You are on page 1of 20

 

 
Heap leaching as a key technology for recovery of values from low-grade ores
– a brief overview

Jochen Petersen

PII: S0304-386X(15)30077-3
DOI: doi: 10.1016/j.hydromet.2015.09.001
Reference: HYDROM 4158

To appear in: Hydrometallurgy

Received date: 15 June 2015


Revised date: 27 August 2015
Accepted date: 3 September 2015

Please cite this article as: Petersen, Jochen, Heap leaching as a key technology for
recovery of values from low-grade ores – a brief overview, Hydrometallurgy (2015), doi:
10.1016/j.hydromet.2015.09.001

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT

Heap Leaching as a Key Technology for Recovery of Values from Low-


grade Ores – A Brief Overview
Jochen Petersen

Department of Chemical Engineering, University of Cape Town, Rondebosch 7700, Cape Town, South Africa,

T
Tel: +27 21 650 5766; email: Jochen.petersen@uct.ac.za

P
RI
Abstract:

SC
Heap Leaching Technology is finding increasingly widespread application to recover values
from low-grade ores, especially in the gold and copper industry. The particular attractiveness
of the process is the relatively low-cost recovery of the target metal at site without the need

NU
for energy intensive comminution, providing the process can be combined with a selective
solution recovery technology (for example solvent extraction). However, this is contrasted
with often slow and inefficient recovery, and technical issues such as poor heap permeability
MA
and post-closure stability. This paper gives a comprehensive overview of the principles of
heap leaching, offers a critical analysis of the economic viability of the process and how
certain technical drawbacks of the technology affect this, as well as providing a brief
ED

overview of emerging and potential future applications of the technology.

Keywords: heap leaching, recovery, leaching kinetics, economics, gold, copper, zinc,
uranium, e-waste
PT
CE

Introduction

Heap leaching forms part of the group of technologies known as percolation leaching, which
AC

includes in situ leaching, dump leaching, heap leaching and vat leaching (Bartlett, 1998;
John, 2011). Common to all these technologies is migration of leach solution through a fixed
bed of ore particles, on its path interacting with the solid, releasing the target minerals into
solution with which they are carried out of the bed. The target species are subsequently
recovered from solution using conventional hydrometallurgical techniques, especially solvent
extraction (SX), before returning the barren solution to the leach process.

The distinction between the different processes lies in the degree of preparation of the ore
bed. In in situ leaching the ore remains underground and solution flows either through the
natural porosity of the ore (typical for example for uranium bearing sandstones), or a pore
structure is created through some form of mechanical cracking (in-situ blasting or hydraulic
fracturing). Dump leaching refers to as-mined ore that is piled in rock depositories
(‘dumps’), which are irrigated with leach solution that percolates through the bed and is
collected at the base for further treatment. This process is applied usually only to very low-
grade waste rock as a ‘value-add’, since the costs of installing the irrigation system and
treating the resulting stream are marginal. Due to the absence of any crushing prior to

1
ACCEPTED MANUSCRIPT

stacking, the particle size ranges from a few centimetres to large boulders of one meter and
more, and consequently dump leaching is associated with poor extractions (20-30%). In heap
leaching the ore is crushed, usually to below 1 in. (25 mm), and heaps are much more
carefully engineered to maximise percolation of leach solution through them. The higher
investment in operating the process is offset by significantly higher recoveries (typically

T
>70%) and the process is applied often to low- to intermediate grade ores. Heaps are also
operated for limited time periods before being removed or abandoned as a permanent deposit.

P
Finally, vat leaching refers to placing more finely crushed ore (1-10 mm) in a large basin

RI
where it is subsequently flooded with leach solution and left to react, before draining the
solution again for treatment and removing the ore for final storage.

SC
Heap leaching is practised all over the world for three key applications: cyanide leaching of
gold ores, acid leaching for copper oxide ores and oxidative acid leaching of secondary

NU
copper sulphide ores, invariably with the assistance of certain microorganisms. A few heap
leach operations for uranium also exist (acid and alkali leaching), although the technology is
less common, partly due to environmental concerns (Scheffel, 2002). Uranium heap leaching
MA
has been pioneered since the 1950s, whereas both, copper oxide and gold/silver heap leaching
emerged in the US from the late 1960s (Kappes, 2002; Bartlett, 1998). Large scale copper
heap leaching commenced in Chile from the 1980s, and sulphide heap leaching emerged
essentially as a consequence of acid leaching from mixed oxide/sulphide ores from the 1990s
ED

