , S2 O2
3 , HSO3 , NH3
HS , S2
X (X Cl. Br, I, SCN)

Cl , S2 O2
3 , SCN , SC(NH2 )2 ,
NH2 (CH2 )COO , NH2 CH(CH3 )COO

For example, coppergold ores consume cyanide whilst


carbonaceous (preg-robbing) material in gold ores
adsorb gold cyanide and prevent the solubilization of
gold. The silica encapsulation and the association of
gold with sulde minerals make the refractory gold ores
resistant to gold cyanidation, even after ne grinding. In
such cases it is essential to pre-treat the ore by roasting,
pressure oxidation, and bio- or chemical oxidation to
degrade the carbonaceous material or oxidize the sulde
material. However, the neutralization of acidic residues
of pre-oxidation treatment, prior to cyanidation, is
uneconomical due to lime requirement and material
handling problems (Wan et al., 2003). Gold recovery in
cyanidation of washed acid-autoclave residue produced
at chloride levels of 100 g/t Cl or higher is lower,
and cannot be improved by extending the autoclave
residence times (Ketcham et al., 1993). In addition to
these technical problems, several catastrophic accidents
caused by the failure of gold tailings dams have alarmed
the public and increased the concern about environmental hazards of cyanide (Miller and Pritsos, 2001).
Despite long term interest in non-cyanide gold lixiviant systems, the validity of some published data for
equilibrium constants are questionable and have lead to
erroneous EhpH diagrams. For example, recent studies
(Senanayake et al., 2003) on the copperammonia
thiosulphate system for gold leaching showed that the
measured equilibrium constant for the ligand displace
2
ment reaction: Au(S2 O3 )3
2 +2NH3 Au(NH3 )2 +2S2 O3
9
11
is in the range 10 10
compared to the predicted
values in a much larger range of 100:4 1017 , based on
the published thermodynamic data. Thus the use of
erroneous thermodynamic data for the construction of
EhpH diagrams shows that Au(NH3 )
2 is a predominant complex at pH > 10 (Zipperian et al., 1988). In
contrast, no attempt has been made to incorporate
AuOH0 and Au(OH)
2 in the EhpH diagrams, despite
the availability of reliable thermodynamic data for these
species. Gold solubility in weathering uids containing
oxygenated-thiosulphate may be enhanced by the presence of dissolved silver, generated by oxidation of coexisting silver mineralization; thus the gold and silver
solubilities are interdependent due to the formation of a

mixed-metal complex of the type AuAg(S2 O3 )2


2 (Webster, 1986). The lack of information on this type of
mixed-metalligand complexes hamper further investigations in this area. Likewise, there is little agreement
amongst researchers regarding the nature of the actual
species present in a given Au(I) sulphide solution:
2
Au(HS)(H2 S)0 , Au(HS)
and AuHSOH
2 , Au2 S(HS)2
0
2
(Belevantsev et al., 1981); AuHS , Au(HS)
2 , AuS2

(Renders and Seward (1989); AuS (Jiayong et al.,
1996). These facts highlight the importance of a reliable
thermodynamic database for Au(0/I/III) systems that
can be used to rationalize the complex chemistry in noncyanide leaching and recovery processes for gold.
Few attempts have been made to compare and contrast the rate data and reaction mechanisms for dierent
lixiviant systems (Pesic and Sergent, 1993; Dasgupta
et al., 1997; Ritchie et al., 2001; Jerey et al., 2001), and
the comparison of actual leaching data has been limited
(Chen et al., 1980; Navarro et al., 2002, Kholmogorov
et al., 2002). Since all the non-cyanide gold processes are
still at the developmental stages, a better understanding
of the thermodynamics and kinetics of the Au(0/I/III)
systems will (i) rationalize the complex chemistry, (ii)
assist the selection and optimization of a suitable noncyanide leaching system with respect to dierent types of
material, and (iii) lead to investigation of novel gold
lixiviants. This paper intends to compare and contrast
the non-cyanide gold lixiviants of current interest on the
basis of (i) thermodynamics including stability constants
of Au(I/III) complexes, EhpH and Ehlog[Cl ] diagrams, disproportionation of Au(I) to Au(III) and Au0 ,
solubility of gold(I/III) salts and gold metal, and (ii) the
published rate data for the dissolution of gold from
rotating gold discs and from gold ores.

2. Non-cyanide lixiviants of current interest


A recent review by Ritchie et al. (2001) highlighted
the importance of a comparative study, using cyanide as
the bench mark, of the characteristics of alternative
leaching systems. Chloride, thiosulphate, sulphide,
ammonia, and sulte have been identied as the low cost
systems; the rst two are more benecial in terms of
health and safety issues and low environmental impact.
For example, Pangum and Browner (1996) and Ferron
et al. (2003) demonstrated that the chloride catalysed
acid-pressure leaching of refractory or sulphidic gold
ores and leach residues produced by preg-robbing concentrates has the potential for direct dissolution of gold
within the autoclave and thus avoiding the cyanidation
treatment. Hunter et al. (1996) summarized the advantages of a sulphide leaching system in a bio-process
which makes use of the natural sulphur cycle to mitigate
the production of acidic solutions and tailings, and to
enhance the recovery of gold from sulphide-hosted gold

G. Senanayake / Minerals Engineering 17 (2004) 785801

ores. The controlled pressure oxidation of gold bearing


sulphides with HNO3 /O2 at relatively low temperature
and low oxygen pressure to produce sulphur allows the
alkaline sulphide lixiviation of gold based on Eqs. (7)
and (8) (Anderson, 2003).
4S0 6NaOH 2Na2 S Na2 S2 O3 3H2 O

2Aus S2
x 2AuS x  2S

Wan et al. (2003) highlighted the continuing research


interest on the use of thiocyanate for gold recovery in an
acidic environment from a bio-oxidized low grade sulphidic ore. The use of the thiourea system with air, O2 ,
H2 O2 or Fe(III) (Groenewald, 1976; Desch^enes and
Ghali, 1988), the mixed thioureathiosulphate system
with ferricyanide (Hiskey, 1984), the mixed ammonia
thiosulphatesulphide system (Eqs. (7) and (8), Jiayong
et al., 1996) for gold leaching have been reported. Han
(2001) reviewed the use of ammonia with a range of
oxidants at temperatures of up to 200 C to leach gold
from refractory ores. Aylmore and Muir (2001), Molleman and Dreisinger (2002) and Grosse et al. (2003)
highlighted the adaptability of the copperammonia
thiosulphate system to a range of technologies including
heap, dump and in situ leaching of gold.
Unlike Cl , OH and NH3 , the sulphur containing
gold ligands are susceptible to degradation in an oxidizing environment (Eqs. (9)(15)), which leads to (i)
high reagent consumption, (ii) lower rates of gold dissolution in leaching or diagnostic tests due to low concentration of reagents and the formation of insoluble
products such as elemental sulphur, Cu2 S, CuS, Au2 S
and other gold salts (Zipperian et al., 1988, Lorenzen and
Van Deventer, 1993; Jiayong et al., 1996; Abbruzzese
et al., 1995; Muir and Aylmore, 2002), and (iii) interference of degradation products in subsequent unit operations for separation and recovery of gold (Nicol and
OMalley, 2001; West-Sells et al., 2003). For the sake of
simplicity the Cu(I/II) complexes with ammonia are not
shown in Eq. (10), but considered later in the discussion
related to the reaction mechanism for gold dissolution.
2Na2 S2 O3 O2 2H2 O Na2 S4 O6 2NaOH

Na2 S2 O3 3 or 4 CuSO4
0:5Na2 S4 O6 Na3 or 5 CuS2 O3 2 or 3 Na2 SO4
10
Na2 S4 O6 2NaOH Na2 S3 O6 Na2 S5 O6

11

Na2 S5 O6 3NaOH 2:5Na2 S2 O3 1:5H2 O

12

2NaSCN Fe2 SO4 3 SCN2 2FeSO4 Na2 SO4


13
3SCN2 4H2 O H2 SO4 HCN 5HSCN

14

2CSNH2 2 Fe2 SO4 3


CSNHNH2 2 2FeSO4 H2 SO4

15

787

Whilst the oxidized forms of ligands such as


L2 (SCN)2 and (CS(NH2 )2 )2 in Eq. (4) are capable of
oxidizing gold (Groenewald, 1976; Li and Miller, 2002;
Wan et al., 2003), the ligand is partially reproduced in
some degradation reactions (Eqs. (12) and (14)). Oxidation of ligands such as SC(NH2 )2 (Chai et al., 1999)
and precipitation of metal sulphides can be avoided by
adding Na2 SO3 (Zipperian et al., 1988), whilst the
addition of amino acids (Michel and Frenay, 1999) and
ethylenediaminetetraacetic acid (EDTA) lowers the
thiosulfate degradation and improves the gold extraction with Cu(II)NH3 S2 O2
3 (Xia et al., 2003). In order
to avoid the thiosulphate degradation and toxic
ammonia, Ji et al. (2003) used O2 /Na2 S2 O3 /pH 1112 at
elevated temperatures of 6080 C and O2 pressures of
10100 psig to leach gold from carbonaceous material.

3. Thermodynamics
3.1. Linear correlations of stability constants
The values of the equilibrium constant bn [AuLn ]/
[Auz ][L]n for the reaction Auz +nL AuLn listed in
Table 2 provide useful information on gold(I/III) speciation in process liquors; z 1 and 3 for Au(I) and
Au(III) corresponds to the coordination numbers n 1,
2 and 3, 4 respectively. The published values of bn are
generally based on the measured reduction potentials
and/or saturated solubility of chlorides, cyanides, sulp2
hides and hydroxides. Unlike Ag(H2 O)
h and Cu(H2 O)h
which are stable in pure water as hydrated ions of the
relevant salts (e.g. AgNO3 , CuSO4 ), the mono-valent

ions Cu(H2 O)
h and Au(H2 O)h are unstable in the absence of strongly complexing ligands due to disproportionation e.g. 3Au(I) 2Au(s)+Au(III). A plot of
logfb2 CuIL2 or AuIL2 g against logfb2 AgIL2 g
shown in Fig. 1, suggested by Finkelstein and Hancock
(1974), can be used for two purposes: (i) to examine the
validity of the published data for Au(I) and Cu(I) systems, and (ii) to predict bn values for systems for which
reliable data are not available.
Linear functions (denoted by LF1 and LF2) used for
this purpose, based on Fig. 1, are shown in Table 3.

