Você está na página 1de 36

llydrometal/urgy, 30 ( 1992) 127-162

Elsevier Science Publishers B. Y., Amsterdam


Hydrometallurgy of precious metals recovery
ABSTRACT
C.A. Fleming
Lakefield Research. Lakefield. Onto KOL 2110. Canada
(Revised version accepted January 10, 1992)
127
Fleming, C.A., 1992. Hydrometallurgy of precious metals recovery. In: W.e. Cooper and D.B. Dreis-
inger (Editors). Hydrometallurgy, Theory and Practice. Proceedings of the Ernest Peters Interna-
tional Symposium.llydrometal/urgy, 30: 127-162.
The hydro metallurgical process for the treatment of gold and silver ores remained unchanged for
the first 70 years of this century, and consisted essentially ofJeaching in cyanide solution followed by
solid-liquid separation, with the solid residues being washed as efficiently as possible, and the leach
liquor being treated by zinc cementation to recover the precious metals. While this process is generally
extremely efficient and fairly cheap, it does have limitations in the treatment of low-grade ores and
certain complex ore types. For example, ores with a high content of clay or other soft, fine minerals
are usually difficult to filter. and losses of soluble gold or silver in the residues can be unacceptably
high. In other situations, where the precious metal host rock contains high concentrations of sulphides
such as pyrite or arsenopyrite, for example. or base-metal oxides or carbonates, the traditional process
often suffers from poor gold recovery (due to encapsulation of the precious metals in the sulphides)
or high cyanide consumption, or both of these. Whereas these occurrences were fairly rare (or were
avoided!) in the first half of this century, they are now assuming great importance, and each year a
higher percentage of world gold production derives from sources such as these.
A number of new hydrometallurgical processes have been developed and implemented in the gold
industry in the last 20 years. and these have transformed gold processing into a chemical "high tech"
industry. and have allowed increasingly complex ore types and progressively lower grades of ore to be
treated economically. As a result, in a period when gold production might have been expected to
decline, world-wide production has almost doubled over the last two decades.
This paper describes the traditional cyanidation and zinc cementation processes, but focuses on the
new developments in the industry. In particular, new leaching technologies such as heap leaching for
low-grade ores and pressure leaching for refractory sulphide ores are discussed, as well as the carbon-
in-pulp and carbon-in-Ieach processes that have effectively replaced filtration and countercurrent de-
cantation on almost every gold plant built since 1980. Some emerging technologies such as bacterial
leaching and resin-in-pulp are also discussed briefly.
INTRODUCTION
Even since the development of the cyanide/zinc cementation process in the
late 19th century, hydro metallurgy has been a dominant force in the process-
ing of gold and silver ores. Indeed, the cyanidation process undoubtedly qual-
ifies even today as the most important and widespread of the various hydro-
Correspondence to: C.A. Fleming, Lakefield Research, Hydrometallurgy Lakefield, Onto KOL
2HO, Canada.
0304-386X/92/$05.00 1992 Elsevier Science Publishers B. Y. All rights reserved.
128 CA. FLEMING
metallurgical technologies used in the treatment of primary ores and
concentrates. The process is so simple, efficient and inexpensive, in fact, that
very few technological changes or innovations took place in the industry dur-
ing the first 60-70 years of the 20th century. Most of the ores treated in this
period, particularly those from the Witwatersrand in South Africa (where al-
most half the gold ever mined has been recovered), were of the free-milling,
quartzitic type, and it was usually possible to achieve good gold liberation
merely by crushing and milling the host rock adequately. Processing of the
milled ore by a combination of gravity separation (to recover coarse gold
particles) and cyanidation generally yielded overall gold recoveries well in
excess of 90%. The only inherent inefficiency in this process was associated
with the separation of the gold-bearing cyanide solution from the solid leach
residue. Consequently, this was the only area that benefitted significantly from
new technology, with the replacement of cumbersome decantation/washing
systems by more efficient, continuous processes, such as vacuum filtration
and countercurrent thickening/washing.
Several factors combined in the late 1960s and early 1970s to trigger the
technological revolution that has transformed the gold industry over the last
20 years. Firstly, and perhaps most importantly, the price of gold was released
in 1968 from the rigid constraints of the $35/ oz gold standard, and was al-
lowed for the first time in modern history to find its own value in response to
free-market supply and demand forces. This resulted in a ten-fold increase in
the price of gold in a remarkably short period of time and this, in turn, had
several important spin-offs. Firstly, the profits realised by established gold-
mining companies increased greatly and, consequently, there was more to
spend on research and development and exploration for new deposits. Sec-
ondly, it became economically viable to treat lower-grade deposits, and eco-
nomically expedient to maximize extraction efficiencies, since even minor
improvements in recovery now had major economic significance. Thirdly,
many new gold mining companies were born in the United States, Canada,
Australia, the Pacific Rim, South America and Africa, and they brought with
them a refreshingly innovative and entrepreneurial spirit, into an industry
that had become so conservative and resistant to change as to be almost
moribund.
The second force that has played a major role in the technological revolu-
tion has been the natural process of depletion of free-milling, easily extracta-
ble, gold resources around the world. In most of the important new deposits
that are being found today, the gold is very finely disseminated and encapsu-
lated in a host matrix that is inert and impermeable to the cyanide leach
solution. In these cases, it is necessary to first break down the host matrix
(usually by hydrometallurgical processing) and only then is it possible to leach
the liberated gold. There have been many technical innovations in this area,
and the issue facing those involved in the processing of these so-called refrac-
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 129
tory gold deposits is not so much whether or not the gold can be recovered
efficiently, but rather, which one of several processing options should be
adopted.
A third force has entered the primary precious metal mining industry and,
depending on the perspective, this force either hangs as a sword of Damocles
( over the developers of new properties) or opens up as a window of oppor-
tunity (for those involved in the development of new technology). This force
is the environmental legislation that is increasingly regulating and controlling
the precious metals industry, particularly in the First World. Environmental
issues are already having a significant influence on process flowsheet selec-
tion and treatment costs, and will probably be the major stimulus for the de-
velopment of new technology in the gold and silver mining industries over
the next 10-20 years.
THE CONVENTIONAL CYAN IDA TION /CEMENTA TION PROCESS
Practical aspects
The first step in the cyanidation process consists of crushing and milling
the gold-bearing ore to a size in which the gold particles are liberated from the
host rock or mineral. For practical as well as environmental reasons, the mill-
ing is performed on a wet slurry of the ore, usually at a low solids density, in
the 5-15% range, and in closed-circuit, with classification by either cyclones
or screens. The milled product is then thickened to between 30 and 50% solids
(depending on the rheology of the slurry), and advances to the cyanide-leach-
ing plant.
Oxygen is introduced to the slurry, usually in the form of compressed air
that is blown from the base of a conical-bottomed mixing tank (called a Pa-
chuca). The air, therefore, provides not only the oxygen that is necessary to
oxidize gold from the metallic state, in which it occurs in nature, to the gold
(I) state, but also generates the hydrodynamic conditions needed to keep the
slurry fully in suspension. In recent years, mechanically agitated, flat-bottomed
leaching tanks have been preferred to Pachucas, particularly since the intro-
duction of the carbon-in-Ieach process, and in certain instances pure oxygen,
or even stronger oxidants such as hydrogen peroxide, have been used in place
of air. .
The slurry flows by gravity through a series of tanks, numbering anywhere
from 6 to 20, with a residence time of 1-4 h in each (ank, with the overall
residence time being determined by the rate of gold leaching from the partic-
ular ore. The various factors affecting the rate ofleaching are discussed briefly
below, but the residence time is generally about 24 h for a free-milling ore.
Cyanide is usually only added to the second tank in the series, to ensure that
the slurry is thoroughly preaerated and the leach solution saturated with ox-
ygen prior to contact with cyanide. Modern gold plants monitor and control
130 C.A. FLEMING
the concentrations of cyanide and oxygen in the leach solution throughout the
leaching plant, to ensure consistent performance and to minimize cyanide
consumption.
The third step in the process involves separating the gold-bearing leach so-
lution from the gold-depleted ore. This operation has always been costly and
inefficient, and was an important stimulus for the development and rapid
implementation of the carbon-in-pulp and carbon-in-leach processes. The
methods used are either vacuum filtration (drum or belt filters) or counter-
current decantation (CCD) and, in both cases, inefficiencies stem from the
difficulty in washing the last traces of gold-bearing solution from the solids
without creating water-balance problems in the overall circuit. In general,
about 1.5-1.8 m
3
of gold-bearing solution is produced for every ton of ore
milled, and up to 2% of the leached gold can be lost in the final leach residue.
Once separated from the washed solids, the filtrate or gold pregnant solu-
tion is first clarified to remove the last traces of suspended solids, and is then
de-gassed under negative pressure to eliminate oxygen from solution. This
operation, which is performed in a Crowe tower, is an important step in the
overall process, because oxygen would otherwise participate in wasteful side-
reactions in the next step in the process. In this next step, gold is recovered
from the pregnant solution by reducing it back to the metallic state. A number
of powerful reductants, such as sodium borohydride and aluminium metal
have been shown to be technically viable in this role. When cost and effi-
ciency are considered, however; nothing compares with zinc metal, and all
conventional gold plants around the world employ zinc cementation to re-
cover gold from the cyanide leach liquor. An excess amount of zinc is re-
quired for efficient gold recovery, and the precipitate therefore consists of a
mixture of gold and zinc metal. This precipitate is recovered by filtration and
the barren filtrate is recycled as wash water to the filtration or CCD plant. In
the final step in the process, the excess zinc in the zinc/gold sludge is either
converted to zinc oxide in a calcining furnace, or is dissolved in a mineral
acid, prior to smelting the gold powder to bullion.
A full description of the conventional process is obviously beyond the scope
of this review, and this quick walk through the flowsheet is intended merely
as an introduction to the steps in the process where hydrometallurgy is fea-
tured. The chemistry involved in these steps is discussed in more detail in the
next section. Readers who are more interested in the practical aspects of the
conventional gold process are referred to the recent reviews by Young [1, pp.
277-330] and Bosley [2,pp. 331-334].
The chemistry of cyanidation
The action of alkaline cyanide solution on gold metal was first described in
a patent by Elkington in 1840, although this was in an electroplating context,
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 131
and did not refer to primary ore treatment. More than 40 years were to pass
before the Forrests and MacArthur conceived the idea of applying this reac-
tion to the dissolution of gold and silver from low-grade ores, and were
awarded a British Patent (No. 13,174) on cyanide leaching in 1887, and an-
other on the use of zinc dust for precipitation, in 1888. The first commercial
use of cyanide for gold extraction was at the Crown Mine in New Zealand in
1889, and applications in South Africa and the United States followed soon
after.
The chemistry of the dissolution of gold and silver in alkaline cyanide so-
lutions has been the subject of considerable investigation ever since these first
practical applications. Although the overall features of the reactions involved
are well-established, there is still uncertainty regarding the details of some
aspects of the mechanism of the process. Undoubtedly, the major advance in
the understanding of cyanidation was made by Kudryk and Kellogg [3], who
conducted experiments to demonstrate that the dissolution of gold in cyanide
solutions is essentially an electrochemical process, with the following overall
stoichiometry:
(1)
This reaction consists of cathodic and anodic half-reactions. The anodic re-
action involves the oxidation of gold (0) to gold (I):
4Au+8CN-=4Au(CN)i +4e (2)
and is accompanied by the cathodic reduction of oxygen at the surface of the
gold particle:
O
2
+2H
2
0+4e=40H- (3)
Those interested in detailed mechanistic studies of the gold cyanidation
reaction are referred to the definitive studies of Cathro and Koch [4], Fink
and Putnam [5] and Nicol [6]. The remarkable success of cyanide as a lixi-
viant for gold can be traced at least partially to the enormous stability of the
dicyanoaurate ion (P2 = 10
38
, [7]). This allows gold to be leached efficiently
at very low concentrations of cyanide 0.0 1 mol dm - 3), and the di-
cyanoaurate complex remains in alkaline solution even when the free-cyanide
concentration falls to zero. This contributes to the selectivity of the process,
because the stability of most of the other metal cyanocomplexes that are pres-
ent in cyanide leach solutions are somewhat lower than aurocyanide. This in
turn benefits the economics of the process, since the consumption of cyanide
via other metal complexation reactions is usually insignificant. The fact that
oxygen is an effective oxidant for gold metal in cyanide solution, even at the
very low concentrations that are typical of air-saturated aqueous solutions
(1.5-3XIO-
4
mol dm-
3
, i.e. 5-10 mg I-I) can also be traced to the great
stability of the dicyanoaurate ion.