(Watling, 2006).
PT

Principles of Heap Leaching

A detailed description of the heap leach process and the underlying mechanisms has been
CE

given elsewhere (Petersen and Dixon, 2007a, 2007b; Watling, 2006; Bartlett, 1998), and only
some key aspects are discussed here. Figure 1 shows in a schematic representation a typical
copper sulphide heap leach circuit – the principle is the same for other heap leach processes.
AC

Figure 1: Schematic of a typical copper sulphide heap leach circuit

2
ACCEPTED MANUSCRIPT

A heap is a constructed pile of crushed, and in most cases agglomerated, rock material built
on an impermeable under-liner fitted with a solution collection system. Where sulphide
minerals are targeted, aeration pipes are also placed underneath the heap. The heap is stacked
by conveyor belt; truck dumping is also practiced, but generally results in undesirable

T
compaction of the heap surface by the trucks. Heap heights are typically 6-10 m, but taller

P
heaps are also common to reduce the footprint of the operation. The heap is irrigated from the

RI
top surface, either by sprinklers, sprays or drip emitters.

Typical irrigation rates are in the order of 5-20 L/m2/h, and aeration, where present, in to

SC
order of 0.1-0.5 m3/m2/h. Irrigation is mostly continuous, but intermittent (on-off) schemes
are used in some operations. Solution percolates downwards through the crushed ore bed
(typical top particle size is ¾ in. to 1 in. (19-25mm)) where it partially saturates pore spaces

NU
and partially migrates through the bed in discrete flow channels. Where aeration is present,
air migrates upwards counter-currently to solution through unsaturated void spaces. Mineral
grains contained in the rock interact with chemical species in the solution (introduced with
MA
the feed) to dissolve and release dissolved species into solution, which are then transported to
the base of the heap and recovered as pregnant leach solution (PLS) in the solution collection
system. Target metals are then recovered from the PLS, and the barren solution is made up
ED

with fresh reagents and returned to the heap.

Three key types of chemistry are employed in heap leaching at present: sulphuric acid
leaching predominantly of copper oxides, alkali cyanide leaching of gold/silver bearing ores
PT

and oxidative sulphuric acid leaching of secondary copper sulphides (usually in the context of
supergene porphyry ores). The latter process is invariably catalysed by microorganisms,
CE

which colonise within the heap bed (Demergasso, 2005), and facilitate the oxidation of
ferrous to ferric by atmospheric oxygen as well as the oxidation of reduced sulphur
intermediates (Sn2-, S2O32-, etc) to sulphate (Sand et al., 2001). Bio-oxidation occurs naturally,
AC

but in some operations is promoted through the deliberate inoculation of heaps with microbial
cultures (Gericke, 2011). Heap bio-leaching has also been demonstrated for the pre-oxidation
of refractory gold ores prior to cyanide leaching (Logan et al., 2007).

At the particle level, heap-leaching is governed by the migration of reactive species into and
dissolved species out of pores and cracks, which is based on molecular diffusion, invariably a
very slow process (Sánchez-Chacón and Lapidus, 1997). Overall leach kinetics in a heap are
a complex interplay between reagent transport to and from site, gas-liquid mass-transfer
between the air and solution phases, as well as migration through stagnant liquid in
agglomerates and through particle pores. Further effects are microbial colonisation behaviour,
mineral location and liberation within particles, mineral reaction kinetics as well as
generation and dissipation of reaction heat (through exothermic sulphide leaching) (Petersen
and Dixon, 2007a).

Leach periods typically extend to 60-100 days for heap cyanidation, 4-6 months of acid
copper leaching, and 1-3 years for secondary copper sulphides. The long leach times,

3
ACCEPTED MANUSCRIPT

especially of sulphide minerals, has resulted in this technology being employed only for low-
grade ores for which other technologies, requiring a higher degree of comminution, are
uneconomical. Further, to achieve meaningful overall production rates, very large inventories
of ore need to be kept under leach at the same time, occupying a sizeable footprint and
solution inventory for operation. As a consequence, many heap operations employ multiple

T
lifts, with new ore being stacked on top of spent heaps (Bartlettt, 1998; Scheffel, 2002).