Surprisingly, the log b2 values of Au(NH3 )


2 and Cu(NH3 )2
show large deviations from LF1 and LF2 respectively.
The log b2 value of Au(SO3 )3
2 also shows a signicant
deviation from LF1, whilst that of Cu(I)L2 complexes of
CH3 CN, SCN and CN show signicant deviations
from LF2. Large deviations of Au(NH3 )
2 complexes
from LF1 have been related to (i) the error in the value of
E0 (Au /Au(0)) used in calculation by some authors, (ii)
diculty in the preparation of aqueous Au(NH3 )
2 due to
disproportionation, and (iii) possibility of interaction of
other anions such as CN with Au(NH3 )
2 (Senanayake
et al., 2003). The substitution of published values of

788

G. Senanayake / Minerals Engineering 17 (2004) 785801

Table 2
Stability constants of gold(I) and gold(III) complexes (log bn for n 1  4) at 25 C
Ligand

Complex

log bn

Chloride

AuCl
2
AuCl
4
AuBr
2
AuBr
4
AuI
2

AuI4
AuOH0

9.71
25.3
12.7
32.8
19.2
47.7
10.2
20.6
24
44
45.6
18, 19.5, 21, 26a
13b
46, 30b
17.2
43.9
38.3
85
1.6
23.3
25.4
18
17.7
24.5
30.1
32.8
38.9
41.1
72.9
26, 24b
28
26.8, 16b

Bromide
Iodide
Hydroxide

Au(OH)
2
Au(OH)03
Au(OH)
4
Au(NH3 )
2
Au(NH3 )
2
Au(NH3 )3
4
Au(SCN)
2
Au(SCN)
4
Au(CN)
2
Au(CN)
4
Au(CH3 CN)
2
Au(SC(NH2 )2 )
Au(SeC(NH2 )2 )
Au(NH2 CH2 COO)
2
Au(NH2 CH(CH3 )COO)
2
AuHS0
Au(HS)
2

Ammonia

Thiocynate
Cyanide
Acetonitrile
Thiourea
Selenourea
Glycinate
Alanate
Bisulphide

Thiosulphate

AuS
Au2 S2
2
Au2 S(HS)2
2
Au(S2 O3 )3
2

Sulphite

Au(SO3 )3
2

Sulphide

Reference

Kissner et al. (1997)


Stefansson and Seward (2003)
Stefansson and Seward (2003)

Skibsted and Bjerrum (1974b, 1977)


Ritchie et al. (2001); Aylmore and Muir (2001)
Skibsted and Bjerrum (1974b)

Sharpe (1976)
Johnson et al. (1978)

Michel and Frenay (1999)


Michel and Frenay (1999)
Renders and Seward (1989)
Renders and Seward (1989)
Belevantsev et al. (1981)
Webster (1986)
Renders and Seward (1989)
Belevantsev et al. (1981)
Pouradier and Gadet (1969)
Skibsted and Bjerrum (1977)
Webster (1986)

CN-

HS-

y = 1.85x
R2 = 0.99

SCN

30

S2O32SeC(NH2)2

I-, SC(NH2)2

NH3
SO32-

40

Cl

CH3CN

log{2(AuL2 or CuL2)}

50

Based on the data from Finkelstein and Hancock (1974) and Bard (1973) unless stated otherwise.
a
A range of values are reported based on dierent values of E0 (Au /Au).
b
Value based on linear free energy correlations in Figs. 1 and 2 (Senanayake et al., 2003).

20

y = 0.96x
R2 = 0.99
Au(I), LF1
Au(I) with large deviations
Cu(I), LF2
Cu(I) with large deviations

H2O

10
0
0

10

15

20

25

log{2(AgL2)}

Fig. 1. Loglog plot of b2 of Au(I) and Cu(I) complexes vs Ag(I)


complexes at 25 C. Data: gold (Table 2), silver and copper (Hancock
et al., 1974; Hogfeldt, 1982); Cu(I)HS (Young et al., 2003); Ag(I)
HS (Schwarzenbach and Widmer, 1966).

logfb2 AgIL2 g in LF1 and LF2 gives rise to

3
logfb2 AuNH3 2 g 13 and logfb2 AuSO3 2 g 16
listed in Table 2.

Fig. 2 shows the variation of logfb4 AuIIIL4 g vs


logfb2 AuIL2 g for the ligands Cl , Br , SCN , OH
and CN and the approximately linear correlation is denoted by LF3 in Table 3. Likewise, the linear correlation
between logfb2 CuIL2 g vs logfb2 AuIL2 g for the
ligands HS , Cl , NH2 CH2 COO and NH2 CH(CH3 )COO in Fig. 2 is denoted by LF4 in Table 3. The
substitution of logfb2 AuCH3 CN
2 g 1:6 and logfb2

AuNH3 2 g 13 in LF3 predicts logfb4 Au-CH3 3


3
CN4 g 4 and logfb4 AuNH3 4 g 30 respec
tively. Thus both Au(OH)
2 and Au(OH)4 have higher
stability constants than their counterparts with L NH3 ,
based on LF1 and LF3 (Table 2). Fig. 3 shows a good
linear correlation of the published values of log bn for the
three sulphide complexes MHS0 , M(HS)
2 and M2 S(HS)2
2 of Cu(I) and Au(I) vs Ag(I), showing the validity
of these constants. In summary, the linear correlation of
stability constants of complex species proposed by Finkelstein and Hancock (1974) paves the way to establish
reliable thermodynamic data for Au(I/III) complexes in
non-cyanide lixiviant systems.

G. Senanayake / Minerals Engineering 17 (2004) 785801

789

Table 3
Linear mathematical relationships between stability constants of Cu(I), Ag(I) and Au(I/III) complexes based on Figs. 1 and 2
No.

Figure

Linear relationship

LF1
LF2
LF3
LF4

Fig.
Fig.
Fig.
Fig.

logfb2 (AuL2 )}
logfb2 (CuL2 )}
logfb4 (AuL4 )}
logfb2 (CuL2 )}

1
1
2
2

Cl-

40
20

CN-

y = 2.28x
R2 = 0.98

Br-

SCN-

60

Cl-

NH2CH2COO

NH2CH(CH3)COO-

0
0

10

15

20

25

30

2xCH3 CN predominant reaction

y = 0.53x
R2 = 1.00
35

40

M2S(HS)2

2-

y = 0.80x + 20.09
R2 = 1.00

y = 0.79x + 15.16
2
R = 1.00
MHS
M(HS)2

log {n(Au(I) or Cu(I) complexes)}

Cu(I)
Au(I)

20
0
0

10

20

30

40

0.99)
0.99)
0.98)
1.00)

3AuCH3 CN
x

16

3H2 O

2Aus AuOH03

100

40

Fig. 2. Loglog plot of b4 of Au(III) and b2 of Cu(I) vs b2 of Au(I)


complexes at 25 C. Data from Refs. shown in Fig. 1 caption.

60

(R2
(R2
(R2
(R2

2AuCH3 CNx H2 O
2Aus 0:5O2 g 2H

log{2(AuL2)}

80

1.85 logfb2 (AgL2 )}


0.96 logfb2 (AgL2 )}
2.28 logfb2 (AuL2 )}
0.53 logfb2 (AuL2 )}

solutions of very low concentration: 0.6% (pH 2), 1.7%


(pH 7) and 1.9% (pH 12). The evolution of oxygen gas
was also observed at pH > 10. The two disproportionation reactions which explain these observations are
represented by Eqs. (16) and (17).

HS-

80

OH-

Au(III), LF3
Cu(I), LF4

H2 O

log{4(AuL4)} or log{2(CuL2)}

100

50

60

70

80

log { n( Ag(I) complexes)}


2
Fig. 3. Loglog plot of b of MHS0 , M(HS)
2 and M2 S(HS)2 for Au(I)
vs Ag(I) complexes at 25 C. Data from Table 2 and Schwarzenbach
and Widmer, 1966.

3.2. Disproportionation of Au(I) and EhpH diagrams


Table 4 lists the disproportionation reactions denoted
by DR1-DR11 for Cu(I) and Au(I) to highlight the large
equilibrium constants (Kd ) in the range 106:5 1013 com5
pared to the small values for CuCl
2 (10 , DR1) and

11
Au(OH)2 (10 , DR4). Kissner et al. (1997) mixed a

solution of Au(CH3 CN)
2 ClO4 prepared by anodicoxidation of gold in acetonitrile with water at 25 C and
examined the disproportionation at pH 2, 7 and 12 at 25
C. Surprisingly, instead of disproportionation, 1 mM
Au(I) was slowly reduced, yielding 99.4%, 98.4% and
98.1% Au(0) respectively with the formation of Au(III)

3H 3xCH3 CN minor reaction

17

Despite the large value of Kd 108 (DR9) for AuCl


2 in
Table 4, Diaz et al. (1993) reported that the rate of

disproprtionation of AuCl
2 to Au(0) and AuCl4 is also
very slow, taking more than 2 months to reach equilibrium. However, Pesic and Sergent (1993) showed that
the dissolution of gold from a rotating disc in Br2 /NaBr
produces AuBr
2 , which is subsequently disproportionated fairly rapidly to Au0 and AuBr
4 . Skibsted and
Bjerrum (1974b) prepared a solution of Au(NH3 )
2
in situ by reducing Au(NH3 )3
with hydrazine in a
4
background medium of 10 M NH4 NO3 and observed
13
the disproportionation of Au(NH3 )
2 (Kd 10 , Table
4). They have also determined equilibrium constants for
the hydrolysis of Au(NH3 )3
to Au(NH3 )3 OH2 and
4
acid dissociation of Au(NH3 )3
to Au(NH3 )3 NH2
4
2
(Skibsted and Bjerrum, 1974a). Table 4 lists the reported
values of the relevant equilibrium constants for hydrolysis (Kh ) and acid-dissociation (Ka ). Due to the lower
stability of Au(NH3 )
2 (Table 2) it may also hydrolyse to
AuOH0 and Au(OH)
2 and subsequently disproportionate to Au(s) and O2 , according to reactions HR3,
HR5, DR5 and DR7 listed in Table 4.
0:6
A comparison between Au(NH3 )
or 103
2 (Kh 10

15
for HR3 and HR5) and AuCl2 (Kh 10
or 103 for
HR1 and HR2) in Table 4 clearly shows that AuCl
2 is
more stable towards hydrolysis. Moreover, as shown by
DR4, DR5 and DR7, the calculated values of Kd for the
disproportionation of AuOH0 and Au(OH)
2 to Au(s)
and O2 (100:2 101 ) is much larger than Kd for the disproportionation that produces Au(s) and Au(OH)
4
(1011 ). Figs. 4 and 5 make use of the information
in Table 4 to incorporate the Au(I/III) species such
2
as AuOH0 , Au(OH)
and Au(NH3 )3 2 , Au(NH3 )3 OH
NH2
in
the
EhpH
diagrams
for
Au(0/I/III)O2 /
2

790

G. Senanayake / Minerals Engineering 17 (2004) 785801

Table 4
Disproportionation (Kd ), hydrolysis (Kh ) and acid dissociation (Ka ) equilibrium constants of Cu(I) and Au(I) in aqueous media at 25 C; calculated
using bn from Table 2 and Hogfeldt (1982) and E0 from Bard (1973) unless stated otherwise
No.