132 C.A. FLEMING
In contrast to cyanide, all the alternative leaching agents that have been
proposed (halides, thiourea, thiosulphate, thiocyanate) yield gold complexes
that are considerably less stable than aurocyanide (Table 1). Therefore, in
order to stabilize these complexes in aqueous solution, and to achieve accept-
able leaching rates, it is necessary to employ far higher lixiviant concentra-
tions (usually 0.1-1 mol dm-
J
), and this, in tum, means that it is necessary
to recycle spent leaching solution if unacceptable costs and environmental
problems are to be avoided. Oxygen is also unsuitable as the oxidant for gold
in these situations, because of its low concentration and the limitation this
places on the rate of leaching. It is therefore necessary to resort to more ex-
pensive oxidants, such as hydrogen peroxide, chlorine, bromine, ozone or the
ferric ion. When all of these factors are considered, it is not surprising that
cyanide leaching has dominated gold processing for over 100 years, and will
probably continue to do so in the future, despite the fact that the cyanide ion
is itself now coming under the close scrutiny of environmental legislators.
There are, however, potential niche technologies for the alternati ve lixi viants,
and these will be discussed below.
Another important advantage of cyanide over most of the other leaching
chemistries (thiosulphate is the only exception) is the fact that the reaction
takes place in an alkaline environment. The dissolution of gold requires the
cyanide to be present as the free cyanide ion, CN-. Therefore, since the com-
position of cyanide solution is determined by the hydrolysis reaction:
(4)
which has a pKa of about 9.3, a high pH (usually in the 10.5-11 range) is
necessary to ensure that most of the cyanide is in the ionic form. A high pH is
also necessary for safety and economic reasons, as HCN is a volatile and poi-
sonous gas, which is purged from the leach slurry during vigorous air agita-
tion. An important advantage of leaching under alkaline conditions is that
TABLE I
Stability constants for a selection of complexes of gold (I) and gold (III) [8]
Gold(l)
fJ2
Gold (III )
P4
complex complex
Au(CN)i 2x 10
38
Au(CN)4"
_10
56
5x 10
28
AuI4" 5x 10
47
Au(CS(NH
2
h)i 2x 1023
Au(SCN)4" 10
42
Auli 4x 10
19
AuBr4"
1032
Au(SCN)i 1.3 X 10
17
AuCI4 10
26
AuBri 10
12
AuCli 10
9
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 133
the dissolution of base metals such as copper, zinc and nickel is substantially
reduced, resulting in cleaner effiuents than those generally produced in acid-
leach systems. The non-corrosive nature of an alkaline cyanide solution also
means that cheaper materials can be used for the construction of a gold plant.
The pH is regulated by the addition oflime to the leach slurry, usually in the
milling circuit.
The one drawback of the cyanide/oxygen leaching chemistry for the disso-
lution of gold is the low solubility of oxygen in water. It can be shown theo-
retically [8] that, under most conditions, the rate of gold leaching is
controlled by diffusion of oxygen to the metal surface. Factors that playa role,
therefore, in the rate of gold dissolution are:
( 1) The concentration of oxygen in water. This, in turn, is a function of air-
pressure, and oxygen solubility varies with altitude in the 5-10 mg I - I range
for air-saturated solutions. Temperature also plays a role, with oxygen solu-
bility varying inversely with temperature. An increase in temperature, there-
fore, can result in a decrease in the rate of gold leaching, despite the fact that
the diffusivity of oxygen improves. There are several methods of increasing
oxygen concentration. One is to employ pressure cyanide leaching [9], as
practised in a pipe-reactor at the Consolidated Murchison gold/antimony
mine in South Africa [10]. The high leaching rates observed in the Kamyr
leaching tower can probably also be traced to the relatively high pressures that
are attained at the bottom of a tall tower, and oxygen concentrations in the
range of 20-30 mg 1- I have been reported for this process [11 ]. Another is
to use oxygen instead of air to increase the partial pressure of oxygen under
atmospheric conditions. Alternatively, if the economics warrant it, hydrogen
peroxide can be used to supply oxidant to the surface of the gold particle, as
proposed by Loroesch [12].
(2) The dispersion of air in the slurry. This, in turn, is determined by fac-
tors such as the viscosity of the slurry and the method by which air is physi-
cally introduced into the slurry.
(3) The mass transfer of oxygen in the slurry. This is a function of the hy-
drodynamic conditions in the mixing tank, and varies with slurry rheology,
mixer design and energy input.
( 4) The presence in the slurry of parasitic oxygen-consuming reactions. These
species vary from naturally occurring organic molecules (humates, lipids) to
inorganic compounds such as the ferrous ion, and their presence can seriously
retard the rate of gold leaching. This is usually countered by employing one
or more stages of air (or oxygen, or hydrogen peroxide) pretreatment, prior
to the introduction of cyanide.
Cyanide concentration is less important than oxygen concentration, and it
can be shown theoretically [8] that the rate of gold leaching is independent
of free cyanide concentration at levels above about 2.5 X 10 -3 mol dm -3 (60
mg I-I CN - ). At concentrations below 2.5 X 10 - 3 mol dm - 3, the rate of
134 C.A. FLEMING
gold leaching is controlled by the mass transfer of cyanide ion to the surface
of the gold particles, rather than by mass transfer of oxygen. This is generally
only a factor when there are cyanide-consuming constituents other than gold
in the ore. Examples include iron sulphide minerals such as pyrrhotite (which
consumes both oxygen and cyanide), copper minerals such as chalcocite,
malachite and azurite, and various arsenic and antimony sulphides. A num-
ber of sulphur-containing inorganic anions such as sulphide, thiosulphate and
polythionate also react with and consume oxygen and cyanide.
Although much work has been done in this area, particularly in attempts to
reduce the reactivity of certain copper minerals in cyanide solution, little suc-
cess has been achieved, and the only method of ensuring consistently good
gold recovery in these situations is to employ high cyanide concentration and
tolerate high cyanide consumption.
The final factor that plays a role in the rate of leaching of gold from free-
milling ores is the size of the gold particles themselves. It can be shown that,
under conditions of oxygen mass-transfer rate control (in air-saturated solu-
tion), gold will dissolve at a rate of 3.25 mg cm -2 h -I. It is possible to cal-
culate from this that spherical gold particles with a diameter of 44 will
take approximately 13 h to dissolve, while 150 particles will take more
than 40 h. It is for this reason that most gold plants include gravity separation
equipment in the milling circuit if there is coarse gold in the deposit.
The mechanism by which silver metal leaches is also electrochemical, with
the same stoichiometry as shown. for gold in eq. (1). The stability of the
Ag(CN)2" complex is considerably less than that of Au(CN)2" , however, so
the reaction is less favourable, requiring higher oxygen and cyanide concen-
trations for comparable leaching rates. Silver is often present in nature as the
sulphide Ag
2
S, and, according to Shoemaker and Dasher [13], reacts with
cyanide as shown in eq. (5):
+S2- (5)
This reaction is slow and reversible, and only proceeds efficiently to the right-
hand side if the cyanide concentration is kept high (usually> 0.04 mol dm -3)
and the sulphide concentration low (usually by simultaneous oxidation to
thiosulphate ):
2S
2
-+20
2
+H
2
0-.S
2
0j- +20H- (6)
Chemistry of cementation
The zinc-cementation process for the recovery of gold and silver was intro-
duced in 1890, at the same time as the cyanide leaching process, and consisted
essentially of passing the gold cyanide pregnant solution through a bed of zinc
shavings. The process was fairly crude and inefficient, and during the follow-
HYDROMETALLVRGY OF PRECIOUS METALS RECOVERY 135
ing 30 years three major modifications were introduced that improved the
efficiency of the process. These innovations involved the addition of lead salts
to the pregnant solution (1894), the use of line dust in place of line shavings
(1897), and deaeration of the pregnant solution prior to cementation (1916).
The historical review compiled by Leblanc [ 14] on the recovery of gold from
pregnant liquors gives an excellent account of these early developments.
Once the major metallurgical and economic shortcomings of the process
had been resolved, the incentive for further research in this area largely fell
away. Fortunately, the rekindled interest in electrochemistry that began in the
1960s led to research into the mechanism of cementation, and much of the
mystique has been stripped away.
The major reactions are the cathodic deposition of gold and the anodic cor-
rosion of line, which occur at the surface of the zinc particles:
Au(CN)i +e=Au+2CN-
+2e
(7)
(8)
Side reactions that influence the economics of the cementation process are
the reduction of water and dissolved oxygen:
2H
2
0+2e=20H-+H
2
(9)
O
2
+2H
2
0+4e=40H- (10)
both of which involve the simultaneous and wasteful consumption of line, by
oxidation to Zn2+ and formation of zinc tetracyanide, i.e. if there is oxygen
present in solution, the following reaction occurs:
2Zn+8CN- +0
2
+40H- (11)
This reaction is minimized in a modem gold plant by de-aeration of the preg-
nant solution in a Crowe vacuum tower. The consumption of zinc via the
hydrogen evolution reaction (eq. (9 can also be controlled to a certain ex-
tent by increasing the potential of the cementation reaction, by decreasing the
concentration of free cyanide, as discussed by Nicol et al. [8]. In practice,
however, if the cyanide concentration is reduced too much, a passivating layer
of zinc hydroxide may form on the surface of the zinc particles [8]. The del-
eterious effects of excessive and insufficient concentrations of free cyanide
were noted by Leblanc [14], who recommended that the concentration of
free cyanide should be maintained between 2X 10-
3
and 6X 10-
3
mol dm-
3
(50 and 150 ppm CN-).
The role of dissolved lead ions (usually lead nitrate) in solution is also
apparently associated with a shift in the redox potential at the surface of the
linc particles, and Pb2+ concentrations as low as 1-2 ppm have been shown
to shift the potential in a positive direction by as much as 0.2 V [15,16]. The
mechanism associated with this considerable enhancement is not clear, but
136 CA, FLEMING
similar effects have been found with thallium, mercury and bismuth ions [16].
The addition of lead to above 10 ppm may actually retard the cementation of
gold [2,8] so this, too, needs to be carefully controlled.
Finally, the pH of the pregnant solution also influences the redox potential,
which shifts in a negative direction with increasing alkalinity. This can be
beneficial if the rate of the cementation reaction is not entirely under diffu-
sion control, as elucidated by Nicol et al. [8]. Leblanc [ 14] reported that the
cementation of gold was practically constant over the pH range 8-11, but that
a marked improvement could be obtained by an increase in pH to a value
between 11.5 and 11.9.
The zinc metal used for cementation must clearly have a large surface area,
because the overall rate of any electrochemical solid-liquid reaction is di-
rectly proportional to surface area. Zinc dust fulfils this requirement, but is
readily oxidized by atmospheric oxygen during storage. The coating of zinc
oxide must be dissolved before cementation can occur:
(12)
which accounts for the frequent necessity for circulation of the pregnant so-
lution for considerable lengths of time at the start-up of a cementation reac-
tion. High concentrations of free cyanide may be beneficial during this period.
The most common impurities in gold pregnant solutions, including sul-
phite, sulphate, thiosulphate, ferrocyanide, thiocyanate, copper, zinc, nickel
and cobalt, appear to have little or no effect on the cementation process
[ 14,15], although deleterious effects can be observed if the concentrations
are very high. Plaskin et al. [17] reported that copper adversely affected the
cementation of gold, which almost ceased at a copper concentration of 200
mg 1- I. However, the deleterious effect of copper decreases as the concentra-
tion of cyanide is increased, from which it can be inferred that the effect of
copper may be due' to a reduction in the concentration of free cyanide as a
result of the formation of cyanide complexes. Hancock and Thomas [18]
reported that nickel at concentrations higher than 200 mg 1-1 has a slight
retarding effect on cementation.