P
The PLS concentrations of the target metals are usually low (2-10 g/L for Cu, ppm levels for

RI
Au) and, especially in acid leach solutions, many impurities are present. Selective recovery of
copper by the highly selective SX reagents, and of gold cyanide onto activated carbon ensure

SC
the success of heap leaching of these commodities. Raffinate solution is, after make-up,
recycled to the heap. As a result there is a gradual build-up of dissolved species (especially
Al, Mg and Fe sulphates), which at critical concentrations may begin to precipitate within the

NU
heap and consequently block pores and sprinkler nozzles. Adverse effects of such dissolved
species on bioleaching microorganisms have also been observed (Ojumu, 2008), but
management of bleed solutions in industrial heaps is not systematically practised.
MA
Irrigation is usually discontinued after around 80-90% extraction for gold and copper oxide
leaching and 70-80% for copper sulphide leaching. The reason is that extraction rates become
so slow that the costs of continued operation begin to outweigh the value of metal recovered.
ED

Spent heaps are either, reclaimed and deposited on a separate waste ore dump to use the
cleared lined ground for a new heap, or a new lift is built on top of the old heap, with a view
to using the existing infrastructure and potentially recovering additional unleached values, as
PT

leach solution from the fresh ore percolates through the spent ore underneath (Scheffel,
2002).
CE

Optimally, heap leaching should be a low-cost technology for the recovery of values from
low-grade ores, suitable especially for remote mine sites. However, the complexities that
govern the leach kinetics are often underestimated, resulting in underperforming heaps.
AC

Historically, treatment of low-grade ores by heap leaching was seen as an add-on process to
recover additional values from a particular ore body without it determining the profitability of
the operation as a whole. Hence, less care was applied in their design, and the understanding
of the process was built more on operator experience than hard science.

As heap-leach-only operations are becoming more frequent, and operating margins even of
operations that employ multiple extractive technologies (i.e. concentration of higher grade
portion of the ore body by flotation and heap leaching of the low-grade rest) are tightening,
more attention needs to be focussed on improving heap performance in terms of rate of
extraction as well as total extraction. For this it is first necessary to understand what limits
heaps under current operating conditions.

Extensive research by the author and others over recent years has clearly identified a complex
scenario of interactions in which different mechanisms are limiting the overall process at
different times and at different places in the heap. It is convenient to distinguish between four
different scales of reaction-transport phenomena in a heap, with different limitations applying

4
ACCEPTED MANUSCRIPT

in each case (Petersen and Dixon, 2007a, 2007b). This is illustrated in Figure 2 for the case of
heap bioleaching.

T
Level Sub-processes Illustration
 Solution flow

P
through solution flow

RI
packed bed
Heap
 Gas advection internal heat
generation
Scale
 Water vapour

SC
gas flow
transport
 Heat balance

NU
 Gas adsorption O2, CO2

 Pore diffusion
adsorption

 Microbial inter-particle
MA
diffusion
Agglo growth
merate  Microbial
Scale attachment intra-particle
diffusion
 Microbial
ED

oxidation attached and floating


micro-organisms

 Particle shape
PT

mineral grains
effects
cracks and pores
 Intra-particle
Particle host rock and gangue
diffusion
Scale
CE

 Particle and particle surface


grain size
distribution
AC

 Ferric/ferrous
Fe3+
reduction
 Mineral MeS
Grain oxidation Fe2+
Scale  Sulphur
oxidation Me2+
 Surface SO42- S0

processes

Table 1: Schematic representation of sub-processes in heap bioleaching (Petersen and Dixon,


2007b)

5
ACCEPTED MANUSCRIPT

Heap-scale:

Solution migrates through the ore bed principally in a downward direction through the void
spaces. While finer pores tend to become saturated, larger pores remain unsaturated and over
time solution flow becomes concentrated in random channels through unsaturated zones. This
does not mean that portions of the heap bed remain dry. Rather, the heap bed contains zones

PT
of wet, but stagnant solution contained in the interstitial spaces between particles, in between
a network of channels of flowing solution and gas pores. An image of this network, taken of a
laboratory scale packed bed column using a high-resolution positron emission tomography

RI
(PET) scanner, is given in Figure 2. It can clearly be seen that the tracer follows discrete
solution channels rather than flowing homogenously through the wetted bed.

SC
NU
MA
D
TE
EP

Figure 2: PET scan of a 10 cm diameter x 40 cm long column filled with a copper porphyry
C

ore irrigated form a central dripper (picture taken at the PET/PEPT facility at iThemba Labs,
AC

Western Cape, South Africa). The colour scale from red to blue directly relates to decreasing
tracer concentration.