Type of equilibrium

Reference

Disproportionation reaction (Kd )


DR1
DR2
DR3
DR4
DR5
DR6
DR7
DR8
DR9
DR10
DR11


2CuCl
2 (aq) Cu + CuCl (aq)+ 3Cl
2Cu (aq) Cu0 + Cu2 (aq)
2
0
2Cu(NH3 )
2 Cu + Cu(NH3 )4


3Au(OH)

2Au(0)
+
Au(OH)
2
4 + 2OH
0

Au(OH)

Au
+
OH
+
0.25O
+
0.5H
2
2O
2


3Au(SCN)
2 2Au + Au(SCN)2 + 2SCN
AuOH0 Au + 0.25O2 + 0.5H2 O
0


3AuBr
2 2Au + AuBr4 + 2Br
0


3AuCl
2 2Au + AuCl4 + 2Cl
3Au 2Au0 + Au3
3
0
3Au(NH3 )
2 2Au + Au(NH3 )4 +2NH3

105
106
106:5
1011
100:2
100:3
101:2
105
108
109
1012

Skibsted and Bjerrum (1974a)

Hydrolysis reaction (Kh )


HR1
HR2
HR3
HR4
HR5


AuCl
2 + 2H2 O Au(OH)2 +2H + 2Cl
0


AuCl
2 + H2 O AuOH + H + 2Cl


Au(NH3 )
2 + 2H2 O Au(OH)2 + 2NH4
2
Au(NH3 )3
+ NH
4 + H2 O Au(NH3 )3 OH
4

0
Au(NH3 )
2 + H2 O AuOH + NH3 + NH4

1015
102:5
100:6
100:2
102:8

Acid dissociation reaction (Ka )


AD1
AD2

Au(NH3 )3
Au(NH3 )3 NH2
4
2 +H

NH

NH
+
H
3
4

107:5
109:2

Au

Au(OH)4

1.5

Au(OH)3(s)
or Au2O3(s)

1.0

3.3. Solubility of gold(I/III) salts

Au(OH)2

0.0

Au(s)

(b)

-0.5
-2

10

12

14

pH

Fig. 4. EhpH diagram for Au(0/I/III)H2 O system at 25 C (lines a


and b represent O2 /H2 O and H /H2 ). [Au (I or III)] 105 M.

2.0
3+

Au

Au(NH3)42+

(a)

Au(NH3)3NH22+

Au(OH)4-

1.0
0
-

AuOH

Au(OH)2

Eh / V

AuOH

0.5
Au(s)

0.0

NH4+
NH3

(b)

Au(s)

-0.5
-2

Skibsted and Bjerrum (1974a)


Hogfeldt (1982)

0.5

1.5

Skibsted and Bjerrum (1974a)

AuOH

Eh / V

(a)

Skibsted and Bjerrum (1974a)

NH3 H2 O systems. Unlike the previously reported Eh


pH diagram (Zipperian et al., 1988), Fig. 5 does not
show Au(NH3 )
2 , as a result of its tendency to hydrolyse
to AuOH0 and/or disproportionate to Au and O2 .

3+

2.0

Skibsted and Bjerrum (1974a)


Skibsted and Bjerrum (1974a)

6
pH

10

12

14

Fig. 5. EhpH diagram for Au(0/I/III)NH3 system at 25 C. [Au (I or


III)] 105 M.

The saturated solubility of salts in hydrometallurgical


systems provide useful information on chemical speciation and plays an important role in selective leaching,
separation and recovery of metals or salts by precipitation. For example, two recent reviews (Muir, 2002; Senanayake and Muir, 2003) have highlighted the eect of
background chloride concentration, temperature and
pH on the solubility of base metal chlorides, depending
on the stability of chloro-complexes, which leads to
important applications in chloride hydrometallurgical
processes. In general the solubility of complex salts such
as KAu(CN)2 , Na3 Au(S2 O3 )2 , NaAuCl4 and KAuCl4
are relatively high compared to Au2 S and Au(OH)3
(Table 5). The descending order of log b4 of Au(III)
complexes with the three ligands: OH > NH3 > Cl
shown in Table 2 and Fig. 2 is not observed with
the solubility data shown in Table 5: NaAuCl4 >
KAuCl4 >AuCl3  Au(NH3 )4 (NO3 )2 > Au2 S/H2 S > Au
(OH)3 /NaOH > Au(OH)3 . It is important to note that
despite the high value of logb3 44 for Au(OH)03 (Table
2) the solubility of Au(OH)3 in water is as low as 6 mg/l
(Table 5) due to the precipitation of Au(OH)3 or Au2 O3
(Fig. 4). Moreover, the 1.9% Au(III) produced during
the disproportionation of 1 mM Au(I) at pH 12 (Kissner

G. Senanayake / Minerals Engineering 17 (2004) 785801

791

Table 5
Solubility of gold salts
Salt

T (C)

Medium

AuCN
Au2 S2 O3 .3BaS2 O3
Au2 S

25

0.4 m KCN
Water
H2 S(aq)

NaAuCl4
KAuCl4
AuCl3
Au(NH3 )4 (NO3 )3
Au(OH)3
Au(OH)3
Au(OH)3

30
30

[S2 ] total (mol/kg)

Au solubilitya

pH

(mg/kg)

25

(g/kg)
70
6.6

0.008
0.009
0.01
0.2
0.02
0.4

2.0
7.3
11.6
11.6
12.0
11.9

0.003
4.1
2.7
39
14
256

Water
Water
Water
Water
0.41 M NaOH
0.1 M NaOH
Water

25
25
25
25

967
495
441
13
0.21
0.02
0.006

Data: Linkey and Seidell (1958), Au2 S data from Renders and Seward (1989).
a
Gold (I/III) dissolved in 1 kg of water.

et al., 1997) according to Eq. (17) corresponds to 4 mg/l


Au(III); this value is in close agreement with the solubility of Au(OH)3 in water (6 mg/l, Table 5).
Solubility of Au2 S depends on the two parameters pH
and total sulphide content [S2 ]t . For example, at a xed
value of S2
t 0:01 M, the concentration of Au(I)
increases thousand-fold from 3 lg/kg at pH 2 to 4 mg/kg
at pH 7 and drops to 3 mg/kg at pH 11.6. However, a
forty-fold increase in [S2 ]t from 0.01 to 0.4 M at
pH 12 causes an eighty-fold increase in the solubility
of Au(I) from 3 to 256 mg/kg. Fig. 6 shows loglog plots
of solubility of Au2 S and Au(OH)3 against pH and
[OH ]/Kw (Kw ionic product of water) respectively to
compare and contrast the behaviour of the two systems.
An approximate slope of 12 for the initial increase in
log[Au(III)] against logfOH
=Kw g shows the formation of complex species according to Eqs. (18) and (19).

In the case of Au2 S, the solubility dependence on both


pH and sulphide concentration has been related to the
formation of the three complex species shown in Eqs.
(20)(22) (Renders and Seward, 1989). Based on the
gold(I) complex Au2 S(HS)2
2 shown in Fig. 3, Eq. (23)
can also be included in the Au2 S solubilization equilibria. It is clear from these equations that a decrease in
concentration of H2 S in solution caused by a decrease in
partial pressure of H2 S favours the precipitation of
Au2 S.
AuOH3 s NaOH NaAuOH4 aq

18

AuOH3 s 2NaOH Na2 HAuO3 2H2 O

19

0:5Au2 Ss 0:5H2 Saq AuHS


0:5Au2 Ss 1:5H2 Saq

20


AuHS2

Au2 Ss Na2 S Na2 Au2 S2


Au2 Ss 2H2 Saq

1.E-01

2
Au2 SHS2

21
22

2H

23

Au(III) / NaOH 0.8 - 8 mol / kg

[Au(I) or Au(III)] / mol/L

1.E-02

Au(I) / hydrogen sulphide 0.004 - 0.01 M

3.4. Solubility of gold metal

Au(I) / hydrogen sulphide 0.02 - 0.1 M

1.E-03
1.E-04
1.E-05
1.E-06
1.E-07
1.E-08
0

12

16

pH or log{[OH-]/Kw}

Fig. 6. Eect of pH or [NaOH] or total sulphide [S2 ]t on solubility of


Au2 S in H2 S or Au(OH)3 in NaOH (25 C). Data: Au2 S (squares,
Renders and Seward, 1989); Au(OH)3 /NaOH (circles, Linkey and
Seidell, 1958).

Fig. 7 and Table 6 show the solubility of gold metal in


dierent oxidizing media O2 /CN , Cu(II)/Cl , Fe(III)/
Cl , NaOH and MgS (McDonald et al., 1987; Liu and
Nicol, 2002; Stefansson and Seward, 2003; Fleet and
Knipe, 2000), to highlight the eect of temperature,
nature of the dissolved oxidant and ligand in Eqs. (24)
(30). The descending order of Au(I/III) solubility in
various redox systems in Fig. 7: S2 (0.030.15 M) >
Fe(III)/Cl (0.010.41 mM) Cu(II)/Cl (0.060.6 mM) >
Cu(II)/NH3 (0.080.41 mM) > OH /(0.022.5 lM) largely reects the composite eect of decreasing stability
of the relevant gold(I/III) complexes and the solubility
of salts listed in Tables 2 and 5.

792

G. Senanayake / Minerals Engineering 17 (2004) 785801

[Au(I) or Au(III)] / mol/L or mol/kg

1.0E+01
NaAuCl4
Au(NH3)4(NO3)2

1.0E+00

Au(s)/MgS

1.0E-01
1.0E-02
1.0E-03

Au(OH)3 /0.4 M NaOH


Au2S /0.2 M Na2S/pH 11.6

1.0E-04
1.0E-05
1.0E-06
1.0E-07

Au(OH)3
Au(s)/Cu(II)/NH3
Au(s)/Cu(II)/HCl/NaCl
Au(s)/Fe(III)/NaCl/H2SO4
Au(s)/NaOH

Au(s)/Au(III)/ClO4CH3CN/H2O/pH 2

1.0E-08
10

100
Temperature / C

1000

Fig. 7. Eect of temperature on gold solubility in dierent lixiviant


systems and comparison with gold salt solubility. Data: NaAuCl4
(Linkey and Seidell, 1958); other gold salts (Table 4); Au/MgS (Fleet
and Knipe, 2000); Au/Fe(III)/H2 SO4 /NaCl (Liu and Nicol, 2002); Au/
Cu(II)/HCl/NaCl (McDonald et al., 1987) Au/Cu(II)/NH3 (Meng and
Han, 1996) Au/NaOH (Stefansson and Seward, 2003).

Aus FeCl3 NaCl NaAuCl2 FeCl2

24

Aus CuCl2 2NaCl NaAuCl2 NaCuCl2

25

NaAuCl2 2FeCl3 or CuCl2


NaAuCl4 2FeCl2 or CuCl

26

Aus H2 O AuOH 0:5H2 g

27

Aus NaOH H2 O NaAuOH2 0:5H2 g


28
0

Aus H2 S AuHS 0:5H2 g

29

Aus 0:5MgS 1:5H2 S


Mg0:5 AuHS2 0:5H2 g

30

The solubility of gold in 0.1 M FeCl3 /2.4 M NaCl at 40


C ( 0.1 mM, Table 6) compares well with the value of

0.05 mM in 0.1 M FeCl3 /1 M NaCl/1.1 M H2 SO4 (Fig.