The only species that have a marked deleterious effect on cementation ap-
pear to be sulphide ions and soluble compounds of arsenic and antimony
[ 14,15,17]. The effects are observed even at very low concentrations (I mg
1-1 and lower). The poisoning effect of sulphide ions is thought to be due to
the precipitation of insoluble zinc sulphide on the surface of the zinc parti-
cles. No satisfactory mechanism has been prpposed for the effect of arsenic
and antimony.
Aluminum is sufficiently electronegative to reduce aurocyanide ions to gold,
and its use as a cementing agent was patented by Moldenhauer in 1893. The
factor mitigating against its use is the dissolution reaction:
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 137
(13 )
which requires the pH value of the solution to be above 12 in order to keep
AI3+ in a soluble form. Furthermore, calcium aluminate has a low solubility,
and the use of sodium hydroxide or sodium carbonate is required to adjust
the pH; the use of lime must be avoided. A historical review of the use of
aluminum for the cementation of gold has been compiled by Nagy et al. [ 19].
NEW PROCESSES IN GOLD AND SILVER RECOVERY
Free-milling ores
Carbon-in-pulp (ClP) and carbon-in-leach (ClL)
The ability of activated carbon to adsorb gold from solutions has appar-
ently been known since 1847 [20], although it was not until 1894 that the use
of carbon as a precipitant for gold in cyanide solution was first proposed and
patented by Johnson [21]. Unfortunately for Johnson, it was to be another
20 years before the process was used commercially for the first time, at the
Yuanmi Mine in Western Australia [22], and a further 60 years before the
technology had really come of age as a viable alternative to the zinc cemen-
tation process.
Progress over this period, particularly the pioneering work performed by
Zadra and others at the United States Bureau of Mines [23,24], has been
reviewed by McDougall and Fleming [25].
The most important development in the widespread commercialisation of
CIP was the introduction of the process at the Homestake Mine in South Da-
kota, in 1973, where about 2200 tons of fine slimes are treated per day [26].
The Homestake Gold Mine always had separate treatment processes for its
sand and slime fractions and when, in 1971, the slime filtration plant needed
to be replaced, a CIP plant was installed on the advice of the US Bureau of
Mines. This highly successful operation transformed the image of the CIP
process from an "experimental" and small-scale process to one that could be
adapted for the treatment of high-tonnage flows.
The process spread rapidly, firstly in South Africa in the late 1970s, and
then throughout the rest of the gold-producing world in the 1980s, and it is
now the preferred process for all new gold plants, probably accounting for
over 50% of world-wide gold production. One of the important contributions
made by the South Africans was their realisation that the process need not be
restricted to the treatment of materials that were difficult to filter, such as the
slimes fraction of an ore, but could be applied to almost any feed, achieving
high overall gold recoveries, with relatively low capital and operating ex-
penses. CIP plants were built in South Africa that treated anything from whole
ores, to flotation concentrates, flotation tailings, roasted calcines, re-pulped
138 C.A. FLEMING
filter plant residues or, most significantly, old residue dumps. Many hundreds
of millions of tons of tailings are scattered throughout the Witwatersrand re-
gion of South Africa and, because of the superior economics of the CIP pro-
cess compared to zinc cementation, it became feasible for the first time to
profitably re-treat this material. Between 40 and 50 million tons of tailings
are now being treated in this way every year, yielding about 20 tons of gold
per annum from those wastes, and exposing, under the old dumps, hundreds
of acres of prime industrial land for development. The economics of the dump
re-treatment process require that very large tonnages be processed, and this
led to the next major contribution made by the South Africans; namely, the
development of the process engineering that was required to take the process
from the 60,000 ton/month scale of the Homestake operation, to the 1.5-2
million ton/month scale of the ERGO operation, for example. One of the
most important developments in this area was the move away from external
interstage screens, as used at Homestake, to internal air-swept screens. In the
external screening concept, all of the forward flow of pulp had to be pumped
from each adsorption tank up onto a vibrating screen, located on a deck above
the tank. The carbon in the screen overflow flowed back into the tank from
which it had been pumped, while the pulp underflow advanced to the next
tank in line. While this approach was viable on the small scale of the Home-
stake plant, it was clearly impractical on the large-scale South African plants.
Following the early development of the EPAC (equalized-pressure, air-
cleaned) screen in South Africa, a number of new and innovative designs,
based on similar principles, appeared on CIP plants around the world, partic-
ularly in Australia. Many of these and other developments have been re-
viewed in papers by Laxen and co-workers [27,28] and Fleming [29], which
also detail the layout of a typical multi-stage, countercurrent CIP adsorption
plant.
The final contributions made in South Africa have been on the metallurgi-
cal processing side. Fundamental work carried out at the Anglo American
Research laboratory (AARL) and at Mintek, resulted, for example, in the
development of a successful alternative to the Zadra method of eluting gold
and silver from carbon [30] and in the design of a new, more efficient type
of electrowinning cell for the recovery of gold from carbon eluates [31 ].
Work carried out at AARL and Mintek also went a long way toward resolv-
ing the mechanisms involved in the adsorption [32-34] and elution [35] of
gold and other competing species on activated carbon, and of the factors af-
fecting adsorption efficiency [36]. Simple mathematical models, based on
the kinetics of gold extraction by carbon, were developed [37], which ad-
vanced the understanding of several aspects of the countercurrent CIP ad-
sorption process. For example, by modelling batch CIP data, it was possible
to design parameters such as the flow rate and concentration of carbon and
the number of adsorption stages required for optimum metallurgical effi-
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 139
ciency in a continuous process. Since then, a more rigorous model has been
developed, which describes the adsorption process more effectively but which
is, by necessity, more complex [38].
The fundamental difference between a CIP flowsheet and the conventional,
zinc-cementation flowsheet lies in the fact that, in order to recover gold by
reduction on zinc dust, it is first necessary to produce a clarified filtrate,
whereas in the CIP process it is possible to extract gold cyanide directly from
the slurry. In this way, the costly and inefficient solid-liquid separation pro-
cesses of a conventional gold plant (filtration or CCD) are replaced by the
relatively simple and inexpensive screening procedures that are used in a CIP
plant to recover the carbon granules from the leach slurry.
Early estimates of the capital and operating costs associated with the CIP
process indicated that savings of 20-50% could be expected, compared to the
conventional process (Table 2), and these estimates have been vindicated,
for the most part, on the great many CIP plants operating around the world
today. This has meant that lower grade gold ores can be economically treated
than would previously have been the case, and this has been a major stimulus
to new gold mine development around the world. Another important benefit
of the CIP process is the generally improved efficiency of gold extraction that
can be achieved. This stems from two factors. Firstly, soluble gold losses from
a conventional filtration or CCD plant usually amount to about 1 % of the
pregnant solution (i.e. 0.03-0.05 mg 1-1), in the case of easily filterable sol-
ids, while even higher soluble losses are suffered in the case of poorly filtering
or settling solids. By contrast, soluble losses ofless than 0.01 ppm can usually
be achieved on a well-managed CIP plant. Secondly, the additional 5 or 6
hours leaching time in the CIP adsorption contactors usually results in extra
gold dissolution. This extra dissolution can be considerable ifthere are "preg-
robbing" constituents in the ore. Evidence for extra dissolution is obtained
by an analysis of the washed solids in the feed and the discharge from a CIP
plant. In most instances, the residue solids are lower in gold than the feed
solids, despite the presence of fine abraded carbon from the CIP tanks in the
residue [27].
The final advantage of the CIP process is that it is far less vulnerable than
TABLE 2
Comparative costs of carbon-in-pulp and conventional (Merrill Crowe) processes
Process
Capital cost Operating cost
Conventional
1.00 1.00
CIP [51)
. 0.68 0.77
CIP [62]
0.75
0.94
CIP [42, p. 382)
0.37-0.58 0.88-1.18
140 C.A. FLEMING
the conventional process to impurities such as sulphide, arsenate and anti-
monate in the leach liquor, and parameters such as cyanide concentration,
pH and oxygen concentration do not have to be carefully controlled, as their
influence on the loading of gold on carbon is fairly minor [39].
The mechanism of adsorption of gold cyanide onto activated carbon has
been studied extensively over the last 10 years, and the results of these studies
have been reviewed recently [34]. Most of the evidence indicates that gold is
reversibly adsorbed on the carbon, as the aurocyanide ion, without undergo-
ing any chemical change. Adsorption occurs on the extended surface of the
porous carbon granules (total surface area -1000 m
2
g-I), and is only lim-
ited by physical access of the aurocyanide ion into the smaller micropores
( < 10-20 A). The adsorption efficiency of different ions and molecules
decreases with increasing hydration and hydrophilicity and, in the case of
inorganic metal complexes, is believed to involve charge transfer from the
graphitic planes at the carbon surface to the metal atom. Therefore, factors
that decrease the hydration and ionicity of the aurocyanide complex, such as
increasing acidity (H++Au(CN)2"-+HAu(CNh) or ionic strength
(Ca2+ + 2Au (CN) 2" -+Ca [Au (CN h h), enhance the adsorption efficiency on
activated carbon. Conversely, an increase in the concentration of hydroxyl or
cyanide ions reduces the adsorption of gold on carbon and this, coupled with
the reduction in adsorption capacity that occurs with increasing temperature,
forms the fundamental basis of both the Zadra and the AARL elution
processes.
In practice, only a small fraction of the total adsorption capacity of carbon
for gold is utilized in normal CIP operation. It has been shown in the labora-
tory, for example, that gold loading values as high as 3 mol kg-I (- 600,000
g t - I ) can be achieved, whereas the loading of gold on carbon in the first stage
of a CIP plant seldom exceeds 0.05 mol kg-I (10,000 g t-
I
). There are sound
practical and economic reasons for this. Firstly, the aurocyanide ion adsorbs
relatively rapidly onto carbon in the initial stages of the extraction reaction,
during which adsorption occurs in the relatively large macro- and meso-pores.
Contact times of weeks or even months are required, however, to fully utilize
the total surface area available for aurocyanide adsorption [36]. Secondly,
carbon breaks down slowly in the adsorption contactors as a result of the ab-
rasive action of the ore in the leach slurry. This carbon is lost from the circuit
once it is reduced to a size smaller than the interstage screens, and this would
result in significant gold losses if the carbon were loaded to excessively high
gold values. This problem is exacerbated by the fact that the rate of the re-
versible desorption reaction is slow and, therefore, a highly-loaded carbon
granule in the first adsorption contactor would only desorb gold slowly if it
fractured and flowed co-currently with the slurry out of the CIP plant [25].
A third reason for keeping gold loadings on the carbon at a relatively low level
relates to lock-up of gold in the CIP plant. Carbon inventories on large CIP
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 141
plants range from 50 to 100 tons or more, and the gold loaded on this carbon
is removed from an income-earning capacity for the lifetime of the CIP plant.
This can have a critical effect on cash-flow during start-up and, ultimately, on
the economic viability of a new operation.
One of the implications of operating CIP plants at gold loadings well below
the loading capacity of the carbon is that plant performance is a function more
of the kinetic than the equilibrium or thermodynamic characteristics of the
carbon. The rate of gold loading is determined by mass transfer of the auro-
cyanide ion both in the pulp and within the pores of the carbon granules [36].
Therefore, factors such as the agitation efficiency in the adsorption contac-
tors, the viscosity and pulp density of the slurry, and the particle-size of the
carbon granules can all have a significant bearing on CIP performance.
The only negative implication of operating CIP plants with fairly low gold
loadings is that carbon flow rates are relatively high. This increases carbon
handling requirements and, moreover, because the capacity of the elution and
reactivation plants are directly proportional to carbon flow rate, capital and
operating costs are affected adversely. Consequently, elution and reactivation
tended to be the most cost-intensive unit operations in the flowsheet, and this
created bottlenecks in the process as engineering companies sought to control
capital costs by specifying minimum plant requirements [27].
Regeneration ofloaded carbon to that offresh activated carbon is generally
a three-step process. In the first step the loaded carbon is treated with a min-
eral acid (usually hydrochloric, although nitric is also used on some plants).