Gas transport through the heap bed is assumed to be principally upwards through unsaturated
void spaces over which it competes with the flowing aqueous phase. In heaps where no active
aeration is present, its still occurs through thermal up-drafting – reaction heat released
through the leach reactions cause heating of the gaseous stream which therefore accelerates,
causing a thermal draft drawing in cold air from the sides of the heap. But heat effects remain
even if aeration is provided by pipes from the base of the heap. Air on its way up through the
heap becomes saturated with water vapour, the more so the higher the temperature. Moist air
is the main mode of heat transport from the bed and contributes to significant solution losses
by evaporation – however aeration can result in a homogenisation of temperatures within a
heap, leading to overall improvement of performance (Dixon, 2000).

6
ACCEPTED MANUSCRIPT

Agglomerate Scale:

A collection of particles held together by stagnant solution is referred to as an agglomerate –


in many heap operations ore agglomeration is practiced for dust/fines control and
introduction of leach reagent (especially curing acid) prior to heap operation. During
stacking, particle agglomerates form further as they roll off the stacking face, often resulting

T
in an accumulation of larger particles and agglomerates near the bottom of the heap. It is not

P
known, however, whether in the course of heap operation agglomerate structures created

RI
during stacking are retained, or merge into larger zones upon wetting. Some consolidation of
the packing certainly takes place, as heaps commonly ‘slump’ over the first few days of

SC
irrigation.

Transport of dissolved chemical species as well as bacteria in the agglomerate structure


progresses primarily by diffusion; bulk migration by capillary action occurs only initially

NU
upon wetting. Hence diffusion of reagents and reaction products through the stagnant
agglomerate zone can be a major rate-limiting effect during heap leaching and is critically
dependent on the average size of the agglomerate zone. This size can be understood to
MA
correspond to the average distance from the nearest flowing solution channel through which
dissolved species can enter and be removed.

Microbial interactions take place within the agglomerate structure, as microbes cannot
ED

penetrate ore particles. Microbes colonise heaps firmly attached to the mineral surface,
surface associated or freely in solution (Govender et al., 2013). Microbial growth relies on
the uptake of CO2 from the gas stream or carbonates in the ore dissolving in the acidic
PT

environment. Microbial speciation in heaps is complex and community structures vary widely
depending on local conditions and over time. Limited availability of CO2 in the gas stream
CE

can be severely rate limiting in tall heaps due to its low concentration in air (Petersen et al.,
2010).
AC

Particle-Scale

Target minerals are embedded within the particles within a matrix of gangue minerals.
During leaching, reagents (acid, cyanide, ferric or oxygen) need to migrate through pores and
(micro) cracks in this matrix to reach the relevant mineral phases, again by a process of
diffusion. Different from the agglomerate scale diffusion, however, the porosity of a rock
particle is substantially lower and hence few and usually tortuous channels for this diffusion
exist, slowing the process. Some of the mineral may in fact be inaccessible altogether. The
study by Ghorbani et al. (2013) has shown that the density of cracks is strongly related to the
mode of crushing and that in larger particles there is an inner core that remains unaffected by
crushing and essentially inaccessible to leaching. Through using X-ray tomography,
Ghorbani et al. (2013) have also shown that leaching progresses in a reaction front migrating
into the inner particle, but stopping short of the core. The progression of a reaction front
within particles is further compromised by the reaction of gangue or non-target minerals with
the diffusing reagent, especially acid. While this may on the one hand increase mineral
porosity, it contributes to reagent consumption relative to metal recovery.

7
ACCEPTED MANUSCRIPT

Mineral grain scale:

At the scale of the individual mineral grain within the ore particle the same chemical
interactions take place as in concentrate/pure mineral leaching, and the kinetics are governed
by the local chemical conditions at the mineral surface in terms of reactant concentrations,
pH, the presence of other ions and especially temperature. However, in whole particle

T
leaching galvanic interactions between different minerals closely inter-grown at the grain

P
scale can play a significant role in the selective dissolution of one mineral over another, or

RI
even cathodic protection is possible, causing a less noble mineral to dissolve in favour of
another. It is also possible that poorly soluble minerals precipitate in the particle pores as

SC
their solubility is exceeded; this is especially pertinent to iron hydroxides and gypsum.

NU
Advantages of heap leaching – an economic analysis

The key advantage of heap leaching lies in the fact that it is the only economic technology
MA
available to extract value metals from low-grade ores. The economic advantage arises out of
the fact that some cost-intensive process steps necessary in the treatment of high-grade ores
can be avoided, primarily comminution through milling to make the ore amenable to
subsequent flotation (copper sulphides) or direct leaching (copper oxides, gold ores). Milling
ED

cost increases proportionally to the tonnage of ore milled, whereas potential revenue declines
with metal grade in the ore. This will result in a cut-off grade at which treating the ore using
conventional technology is no longer viable (often cited to be 0.5% Cu for copper sulphides
PT

and 5ppm for Au ores).