7). The two values from dierent sources represent the
saturated solubility associated with the redox-equilibrium between gold metal and Fe(III) in Eqs. (24) and
(26). McDonald et al. (1987) and Liu and Nicol (2002)
highlighted the importance of considering the thermodynamic data at high temperatures to predict the relevant redox potentials and equilibrium data for gold
oxidation according to Eqs. (24) and (26). Table 7 lists
the equilibrium constant for selected reactions which
involve dissolution of gold by a range of oxidants to
highlight the eect of temperature. The increase in
temperature from 25 to 100150 C increases the equlibrium constant by 45 orders of magnitude in the case
of sulphide and chloride systems. The calculated equilibrium concentration of the gold dissolved in 0.6 M
CuCl2 /3.42 M NaCl/0.1 M HCl at 102 C shows that the

value of [AuCl
2 ] is 10 times larger than that of [AuCl4 ],
whilst the total concentration of these two species: 108
mg/l Au(I) and 12 mg/l Au(III) is in excellent agreement
with the measured value (115 mg/l, McDonald et al.,
1987).
For the purpose of comparison the solubility of
Au(III) compounds such as Au(OH)3 , Au(NH3 )4 (NO3 )3 , Au2 S and NaAuCl4 (s) in water, NaOH(aq) or
Na2 S(aq) solutions (Table 5) are also shown in Fig. 7.
Although the solubility of NaAuCl4 in water at 25 C is
several orders of magnitude larger than that of Au(OH)3
( 105 mg/kg), Au(NH3 )4 (NO3 )3 ( 102 mg/kg) and
Au2 S ( 104 mg/kg) it remains fairly constant at 3.85.9
mol/kg in the temperature range 1060 C. In contrast,
the solubility of gold metal in dierent lixiviant systems
shows large temperature dependence in chloride,
ammonia and hydroxide lixiviants. Thus Fig. 7 summarizes the eect of temperature on gold solubility due to
the change in redox equilibrium constants and chemical

Table 6
Solubility of gold dissolved from metallic gold, goldsilver alloy (al), colloidal gold (col) or mixed goldsilver colloid
System
Au/O2 /KCN
Au/Fe(III)/NaCl
Au/Cu(II)/NaCl
Au/H2 O2 /SC(NH2 )2

Au/O2 /Na2 S2 O3
Au(al) (64 mol% Ag)
Au(col)
Au(col) (12% Ag)
Au(col) (20% Ag)

[Oxidant] (mM)
7
100
100
600
0
0
17.6
8.8
8.8
8.8
100b

[Ligand] (mM)
0.15
2.4
3.4
3.4
8.6
8.6
8.6
8.6
8.6
8.6
pH 8.6

Data from Linkey and Seidell (1958), if not stated otherwise.


a
Ambient.
b
Eh 0.3 V.

Acid (M)

0.1 (HCl)
0.1 (HCl)
None
0.1 (H2 SO4 )
0.1
0.05
0.1
0.2

T (C)

Time (h)

Au solubility (mg/l)

18
40
102
104

1
185
23
1
6
6
6
14
14
14
1512

4100
23
45
170 (McDonald et al., 1987)
0
3
9.5
41
43
41
1.5 (Webster, 1986)
24
0.7
10
11

25

G. Senanayake / Minerals Engineering 17 (2004) 785801

793

Table 7
Eect of temperature on gold dissolution equilibria in selected non-cyanide media
Dissolution equilibrium

K (25 C)

K (150 C)

Reference

Au(s) + H2 S + HS Au(HS)


2 + 0.5H2 (g)
Au(s) + H2 S AuHS0 + 0.5H2 (g)

Au(s) + CuCl + 3Cl AuCl
2 + CuCl2

Au(s) + CuCl02 + 2Cl AuCl
+
CuCl
2
2
Au(s) + H + 2Cl AuCl
2 + 0.5H2 (g)
Au(s) + H2 O AuOH0 + 0.5H2 (g)

106
1011:5
1012
1012
1019
1022

101:6
107:1
108 (at 102 C)
108 (at 102 C)
1012
1013

Benning and Seward (1996)


Benning and Seward (1996)
McDonald et al. (1987)
McDonald et al. (1987)
Shenberger and Barnes (1989)
Stefansson and Seward (2003)

1.4

speciation for the relevant reactions (Eqs. (24)(30)).


Some of the important features in Fig. 7 are noted below:

AuOH0

Cl2(g)
Cl-

AuCl4-

1.2
Au(s)

3.5. Ehlog[Cl ] diagram for FeCuAuClO2 system


The Ehlog[Cl ] relationships provide a convenient
way of summarizing the chloride complex species, such
as those shown in Tables 4 and 7, involved in redox
reactions in chloride media (Eqs. (24)(26)). Fig. 8 plots
the Ehlog[Cl ] relationships for AuCuFeCl system
at 25 C in acidic media of pH 1 to incorporate AuOH0
and AuCl
2 and thus to update the recently reported
diagram (Senanayake and Muir, 2003). The Eh of redox
couples at a given value of [Cl ] 1 M follows the order
Cl2 /Cl > O2 /H2 O > Au(III)/Au(I) > Au(I)/Au(0) > Fe(III)/Fe(II) > Cu(II)/Cu(I) at 25 C.
McDonald et al. (1987) reported that (i) when copper
containing metal alloys are leached with CuCl2 solutions
in the absence of O2 , the solubility of gold is considerably
reduced due to equilibration with Cu(I), (ii) when copper
metal is added to an equilibrium solution of gold dissolved in CuCl2 , gold precipitates in the ask instead of

Fe3+

FeCl2+

FeCl2+

0.2

Cu(s)

FeCl+

2-

Cu

CuCl3

2+

0.4

CuCl2

Fe2+

CuCl

0.6

AuCl2-

Au(s)

0.8

CuCl2-

1.0

Eh / V

Solubility of Au(III) produced by disproportionation


of Au(CH3 CN)
2 in water at pH 2 (Eq. (17)) is as low
as 6 lmol/l at 25 C. In the presence of the oxidant
Fe(III) in mixed acidic chloride/sulfate media, the solubility of Au(I and/or III) in equilibrium with Au(s) is
also of the same order of magnitude (9 lmol/l) at 20 C.
However, the gold solubility increases to 0.41 mM
and 0.27 mM respectively in Fe(III) and Cu(II) chloride media, with an increase in temperature to 90 C.
At 114 C the gold solubility in Cu(II)/Cl increases
to 0.7 mM. In comparison, the solubility of gold in
Cu(II)/NH3 is much lower, but increases from 0.08
mM at 140 C to 0.4 mM at 180 C.
The highest solubility of gold metal is observed in
MgS/H2 O/0.15 GPa pressure and increases from 62
to 150 mM with the increase in temperature from
116 to 200 C, and decreases to 32 mM at 400 C.
In contrast, the solubility of gold in 0.5 mol/kg
NaOH is much lower even at high temperatures:
0.02 lM at 300 C and 2.7 lM at 500 C. This conrms the lower stability of complex species such as
AuOH0 and Au(OH)
2 and their tendency to disproportionate to Au(s) and O2 (Table 4 and Fig. 4).

0.0
-0.2
-2.0

(a)

AuCl4-

FeCl3

FeCl2

(b)

-1.0

0.0

1.0

2.0

log[Cl ]
Fig. 8. Ehlog[Cl ] diagram for Au(0/I/III)Fe(II/III)Cu(0/I/II)O2
Cl system at 25 C based on bn (Hogfeldt, 1982) and E0 (Bard, 1973),
assuming unit activity coecients, [Au(I)] [Au(III)] 105 M and
others at 0.1 M, pH 1.

cementing onto copper metal, (iii) the addition of metallic


copper lowers the Cu(II)/Cu(I) potential by lowering
[Cu(II)] according to the reaction Cu(s) + Cu(II)
2Cu(I) and thus facilitates the reverse reaction of leaching
(Eq. (25)), (iv) the solubility of gold in CuCl2 increases in
the presence of oxygen, but decreases with the addition of
CuCl under nitrogen. All of these observations and
interpretations are consistent with the Ehlog[Cl ] diagram in Fig. 8. McDonald et al. (1987) and Liu and Nicol
(2002) conrmed that the dissolution of gold with Fe(III)
and Cu(II) in chloride systems at elevated temperatures
takes place according to Eqs. (24)(26). The role of O2 is
largely conned to the regeneration of Cu(II) or Fe(III)
(Eq. (31)) so that decrease in Cu(I) or Fe(II) activity
facilitates gold dissolution (Eqs. (24)(26)).
2CuI or FeII 0:5O2 2H
2CuII or FeIII H2 O

31

4. Kinetics and reaction mechanism


4.1. Dissolution of metallic or colloidal gold
Table 6 compares the time taken for [Au(I)] or
[Au(III)] to reach dierent values of equilibrium concentrations depending on the lixiviant system: cyanide,

794

G. Senanayake / Minerals Engineering 17 (2004) 785801

chloride, thiourea and thiosulphate. The general trend in


rate, based on the ratio of gold dissolved/time, follows
the order cyanide > chloride > thiourea > thiosulphate.
Fig. 9 shows the eect of adding dierent acids and
bases on the concentration of gold dissolved by H2 O2 /
thiourea in 6 h. The following points are worth noting:

[Au dissolved]/ mg/L

Thiourea alone in the absence of oxidant or acid does


not dissolve gold (Table 6).
The mixture 8.6 mM SC(NH2 )2 /0.1 M H2 SO4 dissolves 3 mg/l in 6 h, even in the absence of an oxidant
such as H2 O2 .
At the same ligand/acid concentration a further increase in dissolved gold concentration to 9.5 mg/l is
observed in the presence of 17.6 mM H2 O2 in 6 h.
The addition of bases is detrimental to gold dissolution, whilst the addition of acids is benecial
(HNO3 < HCl < H2 SO4 ) (Fig. 9).
The dissolved gold concentration can reach higher
values of 4043 mg/l over a large period of time of
14 h, even with a lower concentration of H2 O2 of
8.8 mM, irrespective of the H2 SO4 concentration in
the range 0.050.2 M (Table 6).
In comparison to the thiourea system, the rate of gold
dissolution in O2 /Na2 S2 O3 seems to be very slow, tak-

10
8
6
4
2
0
0.25 M
NaOH

0.01 M
Ca(OH)2

none

0.16 M
HNO3

0.3 M
HCl

0.1 M
H2SO4

Fig. 9. Eect of acids and bases on gold dissolution in 17.6 mM H2 O2 /


8.6 mM SC(NH2 )2 in 6 h. Data from Linkey and Seidell (1958).