The main purpose of the acid-wash is to strip calcium carbonate from the
carbon, which can build up to levels of 5% or more in the adsorption circuit.
The calcium comes from the lime used to maintain the pH above 10.5, while
the carbonate is believed to derive predominantly from the oxidation of
cyanide:
CN- + 20H- + H
2
0+ 2e
+ HC03"
(14 )
(15 )
This reaction is catalyzed by activated carbon [40], and results in a decrease
in cyanide concentration in the CIP adsorption tanks of about 50% of the feed
value. The acid-wash also removes certain base metals from the carbon, as
well as fine slimes of silica, clay, and iron oxides, and results in more efficient
precious metal elution in the second step in the regeneration process.
As the adsorption of gold on activated carbon is a thermodynamically re-
versible process, chemical and physical factors that inhibit adsorption will
enhance elution, and vice versa. A number of these factors are utilized in the
elution of carbon on all CIP plants. The most important parameter is temper-
ature, since both the kinetics (activation energy 66.5 kJ mol-I) and the ther-
142 C.A. FLEMING
modynamics (exothermic heat of reaction""' 40 kJ mol- I ) of elution improve
with increasing temperature. The modem trend is towards temperatures higher
than 100C, with the use of pressurized equipment. A second important pa-
rameter is ionic strength. An increase in ionic strength generally favours ad-
sorption, and the use of deionized water to achieve enhanced rates of gold
and silver elution in the AARL process is a manifestation of this phenome-
non. However, the rate of elution also increases with increasing concentra-
tions of solvating anions such as cyanide or hydroxide. Therefore, because of
the opposing influences of the anion effect and the overall ionic strength ef-
fect, the rate of elution passes through a maximum as the rate of the eluting
salt is increased [35]. Another important factor that can be utilized to en-
hance the rate of elution is the effect of polar solvents, such as acetonitrile,
acetone, methanol and ethanol. This effect is attributed to an increase in the
activity of the cyanide ion and a decrease in the activity of the aurocyanide
ion in polar solvents, as compared with water [41 ], and results in significant
enhancement of the rate of elution.
Gold is usually recovered from elP eluates by electrolysis, but in some cases
zinc cementation is still preferred. The concentration of gold and silver in a
elP eluate is usually 2 to 3 orders of magnitude higher than in the original
pregnant leach liquor and, owing to the selectivity of the elP process, the
ratio of precious metals to contaminants is also much higher. Consequently,
the final gold recovery process is less critical to the overall efficiency and eco-
nomics of the process than is the case in the conventional cyanidation/ce-
mentation process, and can be "over-designed" with little impact on capital
costs. This part of the process has therefore caused few problems, and has not
attracted much attention in the way of new developments. Electrowinning
and cementation of gold from elP eluates, and the characteristics of the dif-
ferent cells that are used commercially, have been reviewed recently by Bailey
[42, pp. 550-570] and Nicol et al. [8, pp. 890-894].
The third step in the regeneration process involves the thermal reactivation
of the carbon, to either carbonize or volatilize any organic compounds that
adsorb onto the carbon from the cyanide pulp in the elP contactors. The
carbon is heated to 650-750
o
e for up to 30 min, in the absence of air, and is
then quenched in water and screened (to remove fines) before recycling to
the last adsorption contactor in the elP plant.
There are two other processes involving the extraction of gold and silver
from cyanide solution with activated carbon. The first is the carbon-in-Ieach
(elL) process. For most gold ores, the time required for efficient gold leach-
ing is between 24 and 36 h, whereas the time required for efficient extraction
of the gold by carbon is usually only 5-10 h. In elP, the leaching is essentially
complete prior to carbon adsorption, and the elP adsorption section is about
5 times smaller than the leaching plant. In principle, the elP adsorption sec-
tion could be dispensed with and the carbon added to the leaching tanks rather
than to small adsorption tanks at the end of the leaching section, and this is
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 143
the basis of the elL process. In choosing between elP and elL for a particular
application, a number of factors must be considered:
( 1 ) The major advantage of elL over elP is the lower capital cost of the
plant, which results from the removal oHive or six elP adsorption tanks and
ancillary equipment from the flowsheet. The operating costs associated with
the elP section would also be eliminated.
(2) Because the elL leaching tanks are much bigger than elP tanks, elL
has a further advantage over elP in that the surface area available for inter-
stage screening is much greater, and screening is less likely to be a bottleneck
in the process.
(3) A disadvantage of elL is that the carbon concentration in the pulp is
lower than in elP for the same carbon inventory. Hence, more pulp is trans-
ferred upstream during countercurrent transfer of the carbon, which places
an increased load on the pumps or airlifts and on the interstage screens.
(4) The metallurgical efficiency is lower in elL than it is in elP. This is
because leaching is incomplete when the pulp encounters carbon in the first
adsorption stage of elL. Consequently, the concentration of gold in solution
is lower and, to compensate for a decrease in the kinetic driving force, more
carbon is needed to match the metallurgical performance achieved by elP.
This has been demonstrated by mathematical modelling [37, Part II] and,
even for an ore from which the gold is leached at a relatively fast rate, the
carbon inventory in elL should be about 25% higher than in elP, for the
same metallurgical performance.
(5) Most elP plants have performed well below design specification on
start-up, often because the pulp contains poisons that were not exposed dur-
ing piloting. The adverse effect of poisons can be countered by the use of a
larger carbon inventory or a larger number of adsorption stages. Because a
CIP plant has far more flexibility than a CIL plant, it lends itself better to the
introduction of these types of changes at some stage after the initial commis-
sioning of the plant.
The third process utilizing activated carbon involves extraction of gold and
silver from solutions rather than slurries.
The major cost advantage of processes using activated carbon over zinc
cementation lies in the ability of granular carbon to extract gold cyanide di-
rectly from pulp, thus dispensing with the costly countercurrent decantation
or filtration methods of solid-liquid separation. However, there are also in-
stances where the recovery of gold from clarified or unclarified solutions by
activated carbon is preferred to zinc cementation, and a number of plants
around the world recover gold in this way.
The choice between carbon columns and zinc cementation is based on anal-
yses of capital and operating costs and consideration of the metallurgical ef-
ficiency. As a broad generalization, treatment in carbon columns is more eco-
nomical for large volumes of low-grade solutions containing mainly gold,
144 CA. flEMING
whereas zinc cementation is preferred for relatively small volumes of high-
grade solutions, particularly those rich in silver.
Obvious examples of large-volume, low-grade solutions are tailings ponds
from conventional eeD or filtration plants, and there are a number of plants
around the world recovering gold from sources such as this, in packed carbon
columns. The pay-back period for this sort of scavenging operation is usually
no more than a few weeks!
Another example of a situation in which carbon is usually preferred to zinc
is in the treatment of solutions from heap-leaching operations. Here too, the
main advantage of carbon over zinc cementation is that the pregnant solution
need not be clarified or de-aerated prior to feeding to a carbon column.
The downstream processing to strip gold from the carbon and reactivate
the carbon is the same as in the elP and elL processes, although it may not
be necessary to reactivate the carbon as frequently because the concentration
of organics in solution is usually less than in pulps.
Resin-in-pulp
The resin-in-pulp (RIP) process for precious metal recovery is in its in-
fancy in the Western World, although it has apparently been practised widely
in the USSR for many years [43], and was used in the uranium industry in
the USA, in the 1960s and 1970s.
A detailed description of the mechanisms involved in the adsorption and
elution of aurocyanide on ion-exchange resins is beyond the scope of this pa-
per, but the chemistry involved has been covered in a recent review [44].
The history of anion-exchange resins in gold processing, which has been
detailed in a recent review [8], is not as long as that of activated carbon. The
earliest detailed studies were carried out in the UK in the early 1950s and,
since then, there have been sporadic bursts of activity in a number of labora-
tories around the world. The successful commercialization of the eIP process
has undoubtedly contributed to the current surge of interest in the RIP pro-
cess, as it has served to demonstrate the very substantial benefits of "in pulp"
processing. Moreover, a number of the innovations that have taken place in
the engineering oflarge-scale eIP processes will probably benefit the RIP pro-
cess, as the broad engineering features of the two processes are likely to be
similar.
Up to now the RIP process has not received serious consideration by min-
ing or engineering companies during flowsheet evaluation for greenfield op-
erations. However, resins have several characteristics that are quite different
from carbon, which make them potentially more versatile substrates for the
recovery of gold. The following are examples:
(I) Resins are superior to all types of granular activated carbon with re-
spect to the rate and the equilibrium loading of aurocyanide. This means that
the resin inventory will be lower and the size of the elution plant will be smaller
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 145
in RIP than in CIP, and this impacts favourably on both the capital and the
operating costs of the process.
(2) Resins are eluted at ambient pressure and at temperatures of no more
than 60C, whereas carbon elution has to be carried out at much higher tem-
peratures, preferably around 120-130C, in a pressure vessel. This, too, re-
sults in lower capital and operating costs for RIP than CIP.
(3) Activated carbon requires regular thermal reactivation to remove ad-
sorbed organic material-a step that is unnecessary with resins. The cost of
the equipment required for carbon reactivation can be a significant propor-
tion of the total capital cost of a CIP plant, particularly for smaller operations.
Resin elution/regeneration costs, on the other hand, are fairly insensitive to
the scale of the operation (because the major costs are direct costs, such as
the consumption of chemicals) and this indicates that RIP may find its initial
market niche in smaller operations.
(4) Another potential capital cost advantage of resins over carbons stems
from the fact that RIP adsorption tanks operate efficiently at high concentra-
tions of resin in the pulp (20-30% by volume), with no loss in interstage
screening duty, and with no increase in resin breakage. CIP plants, on the
other hand, generally operate with carbon concentrations of no more than 3-
6% by volume. Therefore the adsorption tanks in RIP can be at least 5 times
smaller than in CIP, with no loss in metallurgical efficiency, and this contrib-
utes to lower fixed costs.
( 5) Resins do not appear to be poisoned by organic species such as flota-
tion reagents, machine oils and lubricants, or solvents, all of which can se-
verely inhibit the loading of gold on activated carbon. Similarly, species such
as hematite and shales or clay-type minerals depress the loading of gold on
carbon, but have little effect on resin. Higher gold loadings should be achiev-
able on resins than on carbon in situations where any of these species are
present in the pulp, and this again reduces the size of the resin elution plant.
( 6) Resins can be used to co-extract other metal cyanide complexes effi-
ciently from the gold leach solution, such as those of cobalt, copper, nickel,
zinc and iron. Therefore, the technology can be adapted to produce environ-
mentally acceptable tailings, whilst at the same time presenting an opportu-
nity for recycling excess cyanide to the leaching circuit. This capability is being
researched at several laboratories and, if successful, could open the way to
treat a variety of deposits that presently cannot be economically processed,
because of high cyanide consumption.
(7) Another important consideration in favour of resins is that much can
be done during the synthesis of resins to tailor-make them for specific appli-
cations, and there has been notable success recently, at Mintek in South Af-
rica, in the development of gold-selective resins [45,46].
The many advantages of resins have been amply vindicated in the only op-
erating RIP plant in the western world, at the Golden Jubilee Mine in South
146 C.A. FLEMING
Africa. The conversion of that plant from carbon-in-Ieach to resin-in-pulp
transformed the operation from near bankruptcy in 1987 to an extremely
profitable operation today [47,48].
Since converting to RIP, the mill capacity at Golden Jubilee has been dou-
bled, and overall gold recovery has improved from 65% to 85%, resulting in
an increase in gold production of more than 200%. What is important from
an economic point of view, is that this has been achieved in RIP adsorption
tanks that are one-sixth the size of the old CIL tanks, and with a resin inven-
tory less than half the old carbon inventory. Operating costs in elution and
regeneration have been contained by loading the resin to gold concentrations
of 4000-6000 g t- I, from a feed solution value of less than 1 ppm. By con-
trast, gold loading on the carbon seldom exceeded 1500 g t-
I
, and was usually
less than 1000 g t - I. Resin losses due to breakage, at about 8-10 g per ton of
ore treated, are also considerably lower than carbon make-up in the old CIL
plant (50-100gt-
I
).