Heap leaching operates on coarsely crushed ores and hence completely avoids costs for
CE

milling and concentration. This comes at the cost of substantially longer extraction times,
especially for sulphide leaching, and also of lower overall recoveries of the value metal. The
economic penalty taken for the slow extraction relates to the long time between when the
AC

mining cost is incurred and when revenue through the sale of metal is generated. This ‘delay’
adversely affects the cash-flow of the operation, especially in the start-up of the operation.

This is illustrated in a recent study by Dlamini (2015) where the economics of installing a
green-field heap bioleach process with metal refinery were compared to those of installing a
concentrator and concentrator combined with refinery in the context of the Gamsberg
sphalerite ore in the Northern Cape Province in South Africa. At an average zinc grade of
7.4%, the ore is of intermediate grade for a zinc ore. The following options were compared:

Option A: Heap-bioleaching recovers 73% of zinc in 300 days (conservatively estimated


from data by Ghorbani et al., 2012), followed by zinc extraction via the Hydrozinc™ process
(Lizama et al., 2003) to produce Special High Grade (SHG) cathode on site for export. 96.5%
zinc is recovered as cathode, with the rest being lost in process residues.

Option B: A conventional mill-float concentrator (Davenport, 2002) set-up is used to produce


a sphalerite concentrate for export and refining in Europe using a standard TC/RC (treatment

8
ACCEPTED MANUSCRIPT

charges and refining charges) contract for this type of material. A total zinc recovery of 86%
is assumed in this process.

Option C: The ore is concentrated by milling and flotation as before, but leached and refined
onsite through an atmospheric tank leach, based on the Outotec process (Lahtinen et al.,
2008, Fillipou, 2004), followed by the same Hydrozinc™ refining route as discussed for

T
Option A. A leach efficiency of 98% was assumed and a zinc recovery of 96.5% to cathode,

P
as in Option A.

RI
The study entailed detailed design of each process to handle 3.4 Mt per annum of the zinc ore
to produce either SHG zinc cathode (Routes A and C) or zinc concentrate for external toll

SC
refining. Capital and operating costs were estimated from each of the designs and a complete
cash-flow analysis was conducted for each option. Given the operation is based in South
Africa, all cost analyses were done in South African Rand (1 ZAR = 0.094 USD in mid 2014

NU
when the study was done). Details of this analysis go beyond the scope of this paper and can
be found in Dlamini (2015).
MA
The key findings of this study are summarised in Table 2. The zinc production of Option A
employing heap leaching is by far the lowest due to lower recoveries associated with heap
leaching (within the time frame considered here – higher extractions would be possible with
longer leach periods) – this is offset, however, by considerably lower capital investment and
ED

operating cost when compared to Option C (no need to build and operate a concentrator).
PT

Table 2: Key economic indicators of the comparative study of 3 process options for the
treatment of the Gamsberg zinc ore (Dlamini, 2015)
CE

Option A Option B Option C


Parameter (Heap leach & (Concentrates) (Concentrate
AC

refining) & refining)


Zinc production (kt/a) 176 216 204
TCI (ZAR, Billion) 12.3 7.55 14.9
PBP (Years) 2.73 3.13 3.21
NPV (ZAR, Billion) 1.71 1.17 0.82
IRR (%) 14.6 14.8 13.6
PVR 1.16 1.17 1.06
Nomenclature: TCI – total capital invested; PBP – pay-back period; NPV – net
present value; IRR – internal rate of return; PVR – price value ratio

Capital investment in a concentrator is significantly less than either of the leach-and-refine


options while zinc production is high (albeit in the unrefined form), but this is offset by
significantly lower earnings on a typical refining contract as compared to the much better
price for SHG zinc cathode.

9
ACCEPTED MANUSCRIPT

As a result, the low-efficiency heap leaching option still presents the highest Net Present
Value (NPV) of the three options compared. Internal Rate of Return (IRR) is similar for all
three options, but it should be borne in mind that Option A (heap leaching and refining)
presents a so far untested technology for zinc, for which an investor would usually look at
IRRs >20% to offset the higher risk. This aspect is perhaps a key reason why heap leaching
has not yet found wider application. A Price-Value Ratio (PVR) of unity indicates a break-

PT
even investment. Although all three options offer a PVR above 1, the departures are
marginal, as is typical of primary resource technology – another reason why the industry is

RI
traditionally seen as risk-averse.