ing 6 weeks to dissolve 1.5 mg/l. However, Table 6


shows that the presence of silver increases the solubility from 1.5 to 24 mg/l (gold metal) and from 0.7 to
11 mg/l (gold colloid).
Table 8 compiles the published rate data per unit
surface area (mol/m2 /s) for the dissolution of gold, silver
and goldsilver alloy in dierent lixiviant systems, based
on rotating gold discs. The rate of dissolution of pure

Table 8
Rate of dissolution gold, silver or goldsilver alloy in dierent lixiviant systems
Oxidant/ligand//pH/other conditions

[Oxidant] (mM)

[Ligand] (mM)

T (C)

105 rate (mol/m2 /s)

Reference

Cl2 /1 M HCl
Cl2 /Cl /pH 4
Br2 /Br /pH 4.5
Fe(III)/SC(NH2 )2 /0.1 M H2 SO4
O2 /CN /pH 10.5
Cu(NH3 )2
4 /pH 9.7/0.5 M (NH4 )2 SO4
Co(NH3 )3
6
O2
H2 O2
O2 /CN /100 g/l NaCl
O2 /CN
O2 /CN
Cu(II)/S2 O2
3 /0.4 M NH3
HOCl/NaCl/pH 3.5
c
Cu(II)/S2 O2
3 /0.84 M NH3 /pH 10
NaOCl/NaCl/pH 6
NaOCl/NaCl/pH 6
NaOCl/NaCl/pH 7
Cu(II)/NH3 /pH 10/1.0 M (NH4 )2 SO4
Cu(NH3 )2
4 /0.32 M NaOCl
I2 /I /pH 10/1 M NH3
Fe(III)/NaCl/0.5 M H2 SO4
Fe(III)/NaCl/0.5 M H2 SO4
O2 /S2 O2
3

2
2
25
5

1000
1000
97
26.3
77

10
200
400
200

4.6
2.3
24
2.4
1.2
0.06
0.05
0.04
0.01
)(2.5)a
0.1 (4.1)a
)(9.6)a
3.8 (3.8)
1.4 (1.6)a
)(3.9a , 26.6b )
0.7
14
7
0.06
1.1
0.07
11.7
19.3
0.01

Diaz et al. (1993)

1
1
1
1
0.15(air)
0.26(air)
1.28(O2 )
10
0.8
25
27
135
135
1
157
1
100
100
sat

12
12
25
25
25
75
75
75
75
20
20
20
20
20
20
20
20
20
135
140
75
200
200
30

O2 /S2 O2
3 /0.2 M (NH4 )2 S2 O3
O2 /SC(NH2 )2 /0.2 M (NH4 )2 S2 O3
O2 /SC(NH2 )2 /0.2 M(NH4 )2 S2 O3
O2 /SC(NH2 )2 /0.2 M Na2 S2 O3

sat
sat
sat
sat

200
5
10
5

30
30
30
30

Gold dissolved from goldsilver alloy.


Pure silver.
c
Also contained 0.44 (NH4 )2 SO4 .
b

5
5
5
100
1700
400
85
850
850
2000

0.40
0.55
0.47
0.04

Pesic and Sergent (1993)

Dasgupta et al. (1997)

Jerey et al. (2001)

Jerey (2001)
Tran et al. (2001)

Han (2001)

Liu and Nicol (2002)


Chandra and Jerey
(2003)

G. Senanayake / Minerals Engineering 17 (2004) 785801


0.4

300
Ag metal

Ag-Au alloy

Ag / Au alloy (64% Ag)

[Au(I)] / mmol L-1

Ag colloid
-1

Ag / Au mixed colloid (20 mol% Ag)

[Ag(I)] / mg L

795

200

Ag / Au mixed colloid (11mol% Ag)

100

Ag-Au mixed colloid

0.3

Ag-Au ore
Slope = 0.6

0.2
Slope 0.5
0.1

0.0
0

0
10

20

30

40

50

Time / weeks

Fig. 10. Concentration of silver dissolved from pure silver, silvergold


alloy or mixed silvergold colloid in oxygenated thiosulphate media
(see Table 6, Data: Webster, 1986).

2
silver in Cu(II)NH3 S2 O2
3 (26.6 mol/m /s) is six times
faster than that of gold from AuAg alloy. Fig. 10
shows the concentration of silver dissolved from silver
metal, silver colloid, mixed silvergold colloid and silvergold alloy in the presence of O2 /Na2 S2 O3 /pH 10/Eh
0.3 V. It is of interest to note that the rate of dissolution
of silver from both mixed silvergold colloid and alloy is
slower than that of pure silver. Moreover, the dissolved
silver seems to approach an equilibrium concentration
in the case of mixed silvergold colloid. Whilst pure
silver dissolves faster than pure gold (Table 7), gold
dissolves faster in the presence of silver (Table 6). Thus
silver(I) in solution seems to catalyse the gold dissolution according to Eqs. (32)(34) (c colloid).

Aus or c AgI AuI Ags or c


Aus or c

S2 O2
3

3
AuS2 O3 2

Aus or c

32

AgS2 O
3

Ags or c

0.1

0.2

33

0.3

0.4

0.5

0.6

-1

[Ag(I)] / mmol L

60

Fig. 11. Concentration of dissolved gold(I) against dissolved silver(I).


Data: AuAg alloy and mixed goldsilver colloid with O2 /Na2 S2 O3
(Webster, 1986); AuAgCu ore with Fe(III)/SC(NH2 )2 (Chen et al.,
1980).

NaOCl/NaCl (20 C) > Fe(III)/Cl (200 C) > Cl2 /HCl/
12 C > Cu(II)/NH3 /S2 O2
3 (20 C) > Fe(III)/SC(NH2 )2
(25 C) > Cu(NH3 )2
/NaOCl
(140 C) > O2 /SC(NH2 )2
4
(30 C) > O2 /(NH4 )2 S2 O3 (30 C). However, due to the
fact that rate (R) is expressed by the general equation
R k [Oxidant]a [Ligand]b , where a and b are the
reaction orders with respect to oxidant and ligand
respectively, the rate constant (k) gives a better comparison between dierent lixiviant systems. Thus, Fig. 12
compares the loglog plots of rate vs [Oxidant] in the
three dierent lixiviant systems O2 /CN , Cu(II)/NH3 /
S2 O2
and Fe(III)/SC(NH2 )2 . Table 9 summarizes the
3
kinetic parameters based on Fig. 12. The slopes of linear
relationships in Fig. 12 indicates the reaction order 0.6
with respect to [O2 ] and 0.9 with respect to [Fe(III)] and
[Cu(II)]. In the case of Cu(II) the order decreases to 0 at
higher values of [Cu(II)]. Despite the lower concentration of 5 mM CN compared to 400 mM S2 O2
3 and 140
mM SC(NH2 )2 , the value of k [Ligand]b i.e. the y

AgS2 O3 3
2

AuS2 O3 3
2 Ags or c

34

A plot of the concentration of Au(I) vs Ag(I) at dierent


time intervals during the dissolution of alloy or mixed
silvergold colloid gives a slope 0.6 (Fig. 11). This
value is in reasonable agreement with the equilibrium
constant for Eqs. (33) and (34) calculated using the
published values of DG0f of relevant species reported by
Zipperian et al. (1988) and Aylmore and Muir (2001). It
is of interest to note that the gold and silver leached
from a coppersilvergold ore using Fe(III)/SC(NH2 )2 /
H2 SO4 (Chen et al., 1980), also follow the same linear
trend (Fig. 11).

3
2
log {105 R / mol m-2 s-1}

y = 0.6x + 2.7
R2 = 1.0

0
-1

Slope = 0.9
y = 0.9x + 2.8
R2 = 1.0

Oxidant = Oxygen in cyanide


Oxidant = Cu(II) in ammonia-thiosulphate

-2
-3
-4.0

Oxidant = Fe(III) in thiourea

-3.5

-3.0

-2.5

-2.0

-1.5

-1.0

log{[Oxidant] / mol L-1}

4.2. Comparison between dierent lixiviants


Table 8 shows the descending order of rate of gold
dissolution per unit surface area: Br2 /Br (25 C) >

Fig. 12. Loglog plot of rate of gold dissolution from rotating discs vs
concentration of oxidant. Data: O2 /cyanide (Table 7); Cu(II)NH3
S2 O2
3 /30 C (Jerey, 2001); Fe(III)/SC(NH2 )2 /25 C (Chen et al.,
1980).

796

G. Senanayake / Minerals Engineering 17 (2004) 785801

Table 9
Kinetic parameters in the rate equation rate k [Oxidant]a [Ligand]b , from Fig. 12
Lixiviant


O2 /CN
Cu(II)/S2 O2
3 /NH3
Fe(III)/SC(NH2 )2 /H2 SO4

[Oxidant] (mM)

[Ligand] (mM)

[Base] or [Acid] (mM)

a (slope)

k [Ligand]b (y intercept)

0.151.3
125
7.243

5
400
140

0.6
0.90
0.9

2.7

840 (NH3 )
5 (H2 SO4 )

intercept is comparable (2.72.8) in all three cases O2 /


CN , Fe(III)/SC(NH2 )2 and Cu(II)/S2 O2
3 /NH3 .
4.3. Surface reaction mechanism
It is well established that the oxidation of gold is
electrochemical in nature and thus the kinetics can be
modelled using corrosion theory for metals (Nicol,
1993). For example, the order close to 0.5 for the cyanidation reaction with respect to [O2 ] shown in Table 9
is consistent with the previous results and the electrochemical model (Crundwell and Godorr, 1997). The
detailed analysis of reaction orders based on rotating
disc data which provide useful information on reaction
order with respect to concentration of reagents is well
documented (Jiang et al., 1993; Li and Miller, 2002).
The leaching results reported by Navarro et al. (2002)
show that the initial % extraction of gold is independent
of the concentration of Cu(II) and S2 O2
3 at pH 10, 25
C and higher reagent concentrations: [Cu(II)] > 0.05 M
and [(NH4 )2 S2 O3 ] > 0.1 M. This result, consistent with
the zero order rate of gold dissolution with respect to
[Cu(II)] noted in Fig. 12, can be used to propose a
surface reaction mechanism for gold leaching with
Cu(II) by combining the electrochemical rate equations
with the well known adsorption theory.
Byerley et al. (1973) showed the importance of considering the mixed complex of Cu(II): Cu(NH3 )n S2 O3 in
order to rationalize the kinetics of oxidation of S2 O2
3 by
ammoniacal CuSO4 . Eq. (35) shows the equilibrium
which represents the adsorption of S2 O2
3 and the mixed
complex Cu(NH3 )n S2 O3 onto the gold surface to form
s Au(S2 O3 )2 Cu(NH3 )2
n ads . If the fraction of the gold
surface onto which S2 O2
and Cu(NH3 )n S2 O3 is ad3
sorbed is denoted by h, the adsorption equilibrium
constant Kads for Eq. (35) and h can be expressed by Eqs.
(36) and (37) respectively, where [Cu(II)] is the concentration of the mixed complex in solution.
s Au S2 O2
3 aq CuNH3 n S2 O3 aq
2