The rather spectacular improvement in performance obtained with resins
in this particular application is, to a large extent, a consequence of the fairly
unique feed material being treated, which is heavily contaminated with nat-
ural organics (humates and fulvates) and is very viscous owing to its high
clay content. There may not be many other deposits that will respond as fa-
vourably to RIP relative to CIP, as has the Golden Jubilee deposit, although
it has been shown that the efficiency of gold extraction with resins is generally
similar to or slightly superior to that of carbon [39]. Moreover, it would seem
that, even in those cases where there are no clear-cut advantages for one pro-
cess over the other as far as extraction efficiency is concerned, RIP processing
still deserves serious consideration during flowsheet evaluation, because of
the potential for lower capital and operating costs.
The potential disadvantages associated with the use of resins for gold ex-
traction from pulps are mostly concerned with the engineering aspects of the
process. Examples include:
(1) Resin beads are smaller than carbon granules and, consequently, both
the pre-screening and the interstage screening duties must be performed on
finer screens in RIP than in CIP. Since this is already a bottleneck area on
many CIP plants, it could present a major engineering challenge on large RIP
plants.
(2) The physical strength of resins and their resistance to attrition in the
adsorption tanks has been an unknown factor until relatively recently. Since
resins are considerably more expensive than activated carbon, high resin losses
in RIP would place an intolerable cost on the process. The experience at
Golden Jubilee suggests that this is unlikely to be a serious problem on a well-
designed and managed RIP plant.
(3) Resins are less dense than activated carbon and tend to accumulate
near the surface of the pulp unless it is efficiently agitated. The best methods
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 147
of mixing and pumping resin/pulp slurries without causing excessive resin
breakage need to be established.
( 4) Currently available, commercial, anion-exchange resins are less selec-
tive than activated carbon for aurocyanide over the base-metal cyanides that
are prevalent in cyanide leach liquors. These complexes can load as strongly
as the aurocyanide ion onto anion-exchange resins, and reduce the capacity
of the resins for gold. This disadvantage is, of course, very site-specific, but
there are many examples of cyanide leach liquors that contain high concen-
trations ( > 100 ppm) of copper or zinc, and resins will not compare favour-
ably with carbon in these situations. The treatment of solutions and pulps
with these compositions will have to await the commercialization of gold-
selective resins.
Heap leaching
Although the heap leaching process has been used for many years for the
recovery of copper and uranium from oxidized ores, the application of the
technique to low-grade gold and silver ores is a relatively recent innovation
that was developed by researchers at the United States Bureau of Mines in
the late 1960s [49,50]. By opening up vast reserves of low-grade ores to eco-
nomic exploitation, heap leaching has probably contributed more than any-
thing, apart perhaps from the CIP process, to the surge in gold production
since the early 1980s. The process is very low cost [51] and flexible, being
well-suited both to small (5-10 t d-
I
) and large-scale operation (10,000 t
d - I ). Depending on the permeability of the rock being treated, it can be pro-
cessed with or without crushing. If there is a large amount of clay in the ore,
the fines are usually separated and agglomerated before being piled on the
heap. The host rock must be porous and permeable or of small grain size, and
the gold and silver must be liberated and non-refractory. Refractory ores and
ores containing species that consume excess cyanide (oxidized sulphides of
As, Sb, Zn, Fe and Cu, for example) or oxygen (pyrrhotite, organics, etc.) are
not well-suited to heap leaching, because of difficulties in controlling the
leaching chemistry within the heap.
In essence, the process consists of piling ore to a given height (usually 3-10
m) on a sloping impermeable bed, and spraying dilute cyanide solution onto
the top of the heap. After flowing down through the heap, the leach solution
is directed along the sloping impervious floor ofthe heap to a collection pond.
The pregnant solution, which usually contains 1-3 ppm gold, is pumped
through carbon columns to recover the gold, and flows into a second pond
where the cyanide strength is restored prior to recycling to the top of the heap.
On some heap leach operations, gold is recovered by zinc cementation but, as
pointed out previously, carbon adsorption is usually more effective for the
treatment oflarge volumes oflow-grade solution.
Recycling of the leach solution to the heap is continued until the gold con-
148 C.A. flEMING
centration in the solution from the bottom of the heap drops to an economic
cut-off level, usually about 0.05-0.1 ppm. The kinetics of leaching are ob-
viously slower than in a conventional leaching plant, firstly because of the
generally larger ore particle-size, and secondly because of the slow rates of
percolation of solution and diffusion of cyanide and oxygen. The leaching
cycle for a heap is therefore usually several weeks, and can stretch to months,
and overall gold recovery can be 20-30% lower than in a milling/agitation
leach process.
Once leaching is complete, the barren heap is either discarded to a tailings
disposal area if the same leach pad is to be used over and over again, or it is
left in place, with a new pad being constructed for the next batch of ore. In
certain instances, where the physical characteristics of the ore warrant it, heaps
can be built one on top of another, as often practised in copper heap-leach
operations. Closure of heap-leach operations is becoming an important envi-
ronmental issue, and it is vital to ensure that runoff from abandoned heaps is
isolated from surface and ground water sources, at least until seepage of cya-
nide from the heap ceases.
Refractory gold and silver ores
General
By definition, a refractory ore is one that does not respond well to direct
cyanidation. The actual leach efficiency required to meet this criterion is
somewhat arbitrary, although most people in the industry are using a level of
80% recovery, or less, to categorize a refractory ore. Whatever the definition,
refractory ores have one thing in common: in order to achieve a reasonable
return on investment, it is generally necessary to pretreat the ore in some way
to improve gold recovery. The reasons for refractoriness and the different
methods of pretreating a refractory ore will be examined and evaluated in this
section. Final gold recovery from the pretreated ore is by cyanidation fol-
lowed by CIP, CIL or zinc cementation and, in this regard, the flowsheet is no
different from one that would be used for a free-milling ore, so these aspects
will not be discussed.
The recovery of precious metals from refractory ores has received much
attention in the last 10 years as more and more ore bodies are found that do
not respond to direct cyanidation. This trend is likely to continue as the oxi-
dized free-milling gold reserves close to the earth's surface become mined-out
and depleted. Underlying many of these oxidized ore bodies are sulphidic
zones that have, until recently, been left in tlie ground owing to their poor
response to direct cyanidation.
In response to this trend towards increased refractoriness with existing ore
bodies and most new discoveries, there has been a surge in activity amongst
research organizations, mining, engineering and consulting companies, to-
HYDROMETAllURGY OF PRECIOUS METALS RECOVERY 149
wards the development of technically and economically feasible pretreatment
routes for these ores. These efforts have resulted in the development of sev-
eral new processing options and the number of plants that have been built to
treat refractory ores is increasing rapidly.
There is an increasing awareness today of the great variability between re-
fractory ores, stemming from differences in the mineralogical, metallurgical
and chemical properties of these ores. One consequence of this is that it is
unlikely that anyone pretreatment route will emerge as a universal process
for all refractory ores. To appreciate the validity of this comment, it is neces-
sary to consider the mineralogical characteristics of a refractory ore. In very
general terms, refractoriness is caused by one or both of the following:
( 1 ) Physical encapsulation of very finely disseminated gold particles (usu-
ally less than a few micrometres in size) within a mineral that is unreactive
and impervious in cyanide solution.
(2) Chemical interference by one or more constituents of the ore with the
cyanide leaching process.
In the mechanism of physical encapsulation, the dominant mineral is arse-
nopyrite, which exhibits a strong association with finely disseminated gold in
many ore bodies. Other sulphide minerals that are known to trap gold are
pyrite, pyrrhotite, realgar and orpiment. In order to liberate the gold from this
type of refractory ore, it is generally necessary to chemically oxidize the sul-
phides. Until 10 years ago this was done by roasting in air, which resulted in
the oxidation of the sulphides to sulphur dioxide. More recently, pressure
leaching has been developed to an industrial process and, in most instances,
is now preferred to roasting. In this case, sulphide is oxidized to sulphate,
most of which reports to the aqueous phase as sulphuric acid. Bacterial leach-
ing has been developed to an industrial process on a small scale, and also
involves oxidation of sulphides to sulphuric acid in solution. These three pro-
cesses are discussed in more detail in the following sections. Oxidation of
sulphides with nitric acid has been extensively tested and has a number of
attractive features but, apart from one refractory silver ore operation [52]
has not been developed to an industrial scale. In addition to these chemical
oxidation methods, it is also possible to liberate gold trapped in sulphides
physically by grinding the mineral to extremely fine (less than 10 Jlm) parti-
cle size. Despite the relatively high energy consumption, this method is con-
siderably cheaper than any of the chemical oxidation methods. The process is
being extensively tested and marketed in South Africa and Australia, but will
always be limited in its applicability to refractory ores containing gold parti-
cles that are bigger than a few microns. Unfortunately, these seem to be fairly
few and far between.
There are two important mechanisms in which gold recovery is limited
through chemical interference from the ore body. The first results from the
presence of species that consume cyanide or oxygen-the two essential ingre-
150 C.A. FLEMING
dients for effective gold dissolution. In this case, the simplest expedient is to
increase the concentrations of cyanide and oxygen until these species have
completely reacted, but this is often not economically viable. The main cul-
prits are pyrrhotite and copper minerals such as malachite, azurite, chalcocite
and, to a lesser extent, chalcopyrite. In fact, any base metal, such as copper,
nickel, zinc or cobalt, will react strongly with cyanide if it is present in an
oxidized ore body. The second mechanism is associated with the presence of
graphitic, carbonaceous matter in the ore. This is a major component of many
refractory ores, particularly in the State of Nevada, USA. Carbonaceous ores
reduce gold extraction by virtue of their ability to re-adsorb gold once it has
been leached as gold cyanide, a phenomenon that is known as preg-robbing
in the industry. The initial metallurgical solution to this problem was the use
of chlorine to deactivate the carbonaceous matter under strongly oxidizing
conditions [53], and this approach is still followed on certain feed materials
at the Carlin and Jerritt Canyon Mines in Nevada. However, for many ores
the consumption and cost of chlorine are prohibitive and alternative process-
ing strategies have been sought. Blinding and deactivation of the carbona-
ceous minerals with an organic compound such as kerosene is a cheaper alter-
native to chlorine and has been tested with many ores. It is generally not an
effective approach, however. Pressure oxidation is not an obvious choice in
these situations as the carbonaceous material has the potential to become even
more active as an adsorbent for gold cyanide after this treatment. The effect
of preg-robbing can be partially overcome by the use of CIL rather than CIP,
provided the activated carbon used in CIL is more "active" than the carbon-
aceous matter in the ore. The rate of equilibration of gold between the car-
bonaceous matter and the activated carbon can be very slow, however, and
this approach is seldom completely satisfactory. Obviously, the complete
elimination of graphitic carbon is the most desirable solution. Attempts to
achieve this by mineral dressing techniques (gravity, flotation, etc.) have only
ever been partially successful. In many instances roasting of the ore will achieve
this objective, but the operating parameters have to be carefully chosen, even
in this case, to ensure the graphitic carbon is burnt off and is not converted to
a more active form.
Apart from the selection of the best processing route to adopt, the other
decision that the developer of a refractory gold operation needs to make is
whether or not to preconcentrate the gold-bearing ore. This is usually by flo-
tation of either arsenopyrite, pyrite, or both, and gravity concentration is also
possible, although the opportunities in this area are more limited. Preconcen-
tration should meet two objectives for it to b6 economically justifiable: the
gold recovery to the concentrate should be sufficiently high (probably> 90%)
so that the tailings can be discarded without further treatment, and the mass
ratio of concentrate to tailings should be low (less than 1: 10). The other im-
portant factor to be aware of is that the major capital and operating costs of
HYDROMETALLVRGY OF PRECIOUS METALS RECOVERY 151
pretreatment, whether by roasting, pressure leaching or bacterial leaching, stem
from the amount of sulphide sulphur that must be oxidized, not the concen-
tration of sulphide. Hence, the pretreatment costs will be similar for ore and
concentrate, and the major benefits of preconcentration only occur further
downstream as a result of smaller cyanidation and CIP plants.