A sensitivity study has investigated, amongst other factors, the effect of changing the zinc

SC
price and working capital on NPV (Figure 3). As the zinc price increases, the difference
between the NPVs for Option A (heap leaching and refining) and Option C (concentrating,
leaching and refining) decreases. This is due to the lower annual zinc production in Option A;

NU
hence the zinc price influences the profitability of each process significantly. Conversely, the
difference between Option B and the other options increases. This is due to the price
participation charge found in a typical zinc concentrate refining contract – as the zinc price
MA
increases, the treatment charge is also escalated.

Working capital refers to the capital that needs to be invested to start up, operate the plant and
maintain a stock-pile until revenue is being received. For heap leaching this is considerable,
D

as the long leach time necessary to realise the full value requires a considerable ore inventory
TE

to be under leach – the mining charges for this inventory can only be recovered some months
down the line, as opposed to mere weeks for the other options. For the present study it was
found that an increase in working capital by 30% would result in the NPV of Option A falling
EP

below that of Option B (concentrator). In practice this would mean if the time to recover 73%
of zinc value from the heap increased from 300 to 400 days, the heap leach option would no
C

longer be the most profitable in this comparison.


AC

10
ACCEPTED MANUSCRIPT

Figure 3a: Effect of variation of Zinc price on the NPV of the various process options
considered for the treatment of the Gamsberg zinc ore

PT
RI
SC
NU
MA
D
TE

Figure 3b: Effect of variation of working capital (WC) on the NPV of the various process
EP

options considered for the treatment of the Gamsberg zinc ore


C

As was discussed, the rate of leaching from heaps is critically determined by a number of
AC

factors, including heap permeability, and it is generally observed that leach rates measured in
laboratory columns are rarely matched by those in operational heaps, primarily because of
poorer permeability. Hence a 30% increase in the leach time is not at all unrealistic.

The economic study has shown that using heap leaching over conventional processes can be a
feasible and profitable option, but that such a comparison is highly case specific and good
data is needed to make a meaningful decision in favour of heap leaching.

Heap Leaching Challenges

Heap permeability

As the foregoing analysis has illustrated, the slow rate of recovery from heaps is a major
drawback of this technology in that it requires substantial inventories of ore under leach. A
typical generic copper heap leach curve versus time is shown in Figure 4 – this suggest that

11
ACCEPTED MANUSCRIPT

60-70% extraction can be achieved quite rapidly (accounting for an adaptation delay in
bioleaching), but beyond that the extraction curve begins to slow considerably. The
multitude of processes taking place in a heap contribute to this slowing, but diffusion through
stagnant zones in the ore bed and within particle pores are perhaps the most critical. The
latter can be countered by alternative crushing technology (e.g. HPGR which creates more

T
pores) and crushing to finer sizes, but this is likely to adversely affect the former with more
fines, corresponding to overall lower heap permeability. Further, the length of time over

P
which an ore is kept under leach appears to decrease heap permeability due to decrepitation

RI
of ore particles in the aggressive leach environment and precipitation of leached gangue
minerals such as magnesium, calcium and aluminium sulphates as well as jarosites, clogging

SC
pore spaces (Ahonen and Tuovinen, 1995).

NU
MA
ED
PT
CE
AC

Figure 4: A typical Cu sulphide heap extraction curve (based on data by Garcia et al., 1999)

In some cases the reclamation and re-stacking of ore during heap leaching is practiced to
allow exposure and leaching of unleached particles previously isolated from solution flow.
On-off rinse patterns aim to achieve a similar effect. Building shallower heaps (as little as one
meter have been suggested; Paul and Johnson, 1975) in general also seems to support more
rapid leaching, but at constant mine production this would require a substantially larger
footprint for the heap, the preparation of which might not be economical. Bioleaching of
sulphide minerals enables rapid leaching at optimal temperatures through the exothermic
oxidation reaction, but achieving such temperatures in heaps stacked in cold climates can take
a considerable amount of time during which leach acid may preferentially react with gangue,
resulting in inefficient operation and premature heap clogging. Pre-heating a cold ore could
significantly reduce this delay at relatively small cost, but this is not industrial practice at this
stage.