s AuS2 O3 2 CuNH3 n ads


surface adsorption equilibrium
Kads hf1  hCuII
S2 O2
3
g
h

Kads CuII
S2 O2
3
f1

1

35
36

1
Kads CuII
S2 O2
3
g

37
Due to the redox nature, the surface reaction for the
oxidation of gold by Cu(II) can be represented by the

2.8

simultaneous oxidation Au(0) Au(I) + e and reduction


Cu(II) + e Cu(I). For the sake of simplicity s Au(S2 O3 )2 Cu(NH3 )2
n ads is represented by s Au(0)Cu(II)ads
and the electrode reactions by Eqs. (38) and (39). Thus
the overall redox reaction described by Eq. (42) i.e. the
simplied version of the sum of Eqs. (38) and (39),
represents the apparent surface reaction which is also
the rate determining step (RDS).
s Au0CuIIads
s CuIIads AuIaq e

anodic reaction
38

s Au0CuIIads e
cathodic reaction

s Au0 CuIaq

39
2s Au0CuIIads s CuIIads s Au0
AuIaq CuIaq

overall

s Au0CuIIAds s AuIaq CuIaq

40
41

2

s AuS2 O3 2 CuNH3 n ads


s AuS2 O3 3
2 aq
CuNH3
n aq

rate determining step; RDS

42

According to the well known electrochemical kinetic


model (Nicol, 1993; Pesic and Sergent, 1993) the rate of
anodic (a) and cathodic (c) reactions are dependent on
the relevant electrochemical rate constants (ka and kc )
and the concentration of the electroactive species. In the
present situation the concentration of the electroactive
species can be represented by h; this leads to Ra ka h
and Rc kc h. However, the anodic and cathodic reactions take place at the same rate (Ra Rc R), leading
to Eq. (43). Thus the combination of the adsorption
theory with the rate equations reveals the relationship
between the rate ( reader is referred to other references
e.g. Pesic and Sergent, 1993, for details on electrochemical equations related to this derivation) of gold
oxidation and the rate constants for the anodic, cathodic
and overall (kRDS ) surface reaction described in Eq. (44).
The substitution for h from Eq. (37) leads to Eq. (45)
which shows the relationship between the rate and the
concentration of species in solution.
R 2 k a k c h2

43

R ka kc 0:5 h kRDS h

44

G. Senanayake / Minerals Engineering 17 (2004) 785801

R kRDS h
kRDS Kads CuII
S2 O2
3
=f1 Kads CuII

 S2 O2
3
g

45

R kRDS Kads CuII


S2 O2
3
kp CuII

46

R kRDS

47

At low reagent concentrations Kads [Cu(II)][S2 O2


3 ] may
be assumed to be negligibly small compared to 1 and
thus 1 Kads [Cu(II)][S2 O2
3 ] 1. This simplies Eq. (45)
to Eq. (46) which shows that the dissolution reaction is
pseudo rst order with respect to [Cu(II)] at constant
[S2 O2
3 ] and kp is the pseudo rst order rate constant.
This also agrees with Fig. 12 and the rate data based on
corrosion current reported by Jiang et al. (1993) for the
oxidation of a rotating gold disc, which led to a rst
order reaction with respect to [Cu(II)]. In contrast, at
higher concentrations of Cu(II) and/or S2 O2
where
3
1 Kads [Cu(II)][S2 O2
3 ]  1, Eq. (45) simplies to Eq.
(47). This explains why the initial rate is independent of
the concentration of Cu(II) and S2 O2
3 at higher reagent
concentrations (Fig. 12).
4.4. Leaching of gold ore
Tables 10 and 11 describe the mineralogical constituents, gold and silver grade, particle size and pulp
density of dierent material used by previous researchers, with a range of non-cyanide lixiviants and conditions. For example, the material used in test J12 was a
highly preg-robbing oxide ore of gold and silver grades
910 and <0.5 g/t respectively. Whilst 81.8% gold was
associated with carbonaceous material, only 0.5% gold
was recoverable by direct cyanidation (Ji et al., 2003). In
contrast, the material used in test N11 (Navarro et al.,
2002) was a sulphidic ore with much higher gold (95 g/t)
and silver (235 g/t) grades. Fig. 13 compares the fraction

797

of gold (X) leached from these two ore samples using


the two lixiviant systems CuSO4 /(NH4 )2 S2 O3 /pH 10/
25 C and O2 (100 psig)/Na2 S2 O3 /pH 12/40 C. Due to
the dierent types of material, gold grades, oxidants,
temperature and pH used in the two tests N11 and J12
(Tables 10 and 11), and the absence of Cu(II)/NH3 in the
latter, the fraction of gold extracted (X) vs time (t) follow slightly dierent trends in Fig. 13. It is of interest to
note that the results from test N11 obey the well known
(Sohn and Wadsworth, 1979) shrinking sphere model
1=3
1  1  X kss t, where kss is the apparent rate
constant for the rate controlling surface chemical reaction. The value of kss based on the slope appears to be of
the same order 105 s1 in the two tests (Fig. 13, Table
11).
Fig. 14 compares the fraction of gold leached from
gold bearing material, which contained predominantly
arsenopyrite (Kholmogorov et al., 2002), with the three
lixiviant systems SC(NH2 )2 , SCN or S2 O2
using
3
Fe(III) as the oxidant. Again, the rst two obey a
shrinking sphere model with kss 4  105 s1 and
3 105 s1 respectively (Table 11). The thiosulphate
system follows the same trend as thiocyanate during the
rst 2 h. Thus the surface reaction between gold and the
oxidant appears to be rate controlling in tests K1K3,
N11 and J12.
Fig. 15 compares the fraction of gold leached with
dierent material using a range of non-cyanide lixiviants
described in Tables 10 and 11. The three systems Fe(III)/

S2 O2
(tests K3,
3 , Fe(III)/SC(NH2 )2 and Fe(III)/SCN
U1, W1) show relatively slow rates of gold extraction
even at the initial stages of leaching, largely due to the
low temperature of 2025 C compared to 60 C in other
tests. The initial rate was higher in tests A3 and J11, and
the % Au extraction in 6 h was close to 80%. Amongst
the tests J4, A3, Z4 and B8 which used Cu(II)NH3
S2 O2
3 as the gold lixiviant at 60 C, Z4 and B8 showed
fast initial rates and gave over 90% gold extraction in 1 h,

Table 10
Description of material and conditions used in leaching tests (references in Table 11)
Test/material

Mineralogical constituents

Au g/t

Ag (g/t)

Size (lm)

PD (%)

B8/conc.
Z4/Mn ore
J4/oat conc.
A3/gold ore
N11/conc.

CuFeS2 , FeS2 , ZnS, FeS


Rhyolite, andesite

4.76
3
154
51.6
95

113
113c

24.3

44
74

40
40

4.7

J11-J12/oxide ore
K1K3/conc.
W1/bio-oxidized
U1/oxide ore
a

SiO2 , (K,Na,Al,MgSiAlOOH)
FeS, FeS2 , CuFeS2 ,
(Cu,Fe)12 As4 S13
SiO2 , Al2 O3 , Fe2 O3 , MgO, CaO
FeAsSd
a

SiO2 , CaO, Fe2 O3 , Al2 O3

910
110
2.13
4.1

<0.5

Cu (%)

S (%)

Fe (%)

1
235

3.7
50

40.4
0.39

74
3774

40
40

75b

33
1020
20
20

Not reported.
80% passing (P80 ).
c
Most of Ag associated with MnO2 .
d
Contained 0.10.2% iron sulphide and carbon containing compounds.
b

Cu (g/t)

17

3.17
0.18

3.23
3.82

75b
74

798

G. Senanayake / Minerals Engineering 17 (2004) 785801

Table 11
Leaching conditions and apparent rate constants (kss and kpl ) for gold extraction from dierent types of gold ores using non-cyanide lixiviantsa
Lixiviant/Test/Material

pH

T (C)

[Oxidant] (mM)

[Ligand] (M)

[NH3 ]t (M)

105 kss b (s1 )

0.67
4.12
4.41
4
(pH 10)

105 kpl c (s1 )

Cu(II)/S2 O2
3 /NH3
B8/conc.
Z4/Mn ore
J4/gold. conc.
A3/gold ore
N11/conc.
O2 /S2 O2
3
J11/oxide ore
J12/oxide ore
Fe(III)/S2 O2
3
K3/FeAsS conc.
Fe(III)/SCN
W1/bio-oxidized
K2/FeAsS conc.
Fe(III)/SC(NH2 )2
K1/FeAsS conc.
U1/oxide ore
Cu(II)/O2 /NH3
H1/sulphidic/refractory

10
8.510.5
10

60
60
60
60
25

46.5d
94
63
100
12

0.34
1.96
0.71
2
17

11
12

60
40

0.1
0.1

20

15

0.5

0.8

2
2.2

200
15

0.05
0.4

0.8

20

10

20
20

200

15
0.14 kg/t

0.2
100 g/kg

38

5.5

Information on mineralogical constituents shown in Table 10.


a
Refs.: A3 (Abbruzzese et al., 1995); B8 (Berezowsky and Sefton, 1979); J4 (Jiang et al., 1993); J11, J12 (Ji et al., 2003); K1, K2, K3 (Kholmogorov
et al., 2002); N11 (Navarro et al., 2002); U1 (Ubaldini et al., 1998), H1 (Han, 2001). PD 20%, pO2 600 kPa.
b
Slope of linear relationships in Figs. 13 and 14.
c
Slope of linear relationship in Fig. 16.
d
Air sparged.
e
pO2 100 psig.
f
0.2 M H2 SO4 .

1.0

1.0

0.8

y = 2E-05x
R2 = 0.912

0.6

X or 1-(1-X)

X or 1-(1-X)1/3

0.8

y = 2E-05x
R2 = 0.9977

0.4

K1 /thiourea (X)
K2 / thiocyanate (X)
K3 / thiosulphate (X)
K1 / shrinking sphere model
K2 / shrinking sphere model
K3 / shrinking sphere model

1/3

J12 (X)
J12 (shriking sphere model)
N11 (X)
N11 (shrinking sphere model)

y = 4E-05x
R2 = 0.9893

0.6

0.4

y = 3E-05x
R2 = 0.964

0.2
0.2

0.0
0

10000

20000

30000

40000

Time / s

0.0
0

5000

10000

15000

Time / s
2
Fig. 13. Gold extraction curves with Cu(II)/NH3 /S2 O2
3 and O2 /S2 O3
and applicability of the shrinking sphere model for tests N11 and J12
described in Tables 10 and 11.

showing the eect of the high copper and/or silver content of the material used in test Z4 and B8. Jiang et al.
(1993) achieved a 90% gold extraction in 4 h at 50 C,
compared to <60% in test J4, by using high [Cu(II)] of
0.31 M.
Despite the use of a high temperaturepressure oxidation in test H1 (100 kPa O2 /200C) with Cu(NH3 )2
4 ,
the initial rate was much lower than that in tests B8 and
Z4, leading to 75% Au extraction in 1 h. This clearly
shows the role of thiosulphate in tests B8 and Z4 compared to H1. Nevertheless, the overall gold extraction in

Fig. 14. Gold extraction curves from arsenpyrite concentrate using


Fe(III) in thiourea, thiocynate or thiosulphate media and applicability
of shrinking sphere model for Tests K1K3 described in Tables 10 and
11.