The final point of general interest in relation to the treatment of refractory
gold ores relates to the choice of the most appropriate lixiviant to leach the
precious metals from the pretreated ore or concentrate. Cyanide has been used
almost universally up to now, and will continue to be the preferred choice in
most cases. However, there has been a great deal of research into alternative
leaching chemistries in the last 10 years. The incentive for this work has
stemmed largely from concerns relating to the ability of the industry to meet
more and more stringent environmental regulations. The validity of this fo-
cus is questionable. Firstly, there is every indication that the industry will be
able to meet the most stringent limits with technology that is currently avail-
able for treating cyanide effiuents. Secondly, it is by no means proven or es-
tablished that the alternative leach liquors will be any easier or cheaper to
treat and render environmentally benign.
However, the treatment of refractory ores is a potential niche application
where alternative lixiviants could become important in the future. The reason
for this lies in the fact that preoxidation of sulphide ores, whether by pressure
or bacterial leaching, involves the generation of sulphuric acid and ferric ions
in solution. In an alkaline cyanide leaching process, these ions must always
be neutralized, and the acidic liquors often need to be separated and washed
from the solids (to reduce cyanide consumption). This would not be neces-
sary when leaching gold with either thiourea or thiocyanate, however, as both
sulphuric acid and ferric ions are, in fact, important ingredients in the overall
leaching reactions in these two systems. This particular application therefore
warrants investigation.
Roasting
By far the most commonly used technique for the recovery of gold from
refractory ores is roasting of flotation concentrates prior to cyanidation. When
arsenopyrite is present, a two-stage process is usually applied-a non-oxidiz-
ing first stage roast at 400-450C to remove the arsenic as volatile arsenic
trioxide, followed by an oxidizing roast at 650-750C, to produce hematite
and sulphur dioxide. Roasting results in a reasonably good recovery of gold
from refractory materials, provided close control is maintained over the
roasting atmosphere and temperature, since the calcine produced must be
sufficiently porous for even the finest gold particles to be accessible to the
cyanide leach solution.
In many instances the presence of carbon causes problems in roasting, since
the high temperature required to burn off the carbon yields a calcine that is
152 CA. FLEMING
not sufficiently porous to expose all of the gold. Nevertheless, roasting should
be the best way of dealing with carbonaceous refractory ores as it is the only
method of potentially eliminating all of the carbon. The conditions have to
be carefully controlled, however, since any carbon that is not burnt off will be
even more active in the calcine than it was in the original ore. The presence
of antimony in a roaster feed also has a deleterious effect on subsequent gold
recovery, and this is probably associated with the formation ofa low melting-
point, impervious, glassy phase that traps gold.
Roasting has developed from fixed-bed hearth and rotary furnaces to fluid-
ized-bed type furnaces. The development of fluid-bed roasting has resulted in
significant improvements to the process, as a result of more homogeneous
reaction conditions and better control of the temperature, atmosphere and
residence time of reactants in the roaster.
More recently, there has been a trend towards the roasting of whole ores
rather than concentrates-a development that has been accelerated by two
important technical innovations, the circulating fluid-bed roaster [54] and
the two-stage oxygen roasting process ofIndependence Mining Company [55].
The main technical challenges in roasting lie in meeting the stringent envi-
ronmental limits for gaseous discharge, particularly in the treatment of ar-
senic sulphide ores. It is generally believed to be necessary to volatilize the
arsenic in an ore in order to achieve high gold recovery in the subsequent
cyanidation process, although unconfirmed reports indicate that this may not
necessarily be true in the oxygen-whole-ore roasting process. Arsenic is vola-
tilized as arsenic trioxide, and must then be very efficiently condensed and
filtered from the off-gas. This dust is too impure for direct sale to the chemical
or pharmaceutical industries, and the cost of refining cannot be justified in
most cases. The material is highly toxic, so storage is also an expensive option.
All sulphides are oxidized to sulphur dioxide during roasting. The produc-
tion of sulphuric acid from the off-gas is technically feasible if the sulphide
grade in the roaster feed is high, but the capital cost of an acid plant is signif-
icant, and cannot be justified unless there is a market for the acid nearby.
Therefore, the sulphur dioxide is generally neutralized by scrubbing the off-
gas in lime, and this can be a significant cost factor, particularly in the roast-
ing of whole ore with a high (> 3%) sulphur content. Lime or limestone can
be pre-mixed with the ore prior to roasting, and this traps some of the sulphur
dioxide in the roaster bed, reducing the demands on the off-gas scrubbers.
This option needs to be carefully tested and optimized, however, as there are
indications that gold recovery may be adversely affected. The practice of dis-
charging untreated gas to the atmosphere is no longer an option on a new
plant, anywhere in the world.
Another parameter that needs to be investigated for each ore or concentrate
type is the degree of sulphide oxidation that is required. Some ores appear to
require as little as 40-50% oxidation to achieve optimum gold recovery. In-
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 153
creasing the degree of oxidation may lead to only marginal improvement in
gold recovery, but produces more sulphur dioxide to be neutralized. The var-
ious other sulphide oxidation processes discussed below are no different in
this regard.
Pressure leaching
There has been a shift in direction in recent years in the treatment of re-
fractory gold ores or concentrates from the traditional roasting method to
pressure leaching. This is the result of two factors: firstly, the inability of first-
generation roasting plants to meet stringent new environmental limits on the
discharge of sulphur dioxide and arsenic trioxide to the atmosphere; and sec-
ondly, the fact that gold recovery from roasted calcines is generally inferior to
that achievable after pressure leaching. This latter factor is especially impor-
tant in the treatment of arsenopyritic ores, where the gold is frequently so
finely disseminated that it is not rendered accessible to cyanide, even though
a porous calcine is produced. In pressure leaching, the decomposition prod-
ucts of the oxidation of both pyrite and arsenopyrite are dissolved species
(which then reprecipitate under certain conditions), so any gold occluded in
these mineral phases, no matter how fine, is fully exposed during oxidation.
The pyrite and arsenopyrite decomposition reactions can be represented as
follows:
2FeS
2
+ 15/20
2
+H20-+Fe2(S04h +H
2
S0
4
2 FeAsS + 70
2
+H
2
S0
4
+2H20-+Fe2(S04h +2H3As04
(16)
(17)
Under the conditions prevailing in an autoclave, 60-95% of the iron repre-
cipitates as a basic ferric sulphate (Fe(OH)S04) or as a jarosite,
(MFe3(OH)6(S04b where M=H+, Ag+, 1/2 Pb2+) while a significant
Portion (80-98%) of the arsenic reprecipitates as the very stable ferric arse-
nate complex (FeAs0
4
).
A number of large pressure oxidation plants have been successfully com-
missioned in the last 6 years (Table 3), and these have demonstrated to the
metallurgical industry that the process is rugged and highly efficient for the
pretreatment of refractory gold ores. There are also several plants under con-
struction, and many more in various stages of metallurgical testing, engineer-
ing design or feasibility assessment. Pressure leaching is undoubtedly the
preferred pretreatment route for non-carbonaceous, refractory gold ores at
the present time.
The first step in a pressure leaching operation will generally involve precon-
ditioning with sulphuric acid to neutralize the carbonates (calcium or mag-
nesium) that are often a constituent of refractory gold ores. If this is not done
prior to pressure leaching, the acid generated in the autoclave, from the sul-
phide oxidation reaction, will decompose the carbonates and produce carbon
154 C.A. FLEMING
TABLE 3
Pressure leaching operations for the treatment of refractory gold ores
Project Status Plant capacity Country
(t d-
'
)
McLaughin Operational 3,000 ore USA
Goldstrike phase I Operational 1,500 ore USA
Getchell Operational 3,000 ore USA
Mercur Operational 800 ore USA
Sao Bento Operational 240conc. Brazil
Campbell Red Lake Operational 70conc. Canada
Porgera Operational 3,500conc. Papua New Guinea
Goldstrike phase 2 Construction 10,000 ore USA
Con Construction 100conc. Canada
dioxide in the autoclave. As a result, the partial pressure of oxygen in the
autoclave is reduced and this, in turn, lowers the rate of sulphide oxidation.
Depending on the relative proportions of sulphide and carbonate in the ore
or concentrate, the overall acid balance may be either positive or negative. If
it is positive, the acid generated in the autoclave can be recycled to the pre-
conditioner by separating the solids and the liquids in the autoclave dis-
charge. If it is negative, it will be necessary to import acid to decompose the
carbonates. This is obviously undesirable, and would be an incentive to treat
concentrate rather than ore, since carbonates can usually be rejected quite
effectively to the tailings during flotation. An alternative approach in a situ-
ation where the sulphide grade is low and the carbonate grade is high, is to
operate the pressure leaching reaction under alkaline conditions, and this is
the processing route that has been followed at the Mercur Mine in the USA.
The rate of sulphide oxidation is slower under alkaline conditions than it is
in acidic solution, and therefore it is necessary to employ longer residence
times and higher temperatures to achieve equivalent gold liberation.
Before pumping slurry to the autoclave, the solids density must be adjusted
to an optimum value, which is a function of the sulphide concentration in the
solids. This can be done before, during or after acid-preconditioning, and could
involve either thickening or dilution of the slurry. The relationship between
solids density and sulphide concentration is extremely important, since it in-
fluences both the heat balance in the autoclave and the ultimate concentra-
tion of sulphuric acid in the autoclave solution.
The oxidation of sulphide with oxygen in the autoclave is an exothermic
reaction, and the concentration of sulphide required for an autothermal re-
action can be calculated from the heats of reaction for the different sulphide
minerals. For example, Mason [56] has calculated that a sulphide grade of
5.6% would be needed for autothermal reaction with an arsenopyritic ore, or
HYDROMETAlLURGY OF PRECIOUS METALS RECOVERY 155
12.9% with a pyritic ore, assuming a solids density of 40% and an operating
temperature of 220C. Mixed pyrite/arsenopyrite ores would possess heat-
generating capacities between these values. Clearly, therefore, if the concen-
tration of sulphide in the feed is higher than these values, some degree of
cooling or dilution may be necessary in the autoclave, to maintain tempera-
ture. If the grades are lower, which is usually the case in the treatment of
whole ore, it is necessary either to preheat the feed slurry prior to pumping
into the autoclave, or to inject steam under pressure into the autoclave. Pre-
heating of ore is achieved by recovering heat from the autoclave discharge
and recycling it to the feed, via a series of heat exchangers operating counter-
current to the flow of slurry.
The importance of sulphide grade in relation to sulphuric acid concentra-
tion stems from the fact that the oxidation of elemental sulphur (an inter-
mediate product in the overall sulphide oxidation reaction) is retarded at
higher sulphuric acid concentrations [57]. This results in a build-up of ele-
mental sulphur in the autoclave product, which can cause both physical (ag-
glomeration) and chemical problems (high cyanide consumption). Interest-
ingly, the rate of oxidation of sulphide (to elemental sulphur) apparently
increases with acidity at low sulphuric acid concentrations, and is then rela-
tively constant at high acidities [57], so gold liberation is not a problem un-
less it becomes re-encapsulated in elemental sulphur. In practice, sulphuric
acid concentrations are usually maintained in the 20-60 g 1- 1 range, and it is
only at concentrations of 100 g 1- 1 or more that elemental sulphur formation
can become severe. It is not always possible to predict the sulphuric acid con-
centration directly from the solids density and the sulphide concentration in
the feed, because even if all of the sulphide is oxidized to sulphate, it is usually
distributed between sulphuric acid in solution and insoluble sulphate com-
plexes in the solids (jarosites, for example), with the distribution influenced
by reaction conditions in the autoclave. Laboratory testing is therefore nec-
essary to determine optimum reaction conditions, although an empirical re-
lationship has been developed by Conway and Gale [58]:
solids density = (0.3[S2-] +0.825)-1 (18)
which can reduce the amount of testing required.