12
ACCEPTED MANUSCRIPT

Limitations by commodity

A further key drawback of heap leaching relates to the relatively low concentrations of value
metal in the PLS, requiring relatively more elaborate technology or larger scale of operation
to recover it as compared to concentrate leaching processes. The absence of effective
technologies to recover certain metals selectively from low-grade solution has prevented heap

T
leach technology from being considered for their recovery (John, 2011) – these include zinc

P
and nickel. Although the Hydrozinc™ process in the case study presented above revolves

RI
around an SX process for zinc, it needs to be operated at a relatively high pH, requiring
neutralisation of the acidic PLS from the heap, and iron control remains problematic. The

SC
need for lime or limestone as a neutralisation agent significantly influences the operating
costs of the process. Similar problems exist in the extraction of nickel from mixed leach
liquors. Selective ion exchange resins remain an active field of investigation, and this is the

NU
technology of choice for uranium heap leaching. Their great advantage is the ability to
recover metals even at extremely low solution concentrations. Speciality resins are expensive,
however, and their application remains mostly niche.
MA
Environmental concerns

A final drawback of heap leach technology lies in a lack of environmental control over the
process. The flows of potentially toxic chemical solutions (containing acid or cyanide and
ED

dissolved metals) through the large scale operation present a certain risk for leakage into the
environment – through cracks in the liners underneath heaps and through open channels for
solution collection and ponds for storage. Water losses through evaporation and from wind-
PT

blown irrigation sprays can be considerable. All these aspects require careful management of
heap operation, but the risk of a leak in the heap liner remaining undetected is considerable
CE

and hence double-liner systems are becoming more commonplace.

A further concern is directed at the fate of spent heaps. Two principal approaches exist –
AC

removal of the spent ore from the leach pad and disposal in a waste pile, or stacking of fresh
material on top of the spent layer as a new lift. Both methods have advantages and
disadvantages. A spent heap is rarely completely exhausted, given the generally poor
extraction discussed previously. Hence leaching will continue at very slow rates from the
spent material, regardless of where it is deposited. This could pose problems in an un-lined or
unmonitored waste pile, especially in areas of high rainfall. Multiple-lift heaps, on the other
hand, intend to harness this slow leaching by contacting leach liquor from the upper, fresh
pile with the spent lower pile allowing additional value to be released. Poor permeability of
an old heap can interfere, however, with efficient leaching from the fresh one, and thus
effectively cancel the gains made by not having to move the spent heap.

A criticism is often levelled at heaps in that they require large solution inventories which are
permanently held up in the bed, even after the heap is spent. While this is correct, compared
to conventional processing, heaps are actually a lesser sink to water: Tailings dams of finely
ground material usually retain in the order of 30% moisture, whereas in heaps it is only
around 10% due to the larger, unsaturated pore spaces present. In direct comparison

13
ACCEPTED MANUSCRIPT

therefore, heap processing appears less water intensive than conventional processing, but the
author is not aware of such a study having been conducted systematically.

Future directions of heap leaching

T
Heap leaching technology is well established for leaching from low-grade copper oxide and

P
sulphide ores as well as for gold leaching. A few uranium plants are in operation, but these

RI
are relatively uncommon. Although the technology has been explored for zinc at the
demonstration scale (on which the economic study above is based), there are no existing

SC
operations that employ this process. Heap leaching is also employed in some operations for
the dissolution of nitrates from caliche minerals (Galvez et al., 2012).

NU
However there have been a significant number of process innovations looking at heap
leaching using novel reagents, such as the ammonia/thiosulphate system for gold (Grosse et
al., 2003), ammonia for copper oxides (AmmLeach; Welham, 2007 ), chloride for Cu-
MA
sulphides (Aroca et al., 2012) and targeted thermophile bio-leaching for primary copper
sulphides (Dew et al. , 2011). Different commodities considered are zinc (as discussed
above), Ni from low-grade laterites (Agatzini-Leonardou et al., 2009; Readett and Fox,
2010), bio-leaching of poly-metallic shales (Saari and Riekkola-Vanhanen, 2011), rare earths
ED

from their relevant ores, as well as copper, zinc, tin, lead etc. from electronic scrap (by
bioleaching and chemical leaching; Ilyas et al., 2013)
PT

Re-leaching of tailings and residue materials, either to liberate additional value left behind
due to incomplete extraction (especially in the context of historic gold tailings) or to extract
other values not originally targeted in the primary extraction process (for example, uranium
CE

in gold tailings, rare earth elements in titanium sand residues, zinc in steel making furnace
dusts, copper in smelter slags), are other applications where heap leaching could be employed
for the economic recovery of these values. The Geocoat™ technology (Harvey and Bath,
AC

2003), which envisages coating concentrates onto coarse support rock, provides an
inexpensive route to make fine material ‘heap-leachable’ in this context.