23 h was comparable in the two tests H1 and Z4,


highlighting the success with Cu(II)/NH3 /O2 -pressure
leaching. Fig. 16 plots 1  31  X 2=3 21  X for
tests W1 and K3 to examine the validity of the well
known shrinking sphere model with a solid layer:
2=3
1  31  X 21  X kpl t, where the diusion of
reactants or products through a solid product on the
surface is rate controlling. The apparent rate constant
kpl obtained from the slope is of the order 105 s1

G. Senanayake / Minerals Engineering 17 (2004) 785801


100%
B8

Z4

Gold Extraction

J11

H1

80%

60%
J4

W1
K3

40%

J11
J4
Z4
A3
B8
W1
K3
H1
U1

A3

U1

20%

0%
0

4
5
Time / hours

Fig. 15. Gold extraction curves from dierent materials described in


Tables 10 and 11 using dierent lixiviants at 60, 20 and 200 C.

0.2

2/3

1-3(1-X) +2(1-X)

W1

0.2

K3

Slope 8 x 10
R2 = 0.98

0.1

-6

0.1

799

temperature, but increases with increasing temperature.


Solubility of gold salts follow the order NaAuCl4 >
KAuCl4 > AuCl3  Au(NH3 )4 (NO3 )2 > Au2 S/H2 S >
Au(OH)3 /NaOH > Au(OH)3 .
Results from rotating disc studies in dierent lixiviants
at dierent temperatures reveal the order of rate of
gold oxidation: Br2 /Br (25 C) > Fe(III)/Cl /H2 SO4
(200 C) > NaOCl/Cl (20 C) > Cl2 /HCl/12 C > Cu(II)/NH3 /S2 O2
3 (20 C) > Fe(III)/SC(NH2 )2 (25 C)
> Cu(NH3 )2
/NaOCl
(140 C) > O2 /SC(NH2 )2 (30
4
C) > O2 /S2 O2
(30
C),
but do not reect the rate con3
stants of the surface reactions.
Silver dissolves faster than gold and thus Ag(I) catalyses the dissolution of gold by redox-displacement.
Whilst the surface reaction can be rationalized on the
basis of a combined adsorptionelectrochemical
model, the analysis of gold leaching in dierent lixiviant systems shows the validity of a shrinking particle
model with apparent rate constants (kss and kpl ) of the
order 105 s1 .

Acknowledgements

0.0
0

5000

10000

15000
Time / s

20000

25000

Fig. 16. Applicability of the shrinking sphere model with product layer
for gold extraction data in Tests W1 and K3 described in Tables 10 and
11.

(Table 11). These results highlight the importance of


considering the role of solid products on gold surface
that would retard the leaching reaction.

5. Summary and conclusions


Fundamental studies are vital for the rationalization
of the complexity of gold leaching in non-cyanide
media.
Linear free energy correlations of stability constants
of Au(I) complex species with respect to relevant species of Ag(I) and Cu(I) show the stability order of
Au(I) complexes with ligands: CN > HS > S2 O2
3 >
SC(NH2 )2 > OH > SCN > NH2 CH2 COO > SO2
3 >
NH3 > Br > Cl > CH3 CN. In the case of Au(III)
complexes the stability order is CN > OH >
SCN > Br > Cl .

Species such as AuOH0 , Au(OH)
2 and AuCl2 can be
incorporated as intermediates in EhpH and Eh
log[Cl ] diagrams.
Solubility of gold metal in the presence of oxidants
decreases in the order MgS/H2 S  Fe(III)/Cl >
Cu(II)/Cl > Cu(II)/NH3  NaOH/H2 O at a given

The author gratefully acknowledges the continuous


support from Professor M.J. Nicol (Murdoch University), Adjunct Professor D.M. Muir (CSIRO Minerals,
Perth) and the A.J. Parker Cooperative Research Centre
for Hydrometallurgy.

References
Abbruzzese, C., Fornari, P., Massidda, R., Vegli
o, F., Ubaldini, S.,
1995. Thiosulphate leaching for gold hydrometallurgy. Hydrometallurgy 39, 265276.
Anderson, C.G., 2003. Alkaline sulde recovery of gold utilizing
nitrogen species catalyzed pressure leaching. In: Young, C.A.,
Alfantazy, A.M., Anderson, C.G., Dreisinger, D.B., Harris, B.,
James, A. (Eds.), Hydrometallurgy 2003Fifth International
Conference in Honor of Prof. Ian Ritchie, vol. 1. TMS, pp. 75
87 and references therein.
Aylmore, M.G., Muir, D.M., 2001. Thiosulfate leaching of goldA
review. Minerals Engineering 14, 135174.
Bard, A.J., 1973. Encyclopedia of Electrochemistry of Elements.
Marcel Dekker, New York.
Belevantsev, V.I., Peshchevitsky, B.I., Shamovskaya, G.I., 1981.
Gold(I) sulde complexes in aqueous solutions. Isvest. Sib. Otd.
Akad. Nauk. SSR. Ser. Khim. 1, 8187.
Benning, L.G., Seward, T.M., 1996. Hydrosulphide complexing of
Au(I) in hydrothermal solutions from 150400 C and 5001500
bar. Geochimica et Cosmochimica Acta 60, 18491871.
Berezowsky, R.M.G.S., Sefton, V.B., 1979. Recovery of gold and silver
from oxidation leach residues by ammoniacal thiosulphate leaching. Paper presented at AIME annual Meeting, New Orleans, LA,
pp. 102105.
Byerley, J.J., Fauda, S.A., Rempel, G.L., 1973. Kinetics and mechanism of the oxidation of thiosulphate ions by copper(II) ions in

800

G. Senanayake / Minerals Engineering 17 (2004) 785801

aqueous ammonia solutions. Journal of Chemical Society, Dalton


Transactions, 889893.
Chandra, I., Jerey, M., 2003. Can a thiosulfate leaching process be
developed which does not require copper and ammonia? In:
Young, C.A., Alfantazy, A.M., Anderson, C.G., Dreisinger, D.B.,
Harris, B., James, A. (Eds.), Hydrometallurgy 2003Fifth International Conference in Honor of Prof. Ian Ritchie, vol. 1. TMS,
pp. 169182.
Chai, L., Okido, M., Wei, W., 1999. Eect of Na2 SO3 on electrochemical aspects of gold dissolution in alkaline thiourea solution.
Hydrometallurgy 53, 255266.
Chen, C.K., Lung, T.N., Wan, C.C., 1980. A study of the leaching of
gold and silver by acidothioureation. Hydrometallurgy 5, 207
212.
Crundwell, F.K., Godorr, S.A., 1997. A mathematical model of the
leaching of gold in cyanide solutions. Hydrometallurgy 44, 147
162.
Dasgupta, R., Guan, Y.C., Han, K.N., 1997. The electrochemical
behavior of gold in ammoniacal solutions at 75 C. Metallurgical
and Materials Transactions 28B, 512.
Desch^enes, G., Ghali, E., 1988. Leaching of gold from a chalcopyrite
concentrate by thiourea. Hydrometallurgy 20, 179202.
Diaz, M.A., Kelsall, G.H., Welham, N.J., 1993. Electrowinning
coupled to gold leaching by electrogenerated chlorine I. Au(III)
Au(I)/Au kinetics in aqueous Cl2 /Cl electrolytes. Journal of
Electroanalytical Chemistry 361, 2538.
Ferron, C.J., Fleming, C.A., Dreisinger, D., OKane, T., 2003.
Chloride as an alternative to cyanide for the extraction of gold
going full circle? In: Young, C.A., Alfantazy, A.M., Anderson,
C.G., Dreisinger, D.B., Harris, B., James, A. (Eds.), Hydrometallurgy 2003Fifth International Conference in Honor of Prof. Ian
Ritchie, vol. 1. TMS, pp. 89104.
Finkelstein, N.P., Hancock, R.D., 1974. A new approach to the
chemistry of gold. Gold Bulletin 7, 7277.
Fleet, M.E., Knipe, S.W., 2000. Solubility of native gold in HOS
uids at 100400 C and high H2 S content. Journal of Solution
Chemistry 29, 11431157.
Groenewald, T., 1976. The dissolution of gold in acidic solutions of
thiourea. Hydrometallurgy 1, 277290.
Grosse, A.C., Dicinoski, G.W., Shaw, M.J., Haddad, P.R., 2003.
Leaching and recovery of gold using ammonium thiosulfate leach
liquors (a review). Hydrometallurgy 69, 121.
Han, K.N., 2001. Electrochemical behavior of the dissolution of gold
in ammoniacal solutions. In: Young, C.A., Twidwell, L.G.,
Anderson, C.G. (Eds.), Cyanide: Social, Industrial and Economic
Aspects. TMS, pp. 485499.
Hancock, R.D., Finkelstein, N.P., Evers, A., 1974. Linear free energy
relationships in aqueous complex-formation reactions of the d10
metal ions. Journal of Inorganic and Nuclear Chemistry 36, 2539
2543.
Hiskey, J.B., 1984. Thiourea leaching of gold and silver technology
update and additional applications. Minerals and Metallurgical
Processing 1, 173179.
Hiskey, J.B., Atluri, V.P., 1988. Dissolution chemistry of gold and
silver in dierent lixiviants. Mineral Processing and Extractive
Metallurgy Reviews 4, 95134.
Hogfeldt, E., 1982. Stability constants of metal-ion complexes, 2nd
supplement, IUPAC Chemical data series no. 2, part A, Inorganic
ligands, Oxford, Pergamon.
Hunter, R.M., Stewart, F.M., Darsow, T., Fogelsong, M.L., Mogk,
D.W., Abbott, E.H., Young, C.A., 1996. A new alternative to
cyanidation: Biocatalyzed bisulde leaching. In: Proceedings of the
Third International Conference on Minerals Bioprocessing and
Biorecovery/Bioremediation in Mining, Big Sky, Montana, August,
pp. 2530.
Jerey, M.I., 2001. Kinetic aspects of gold and silver leaching in
ammoniathiosulfate solutions. Hydrometallurgy 60, 716.