The important operating parameters to control in the autoclave are tem-
perature, oxygen partial-pressure, redox potential, residence time and, as
mentioned earlier, sulphuric acid concentration. Temperature, oxygen par-
tial-pressure and acid concentration all influence the rate of sulphide oxida-
tion, whereas redox potential is used to monitor the extent of oxidation. The
optimum conditions required for a particular refractory ore are established
in laboratory batch and continuous autoclave tests, and the results obtained
on a small scale in the laboratory generally reproduce extremely well in a full-
scale operation. The size of the autoclave is a function of the residence time
156 C.A. FLEMING
required for sulphide oxidation and, since this can dramatically influence the
total capital cost of the facility, and ultimately the economic feasibility of the
process, it is important to maximize the rate of oxidation. The temperature
in the autoclave ranges from a low of 180C at the McLaughin plant to 225C
at the Goldstrike plant, and oxygen partial pressures range from 200 to 550
kPa. Under these conditions, residence times in the autoclave range from 60
to 120 min for pyritic and arsenopyritic refractory ores and concentrates.
The discharge from the autoclave is cooled in a series of flashing steps in
which, depending on the heat balance in the autoclave, heat is either re-
covered and recycled, or is vented as steam to the atmosphere, through a suit-
able scrubber. Depending on the acid concentration in the autoclave liquor,
and whether or not acid-recycling to the feed is necessary, the autoclave dis-
charge slurry will either be neutralized directly prior to cyanidation and CIP
(as practised at McLaughin and Goldstrike), or the autoclave liquor will be
separated from the solids by filtration or CCD. This is an expensive option
because the solid-liquid separation equipment must all be acid-proof, but it
does benefit the down-stream cyanidation and CIL or CIP operations in sev-
eral ways. Firstly, separation of the autoclave liquor removes base metals that
would otherwise consume cyanide and could also cause problems in CIP and
elL through formation of slimy hydroxide precipitates. Secondly, the re-
moval of most of the acid with the autoclave liquor reduces the amount of
gypsum formed during neutralization of the slurry prior to cyanidation. This
reduces the viscosity of the slurry and results in improved CIP JCIL perform-
ance. Finally, the feed to cyanidation is a lot cooler after solid-liquid separa-
tion than when the autoclave discharge is neutralized directly, and the lower
temperature also benefits cyanidationjCIL performance significantly.
Bioleaching
With the world-wide interest in low-grade refractory gold ores, the "new"
technology of bacterial heap leaching is being actively investigated by several
mining companies. The greatest interest, however, is focussed on the treat-
ment of high-grade concentrates, by bacterial oxidation of finely-milled solids
in stirred-tank reactors. The fundamental difference between the refractory
gold applications and those involving copper and uranium minerals lies in
the fact that the metal of interest, gold, does not dissolve during bacterial
leaching but, as in roasting and pressure leaching, is liberated for subsequent
recovery by cyanidation.
The chemical reactions involved are the same as those shown previously
for pyrite and arsenopyrite in pressure leacHing (eqs. (16) and (17. Ar-
senic and iron dissolve, liberating gold, but unlike the situation in an auto-
clave, they do not reprecipitate under the ambient conditions prevailing in a
bioleach reactor. The build-up of iron, arsenic and sulphuric acid (and tem-
perature) must be controlled, in fact, because they limit the rate of bacterial
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 157
reproduction. For this reason, bioleaching of sulphide concentrates is con-
ducted at low solids densities, usually around 10-15%. The pH needs to be
controlled in the 1.5-2.0 range, and the excess sulphuric acid generated by
oxidation of sulphides is neutralized by adding lime slurry to the bioleach
reactor. The temperature must be controlled in the 35-40C range for Thioh-
acillus ferrooxidans and, because of this limitation, some research work is
being done with su/folobus, a bacterial strain preferring higher temperatures.
Iron and arsenic levels must also be controlled.
A typical plant treating concentrate will receive a feed containing anywhere
from 10% to 30% sulphide and up to 25% arsenic. The iron content will de-
pend on the relative amounts of pyrrhotite, pyrite and arsenopyrite, but there
should be at least a stoichiometric equivalent of iron to arsenic to ensure com-
plete precipitation offerric arsenate rather than calcium arsenate, during neu-
tralization after bioleaching. The stability of the precipitate apparently in-
creases with increased iron to arsenic ratio, and a molar ratio of 4: 1 is
recommended [59]. The feed to bioleaching may be pretreated with fresh or
recycled acid if it contains more carbonates than sulphides. The oxidation
tanks are usually configured in four stages, with additional tanks employed in
the first stage to accommodate the increased oxidation rates and aeration re-
quirements in the initial phases of the reaction. Overall residence time in the
bioi each reactors is of the order of 3-4 days. The air supply to the reactor
must supply both the oxygen needed to oxidize the sulphides and the agita-
tion needed to suspend the ore particles, and the demand by the former is
much greater than the latter, particularly in the early stages. Consequently,
there has been a lot of research into the method of distributing air into the
slurry and maximizing mass transfer of oxygen in the slurry. Oxygen transfer
efficiency is obviously lower for a process that uses air than for one that uses
pure oxygen, such as pressure leaching. However, air is preferred to oxygen as
the source of oxygen, firstly because of cost, and secondly because the bacteria
also require carbon dioxide to survive. A typical range of theoretical oxygen
demands would be 100-500 kg of oxygen per ton of concentrate. The supply
of air to the reactor will also remove some of the heat generated by the oxi-
dation reaction, but in practice additional heat removal (heat exchangers)
will probably be necessary to control the temperature at -- 40 C.
Nutrients to stimulate bacterial reproduction are added with the feed to the
first reactor. The optimum nutrients and their dosage need to be established
for each application, but will usually consist of a mixture of ammonium, po-
tassium and phosphate ions. Some of these may occur naturally in the ore,
and the balance can generally be supplied in the form of a commercially avail-
able fertilizer.
A solid-liquid separation step is generally carried out after bioleaching be-
cause of the low solids density of the slurry, and thickening would probably
be preferred because oxidized ores do not filter well. Part of the overflow
158
C.A. FLEMING
could be recycled to the feed as a source of make-up acid and bacteria, al-
though this is not critical, because the bacteria reproduce very rapidly in the
bioleach reactors. The remaining liquor would be neutralized with lime to
produce the stable, basic ferric arsenate complex, (FeAs0
4
)Fe(OHh, and
gypsum. The thickener underflow could be washed in a CCO circuit if cya-
nide consumption is high, or could be treated directly by lime neutralization,
cyanidation and CIP.
An important characteristic of bacterial leaching reactions is that they are
autocatalytic, i.e. the bacteria are themselves a product of the sulphide oxi-
dation reaction, and much of the progress that has been made in the last 20
years for the treatment of concentrates has been in the development of bac-
terial strains that multiply rapidly in the aggressive conditions that prevail in
the exothermic sulphide oxidation reaction. This was an essential develop-
ment before even small-scale bioleaching of milled concentrate in stirred re-
actors could be considered for industrial application. For example, in the early
pilot plant work at the Gencor Laboratory in South Africa, it was found that
a 10 day residence time was needed to achieve sufficient oxidation of an ar-
senopyrite concentrate for 97% gold recovery. When this is compared with
the 1-2 h residence time in a typical pressure leaching operation, some insight
into the relative sizes of a pressure leaching and a bioleaching plant can be
gained. After 2 years of operation of the Gencor pilot plant, the bacteria had
adapted and mutated to the extent that the retention time had decreased to 4
days [60]. At that time a decision was taken to build an industrial bioleach-
ing plant to treat this arsenopyrite concentrate at the Fairview Gold Mine in
South Africa. The plant, which was the first of its kind in the world, was com-
missioned in 1986, with a design capacity of lOt of concentrate per day. This
represented 40% of the concentrate being produced at Fairview. In the 5 years
that the plant has been operating, further improvements and refinements have
allowed the capacity of the original plant to be increased to about 17 t d - I,
with gold recoveries typically around 95% (compared with <90% recovery
from the roaster calcine at the same mine). Gencor metallurgists are suffi-
ciently confident with the process that bioleach capacity is currently being
expanded at Fairview to treat all of the concentrate produced, which will al-
low the old Edwards roasters to be decommissioned. Gencor are also building
a bioleaching plant at their Sao Bento operation in Brazil, which will probably
be operated in series with the pressure leaching facility that is already there,
to accommodate the planned increase in production at that mine [61 ].
Another bioleaching operation in South Africa is the 20 t d - 1 pilot plant at
Vaal Reefs, which oxidizes a pyrite concentrate containing about 5 g t-
I
of
gold, prior to uranium and gold recovery. This operation was developed
through ajoint bioleaching research program involving both the Anglo Amer-
ican Corporation and Mintek. A third commercial undertaking, at Tonkin
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 159
Spring in Nevada, was operated for a short while in 1990, but ran into tech-
nical difficulties. This plant is currently mothballed while new partners in the
venture re-evaluate the technology and the ore reserves. The treatment of
concentrate in stirred reactors is also being piloted at the present time at
Dickenson Mine in northwestern Ontario. This will be a small-capacity plant,
and it is therefore well-suited to bioleaching because of the lower capital costs,
on this scale, compared to pressure leaching and roasting. In fact, biological
oxidation may well find its niche in the treatment of small tonnages of flota-
tion concentrate, particularly arsenopyrite, because of the significant capital
cost advantages. The other situation where bioleaching economics may com-
pare favourably with other processes, is one in which only partial sulphide
oxidation is needed for good gold liberation. Because of the very much slower
kinetics of sulphide oxidation in a bioleach reactor, partial oxidation is far
easier to control than in a roaster or an autoclave.
CONCLUSIONS
The cyanide leaching process will continue to playa dominant role in the
processing of free-milling and refractory (after pretreatment) gold ores for
the foreseeable future. New leaching processes may find limited application
in niche areas, such as the treatment of acidic slurries generated in pressure
leaching or bioleaching.
It is unlikely that any more plants incorporating solid-liquid separation
(by either filtration or CCD) will be built, unless gold and silver values are
very high (> 20g t-
1
combined Au, Ag). Carbon-in-pulp and carbon-in-Ieach
will be the preferred processing routes for the treatment of cyanide leach slur-
ries for at least the next 5 years, and possibly longer. Resin-in-pulp has many
attractive features, and could gradually replace carbon-in-pulp if the process
can be successfully engineered on a large scale.
The trend towards increasing refractoriness in gold and silver ores will con-
tinue, and more and more gold plants will incorporate pretreatment process-
ing to liberate precious metals. The choice of pretreatment will depend on ore
mineralogy and site-specific economic and environmental factors. New pre-
treatment technologies such as bioleaching, nitric acid leaching and ultra-fine
grinding have many attractive features and will increasingly challenge the
"established" processes of roasting and pressure leaching.
REFERENCES
Young, G.J.C., Cyanidation. In: G.G. Stanley (Editor), The Extractive Metallurgy of Gold
in South Africa. S. Afr. Inst. Min. Metall., Monogr. Ser., M7 ( 1987): 277-330.
160 C.A. FLEMING
2 Bosley, D., Recovery of gold from solution by cementation. In: G.G. Stanley (Editor), The
Extractive Metallurgy of Gold in South Africa. S. Afr. Inst. Min. Metall., Monogr. Ser., M7
(1987): 331-344.
3 Kudryk, V. and Kellogg, H.H., Mechanism and rate-controlling factors in the dissolution
of gold in cyanide solution. J. Metals, 6 (5) (1954): 541-548.
4 Cathro, K.J. and Koch, D.F.A., The dissolution of gold in cyanide solution-an electro-
chemical study. Proc. Australas. Inst. Min. Metall., 210 ( 1964): 111-126.
5 Fink, e.E. and Putnam, E.L., The action of sulphide ions and of metal salts on the distri-
bution of gold in cyanide solutions. Trans AIME, 187 (1950): 952-955.
6 Nicol, M.J., The anodic behaviour of gold, Part II. Oxidation in alkaline media. Gold Bull.,
13 (37) (1980): 105-111.
7 Hancock, R.D. and Finkelstein, N.P., EO / EO diagrams: their uses in estimating unknown
standard reduction potentials with particular reference to d 10 electronic configuration metal
ions. Nat. Inst. Metall., Johannesburg, Rep. 1153 (1970).
8 Nicol, M.J., Heming, e.A. and Paul, R.L., The chemistry of the extraction of gold. In: G.G.
Stanley (Editor), The Extractive Metallurgy of Gold in South Africa. S. Afr. Inst. Min.
Metall., Monogr. Ser., M7 (1987): 831-906.
9 Pietsch, H., Tuerke, W. and Bareuther, E., Leaching of gold and silver. German Pat. DE
3,126,234 (1983).
10 Muir, e. W.A., Hendricks, L.P. and Gussman, H. W., The treatment of refractory gold bear-
ing flotation concentrates using pressure leaching techniques. In: V. Kudryk, D.A. Corrigan
and W.W. Liang (Editors), Precious Metals Mining, Extraction and Processing. AIME,
Warrendale, Pa. (1984), pp. 309-322.
II Murray, R.H., Van Aswegen, P.e. and Elmore, e.L., Kamyr countercurrent tower gold
extraction process. S. Afr. Inst. Min. Metall., Colloq. on Leaching, MINTEK (November
1988), Pap. 3.
12 Loroesch, J., Peroxide assisted leach; three years of increasing success. Proc. Randol Gold
Forum (Squaw Valley, Calif.) (Sept. 1990),pp. 215-220.
13 Shoemaker, R.S. and Dasher, J., Recovery of gold and silver from ores. Int. Precious Met-
als Inst. Seminar on Refining of Precious Metals. Skytop, Pa, 669.21/23 (082) Int. (1980).
14 Leblanc, R., Precipitation of gold from cyanide solution by zinc dust. The influence of
certain factors and impurities. Can. Min. J., 63 (1942): 213-219; 297-306; 371-379.
15 Nicol, M.J., Schalch, E. and Balestra, P., A modern study of the kinetics and mechanism of
the cementation of gold. J. S. Afr. Inst. Min. Metall., 79 (1979): 191-197.
16 McIntyre, J .D.E. and Peck, W.F., Electrodeposition of gold. Depolarisation effects induced
by heavy metal ions. J. Electrochem. Soc., 123 (1976): 1800-1813.
17 Plaskin, LN., Suvorovskaya, N.A. and Budnikova, O.K., Precipitation of metals from so-
lution by metals. Izv. Akad. Nauk SSSR, Otd. Tekh. Nauk, (1948): 131-138.
18 Hancock, H.A. and Thomas, G., The effect of nickel and hydrogen overvoltage on the pre-
cipitation of gold. Can. Min. Metall. Bull., 57 (1954): 337-346.
19 Nagy, I., Mrkusic, P. and McCullock, H.W., Chemical treatment of refractory gold ores.
Nat. Inst. Metall., Johannesburg, Rep. 38, 1966.
20 Lazowski, M., On some properties of carbon. Chem:Gaz, 6 ( 1948): 43.
21 Johnson, W.D., Abstraction of gold and silver from their solutions in potassium cyanide.
US Pat. 522260 (1894).
22 Gross, J. and Scott, J.W., Precipitation of gold and silver from cyanide solution on char-
coal. US Bur. Mines, Washington, Tech. Pap., 378 (1927).
23 Zadra, J.B., A process for the recovery of gold from activated carbon by leaching and elec-
trolysis. US Bur. Mines, Reno, Nev., Rep. Invest. 4672 (1950).
24 Zadra, J .B., Engel, A.L. and Heinen, H.J. Process for recovering gold and silver from acti-
vated carbon by leaching and electrolysis. US Bur. Mines, Reno, Nev., Rep. Invest., 4843
(1952).
HYDROMETALLURGY OF PRECIOUS METALS RECOVERY 161
25 McDougall, G.J. and Fleming, C.A., Extraction of precious metals on activated carbon. In:
M. Streat and D. Naden (Editors), Ion Exchange and Sorption Processes in Hydrometal-
lurgy. Soc. Chemical Industry, Chichester, 19 (1987), pp. 56-126.
26 Hall, K.B., Homestake carbon-in-pulp process. Am. Mining Congress Meeting (Las Vegas.
Nev.) (1974), 16pp.
27 Laxen, P.A., Becker, G.S.M. and Rubin, R., Developments in the application of carbon in
pulp to the recovery of gold from South African ores. J. S. Afr. Inst. Min. Metall., 79 ( II )
(1979): 315-326.
28 Laxen, P.A., Carbon in pulp in South Africa. Hydrometallurgy, 13 (2) (1984): 169-192.
29 Fleming, C.A., Recent developments in carbon-in-pulp technology in South Africa. In: K.
Osseo Asare and J.D. Miller (Editors), Hydrometallurgy-Research, Development and
Plant Practice. Metall. Soc. AIME, Warrendale, Pa. (1982), pp. 839-857.
30 Davidson, R.J., Metal, particularly gold recovery. S. Afr. Pat. 73/8939 (23 Nov. 1973).
31 Paul, R.L., Filmer, A.O. and Nicol, MJ., The recovery of gold from concentrated auro-
cyanide solutions. In: K. Osseo Asare and J.D. Miller (Editors), Hydrometallurgy-Re-
search, Development and Plant Practice. Metall. Soc. AIME, Warrendale, Pa. ( 1983), pp.
689-704.
32 Davidson, RJ., The mechanism of gold adsorption on activated charcoal. J. S. Afr. Inst.
Min. Metall., 75 (4) (1975): 67-76.
33 McDougall, GJ., Hancock, R.D., Nicol, M.J., Wellington, O.L. and Copperthwaite, R.G.,
The mechanism of the adsorption of gold cyanide on activated carbon. J. S. Afr. Inst. Min.
Metall., 80 (9) (1980): 344-356.
34 Adams, M.D. and Fleming, C.A., The mechanism of adsorption of aurocyanide onto acti-
vated carbon. Metall. Trans. B., 20B (1989): 315-325.
35 Adams, M.D. and Nicol, MJ., The kinetics of elution of gold from activated carbon. In:
C.E. Fivaz and R.P. King (Editors), Gold 100. Proc. Int. Conf. on Gold, Extractive Me-
tallurgy of Gold. S. Afr. Inst. Min. Metall., Johannesburg, Vol. 2 (1986), pp. 111-121.
36 Fleming, C.A. and Nicol, M.J., The adsorption of gold cyanide onto activated carbon, III.
Factors influencing the rate ofloading and equilibrium capacity. J. S. Afr. Inst. Min. Me-
tall.,84 (4) (1984): 85-93.
37 Nicol, M.J., Fleming, C.A. and Cromberge, G., The adsorption of gold cyanide onto acti-
vated carbon, I. The kinetics of adsorption from pulps; II. Application of the kinetic model
to multi-stage adsorption circuits. J. S. Afr. Inst. Min. Metall., I, 84 (2) (1984): 50-54; II.
84 (3) (1984): 70-78.
38 Wollacott, L.C., Stange, W. and King, R.P., Towards more effective simulation ofCIP and
CIL processes. Part I. The modelling of adsorption and leaching. Part 2. A population-
balance-based simulation approach. Part 3. Validation and use of a new simulator. J. S.
Afr. Inst. Min. Metall., 1,90 (10) (1990): 275-282; 2, 90 (11) (1990): 307-314; 3, 90
(12) (1990): 323-332.
39 Fleming, C.A. and Cromberge, G., Small-scale pilot plant tests on resin-in-pulp extraction
of gold from cyanide media. J. S. Afr. Inst. Min Metall., 84 (1984): 369-378.
40 Adams, M.D., The chemical behaviour of cyanide in the extraction of gold. Part 1. Kinetics
of cyanide loss in the presence and absence of activated carbon. J. S. Afr. Inst. Min. Metall.,
90 (2) (1990): 37-44.
41 Tsuchida, N., Ruane, M. and Muir, D.M., Studies on the mechanism of gold adsorption on
carbon. In: L.F. Haughton (Editor), Mintek 50, Int. Conf. on Recent Advances in Mineral
Science and Technology, Sandton, Vol. 1 (1984), pp. 647-656
42 Bailey, P.R., Application of activated carbon to gold recovery. In: G.G. Stanley (Editor),
The Extractive Metallurgy of Gold in South Africa. S. Afr. Inst. Min. Metall., Monogr. Ser.,
M7 (1987): 379-614.
43 Anon, Gold from Russia's Muruntau deposit. Coal Gold Base Miner., 26 (1978): 75-81.
162 C.A. FLEMING
44 Fleming, C.A., Resin-in-pulp as an alternative process for gold recovery from cyanide leach
slurries. Proc. 23rd Annu. Meet. Canadian Mineral Processors, Can. Inst. Min. Metall.,
Ottawa (1991 ).
45 Schwellnus, A.H. and Green, B.R., Structural factors influencing the extraction of gold
cyanide by weak-base resins. In: M. Streat (Editor), Ion Exchange for Industry. Soc. Chern.
Ind., London ( 1988), pp. 207-218.
46 Green, 8.R., Ashurst, K.G. and Chanston, T.E., Dedicated resins for gold. Gold Forum on
Technology Practices, World Gold '89. Soc. Min. Metall. Exploration, Colo. (1989), pp.
341-353.
47 Fleming, C.A. and Seymore, D., Golden Jubilee resin-in-pulp plant for gold recovery. Proc.
Randol Gold Conf. (Sacramento, Calif.) (Nov. 1989), pp. 297-307.
48 Fleming, C.A. and Seymore, D., The Golden Jubilee RIP plant-recent changes and im-
provements in plant performance. Proc. Randol Gold Forum (Squaw Valley, Calif.) (Sept.
1990),pp.237-242
49 Merwin, R.M., Potter, G.M. and Heinen, H.J., Heap leaching of gold ores in Northeastern
Nevada. AIME, New York, Repr., 69-AS-79 (1969).
50 Heinen, H.J. and Porter, B., Experimental leaching of gold from mine waste. US Bur. Mines,
Rep. Invest., RI 7250 ( 1969).
51 Potter, G.M. and Salisbury, H.B., Innovations in gold metallurgy. Min. Congr. J., (July
1974): 54.
52 Ackerman, J.B. and Bucans, C.S., Plant and process startup of the Sunshine silver refinery.
Miner. Metall. Process., (Feb. 1986): 20-32.
53 Guay, W.J. and Peterson, D.G., Recovery of gold from carbonaceous ores at Carlin, Ne-
vada. Trans. S.M.E., AIME, 254 (1973): 102-104.
54 Folland, G. and Peinemann, B., Lurgi's circulating fluid bed applied to gold roasting. Eng.
Min.J., (October 1989).
55 Deter, K.W. and McCord, T.H., Oxygen whole ore roasting at Jerritt Canyon Joint Ven-
ture. SME Annu. Meet. (Denver, Colo.) (Feb. 1991).
56 Mason, P.G., Energy requirements for the pressure oxidation of gold-bearing sulphides. J.
Metals, (Sept. 1990): 15-18.
57 McKay, D.R. and Halpern, J., A kinetic study of the oxidation of pyrite in aqueous suspen-
sion. Trans. Met. Soc. AIME, (June 1958): 301-309.
58 Conway, M.H. and Gale, D.C., Sulphur's impact on the size of pressure oxidation auto-
claves. J. Metals, (September 1990): 19-22.
59 Haines, A.K., Factors influencing the choice of technology for the recovery of gold from
refractory arsenical ores. In: C.E. Fivaz and R.P. King (Editors), Gold 100. Proc. Int. Conf.
on Gold, Extractive Metallurgy of Gold. S. Afr., Inst. Min. Metall., Johannesburg, Vol. 2
(1986),pp.227-234.
60 van Aswegen, P.c., Haines, A.K. and Marais, H.J., Design and operation ofa commercial
bacterial oxidation plant at Fairview. Proc. Randol. Gold Symp., Perth (1988), pp. 144-
147.
61 Van Aswegen, P.c., Godfrey, M.W., Miller, D.M. and Haines, A.K., Developments and
innovations in bacterial oxidation of refractory ores. SME Annu. Meet. (Denver, Colo.)
(Feb. 1991).
62 Bhappu, R.B., Lewis, M.F. and McAllister, J.A., Leaching of low-grade ores, economic
evaluation of available processes. AIME Annu. Meet. (Dallas, Texas) (Feb. 1974).

Você também pode gostar