Jerey, M.I., Breuer, P.L., Choo, W.L., 2001. A kinetic study that
compares the leaching of gold in the cyanide, thiosulfate and
chloride systems. Metallurgical and Material Transactions 32B,
979986.
Jiang, T., Chen, J., Xu, S., 1993. A kinetic study of gold leaching with
thiosulfate. In: Hiskey, J.B., Warren, G.W. (Eds.), Hydrometallurgy, Fundamentals, Technology and Innovations. AIME, pp.
119126. Chapter 7.
Ji, J., Fleming, C., West-Sells, P.G., Hackl, R.P., 2003. A novel
thiosulfate system for leaching gold without the use of copper and
ammonia. In: Young, C.A., Alfantazy, A.M., Anderson, C.G.,
Dreisinger, D.B., Harris, B., James, A. (Eds.), Hydrometallurgy
2003Fifth International Conference in Honor of Prof. Ian
Ritchie, vol. 1. TMS, pp. 227244.
Jiayong, C., Tong, D., Guocai, Z., Jin, Z., 1996. Leaching and
recovery of gold in thiosulfate based systemA research summary
at ICM. Transactions of Indian Institute of Metallurgy 49, 841
849.
Johnson, P.R., Pratt, J.M., Tilley, R.I., 1978. Experimental determination of the standard reduction potential of gold(I) ion. Journal of
Chemical Society Chemical Communications, 606607.
Ketcham, V.J., OReilly, J.F., Vardill, W.D., 1993. The Lihir gold
project; process plant design. Minerals Engineering 6, 10371065.
Kholmogorov, A.G., Kononova, O.N., Pashkov, G.L., Kononov,
Y.S., 2002. Thiocyanate solutions in gold technology. Hydrometallurgy 64, 4348.
Kissner, R., Welti, G., Geier, G., 1997. The hydrolysis of gold(I) in
aqueous acetonitrile solutions. Journal of Chemical Society,
Dalton Transactions, 17731777.
La Brooy, S.R., Linge, H.G., Walker, G.S., 1994. Review of gold
extraction from ores. Minerals Engineering 7, 12131241.
Li, J., Miller, J.D., 2002. Reaction kinetics for gold dissolution in acid
thiourea solution using formamidine disulde as oxidant. Hydrometallurgy 63, 215223.
Linkey, W.F., Seidell, A., 1958. Solubility of Inorganic and Metal
Organic Compounds, fourth ed. Van Nostrand, Princenton, NJ.
Liu, J.Q., Nicol, M.J., 2002. Thermodynamics and kinetics of the
dissolution of gold under pressure oxidation conditions in the
presence of chloride. Canadian Metallurgical Quarterly 41, 409
416.
Lorenzen, L., Van Deventer, J.S.J., 1993. The identication of
refractoriness in gold ores by the selective destruction of minerals.
Minerals Engineering 6, 10131023.
MacArthur, J.S., Forrest, R.W., Forrest, W., 1888. British Patent No.
14, 174.
McDonald, G.W., Saud, A., Barger, M.S., Koutsky, J.A., Langer,
S.H., 1987. The fate of gold in cupric chloride hydrometallurgy.
Hydrometallurgy 18, 321336.
Meng, X., Han, K.N., 1996. The principals and applications of
ammonia leaching of metalsA review. Mineral Processing and
Extractive Metallurgy Review 16, 2361.
Michel, D., Frenay, J., 1999. Integration of amino acids in the
thiosulfate gold leaching process. In: Proceedings of the Randol
Gold and Silver Forum. Randol International, pp. 99103.
Miller, G.C., Pritsos, C.A., 2001. Unresolved problems with the use of
cyanide in open pit precious metals mining. In: Young, C.A.,
Twidwell, L.G., Anderson, C.G. (Eds.), Cyanide: Social, Industrial
and Economic Aspects. TMS, pp. 7381.
Molleman, E., Dreisinger, D., 2002. The treatment of coppergold ores
by ammonium thiosulfate leaching. Hydrometallurgy 66, 121.
Muir, D.M., 2002. Basic principals in chloride hydrometallurgy. In:
Peek, E., Van Weert, G. (Eds.), Chloride Metallurgy 2002. CIM,
Montreal, pp. 255276.
Muir, D.M., Aylmore, M., 2002. Thiosulphate as an alternative to
cyanide for gold processingissues and impediments. In: Proceedings of the Green Processing, 2002, AusIMM, Cairns, QLD,
pp. 125134.

G. Senanayake / Minerals Engineering 17 (2004) 785801


Navarro, P., Vargas, C., Villarroel, A., Alguacil, F.J., 2002. On the use
of ammoniacal/ammonium thiosulphate for gold extraction from a
concentrate. Hydrometallurgy 65, 3742.
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 in South Africa. SAIMM, Johannesburgh,
RSA, pp. 831905.
Nicol, M.J., 1993. The role of electrochemistry in hydrometallurgy. In:
Hiskey, J.B., Warren, J.B. (Eds.), Hydrometallurgy Fundamentals,
Technology and Innovation. The Metallurgical Society, Littleton,
CO, pp. 4362.
Nicol, M.J., OMalley, G.P., 2001. Recovery of gold from thiosulfate
solutions and pulps with ion-exchange resins. In: Young, C.A.,
Twidwell, L.G., Anderson, C.G. (Eds.), Cyanide: Social, Industrial
and Economic Aspects. TMS, New Orleans, Louisiana, pp. 469
483.
Pangum, L.S., Browner, R.E., 1996. Pressure chloride leaching of a
refractory gold ore. Minerals Engineering 9, 547556.
Pesic, B., Sergent, R.H., 1993. Reaction mechanism of gold dissolution
with bromine. Metallurgical Transactions 24B, 419431.
Pouradier, J., Gadet, M.-C., 1969. Electrochemie des sels dor VIIAurothiosulfate. Journal De Chimie Physique 66, 109112.
Renders, P.J., Seward, T.M., 1989. The stability of hydrosulphido-and
sulphido-complexes of Au(I) and Ag(I) at 25 C. Geochimica et
Cosmochimica Acta 53, 245253.
Ritchie, I.M., Nicol, M.J., Staunton, W.P., 2001. Are there realistic
alternatives to cyanide as a lixiviant for gold at the present time. In:
Young, C.A., Twidwell, L.G., Anderson, C.G. (Eds.), Cyanide:
Social, Industrial and Economic Aspects. TMS, New Orleans,
Louisiana, pp. 427440.
Schwarzenbach, G., Widmer, M., 1966. Die L
oslichkeit von Metallsulden II. Silbersuld [1]. Helvetica Chimica Acta 49, 111123.
Senanayake, G., Muir, D.M., 2003. Chloride processing of metal
sulphides: Review of fundamentals and applications. In: Young,
C.A., Alfantazy, A.M., Anderson, C.G., Dreisinger, D.B., Harris,
B., James, A. (Eds.), Hydrometallurgy 2003Fifth International
Conference in Honor of Prof. Ian Ritchie, vol. 1. TMS, pp. 517
531.
Senanayake, G., Perera, W.N., Nicol, M.J., 2003. Thermodynamic
studies of the gold (III)/(I)/(0) redox system in ammoniathiosulphate solutions at 25 C. In: Young, C.A., Alfantazy, A.M.,
Anderson, C.G., Dreisinger, D.B., Harris, B., James, A. (Eds.),
Hydrometallurgy 2003Fifth International Conference in Honor
of Prof. Ian Ritchie, vol. 1. Leaching and Solution Purication,
TMS, pp. 155168.
Sharpe, A.G., 1976. The Chemistry of Cyano Complexes of the
Transition Metals. Academic Press, London. pp. 265285.
Shenberger, D.M., Barnes, H.L., 1989. Solubility of gold in aqueous
sulde solutions from 150 to 350 C. Geochimica et Cosmochimica
Acta 53, 269278.

801

Skibsted, L.H., Bjerrum, J., 1974a. Studies on gold complexes. I.


Robustness, stability and acid dissociation of the tetramminegold(III) ion. Acta Chemica Scandinavica A 28, 740746.
Skibsted, L.H., Bjerrum, J., 1974b. Studies on gold complexes. II. The
equilibrium between gold(I) and gold(III) in the ammonia system
and the standard potentials of the couples involving gold,
diamminegold(I), and tetramminegold(III). Acta Chemica Scandinavica A 28, 764770.
Skibsted, L.H., Bjerrum, J., 1977. Studies on gold complexes. III. The
standard electrode potentials of aqua gold ions. Acta Chemica
Scandinavica A 31, 155156.
Sohn, H.Y., Wadsworth, M.E., 1979. Rate Processes of Extractive
Metallurgy. Plenum Press, New York.
Sparrow, G.J., Woodcock, J.T., 1995. Cyanide and other lixiviant
leaching systems for gold with some practical applications. Mineral
Processing and Extractive Metallurgy Reviews 14, 193247.
Stefansson, A., Seward, T.M., 2003. The hydrolysis of gold(I) in
aqueous solutions to 600 C and 1500 bar. Geochimica et
Cosmochimica Acta 67, 16771688, and references therein.
Tran, T., Lee, K., Fernando, K., 2001. Halide as an alternative
lixiviant for gold processingan update. In: Young, C.A.,
Twidwell, L.G., Anderson, C.G. (Eds.), Cyanide: Social, Industrial
and Economic Aspects. TMS, pp. 501508.
Ubaldini, S., Fornari, P., Massidda, R., Abbruzzese, C., 1998. An
innovative thiourea gold leaching process. Hydrometallurgy 48,
113124.
Wan, R.Y., Brierley, J.A., Acar, S., LeVier, K.M., 2003. Using
thiocyanate as lixiviant for gold recovery in acidic environment. In:
Young, C.A., Alfantazy, A.M., Anderson, C.G., Dreisinger, D.B.,
Harris, B., James, A. (Eds.), Hydrometallurgy 2003Fifth International Conference in Honor of Prof. Ian Ritchie, vol. 1. TMS,
pp. 105121.
Webster, J.G., 1986. The solubility of gold and silver in the system Au
AgSO2 H2 O at 25 C and 1 atm. Geochimica et Cosmochemica
Acta 50, 18371845.
West-Sells, P.G., Ji, J., Hackl, R.P., 2003. A process for counteracting
the detrimental eect of tetrathionate on resin gold adsorption
from thiosulfate leachates. In: Young, C.A., Alfantazy, A.M.,
Anderson, C.G., Dreisinger, D.B., Harris, B., James, A. (Eds.),
Hydrometallurgy 2003Fifth International Conference in Honor
of Prof. Ian Ritchie, vol. 1. TMS, pp. 245256.
Xia, C., Yen, W.T., Deschenes, G., 2003. Improvement of thiosulphate
stability in gold leaching. Minerals and Metallurgical Processing
20, 6872.
Young, C.A., Dahlgren, E.J., Robins, R.G., 2003. The solubility of
copper suldes under reducing conditions. Hydrometallurgy 68,
2331.
Zipperian, D., Raghavan, S., Wilson, J.P., 1988. Gold and silver
extraction by ammoniacal thiosulphate leaching from a rhyolite
ore. Hydrometallurgy 19, 361375.

You might also like