However, for any of these new directions to be successful, some of the key drawbacks of
heap leaching discussed above need to be addressed through further research and careful
engineering planning. In this sense heap and ore permeability need to be managed through
ore crush size, heap stacking and solution application management, selective extraction of
target metals from low-tenor solution, and environmentally safe operation and disposal of
spent ore. Process improvements need to be evaluated on an economic – and ideally also
environmental – basis against competing technologies. Implementation of any heap leach
process needs to follow the route from lab-scale evaluation, pilot and demonstration scale
trials, before embarking on a fully commercial operation, as would be good engineering
practice.

Conclusions
14
ACCEPTED MANUSCRIPT

This brief overview has illustrated that heap leach technology can be competitive against
‘conventional’ extractive metallurgy, especially in the context of low-grade ores.
Considerable potential exists for its use for secondary resource materials. However, heap
leaching needs to be appreciated as a relatively complex reactor technology and operated
with this understanding to harness maximum value. The large physical scale of a heap leach

T
operation makes it difficult to develop the technology exclusively in a laboratory, and a close
collaboration between a mining operation and an R&D organisation is required at a relatively

P
early stage of development. Given the relative risk averseness of the mining industry, this

RI
aspect has perhaps prevented a broader uptake of the technology so far, although interest
remains considerable.

SC
Acknowledgement
NU
This paper represents the written version of a keynote address delivered to IC-LGO 2015 in
MA
Jamshedpur, India. Travel to this conference was generously sponsored by the National
Research Foundation (NRF) of South Africa through their KIC Grant (KIC14081290409 ).
ED

References
PT

Agatzini-Leonardou, S., Tsakiridis, P.E., Oustadakis, P., Karidakis, T., Katsiapi, A., 2009.
Hydrometallurgical process for the separation and recovery of nickel from sulphate heap
leach liquor of nickeliferous laterite ores, Minerals Engineering 22, 1181-1192.
CE

Ahonen, L., Tuovinen, O.H., 1995. Bacterial leaching of complex sulphide ore samples in
bench-scale column reactors. Hydrometallurgy 37, 1-21.
AC

Aroca, F., Backit, A., and Jacob, J., 2012. CuproChlor®, a hydrometallurgical technology for
mineral sulphides leaching, Proceedings of the 4th International Seminar on Process
Hydrometallurgy, Editors, Casas, J.M., Frías, S.,C., Ciminelli, V.S.T., Montes-Atenas, G.,
Stubina, N., 11-13 July 2012, Santiago, Chile, 96-108.

Bartlett, R.W., 1998. Solution mining, Leaching and fluid recovery of materials, 2nd Edition,
Gordon and Breach Science Publishers, ISBN 90-5699-633-9.

Davenport, W.G., King, M., Schwarz, M. & Biswas, A.K. 2002. Concentrating Copper Ores.
In Extractive Metallurgy of Copper. 4th ed. The Boulevard, Langford Lane Kidlington,
Oxford OX5 IGB, UK: Elsevier Science Ltd. 31-54.

Demergasso, C., Galleguillos, P., Escudero, L., Zepeda, V., Castillo, D., and Casamayor, E.
2005. Molecular characterization of microbial populations in low-grade copper ore
bioleaching test heap, Hydrometallurgy 80, 241-253.

15
ACCEPTED MANUSCRIPT

Dew, D.W., Rautenbach, G.F., van Hille, R.P., Davis-Belmar, C.S., Harvey, I.J., Truelove,
J.S., 2011. High temperature heap leaching of chalcopyrite: method of evaluation and process
model validation. In International Conference: Percolation leaching: The status globally and
in Southern Africa. Misty Hills, Muldersdrift, South Africa, 8-9 Nov; SAIMM Symposium
Series S69, ISBN 978-1-920410-24-7, 201-220.

T
Dixon, D.G., 2000. Analysis of heat conservation during copper sulphide heap leaching,

P
Hydrometallurgy 58, 27-41.

RI
Dlamini, Z., 2015. Techno-economical comparison of three process routes for the treatment
of Gamsberg zinc ore. MSc thesis, University of Cape Town.

SC
Filippou, D. 2004. Innovative Hydrometallurgical Processes for the Primary Processing of
Zinc. Mineral Processing and Extractive Metallurgy Review. 25(3): 205-252.

NU
Gálvez, E.D., Moreno, L., Mellado, M. E., Ordóñez, J. L., Cisternas, L.A., 2012. Heap
leaching of caliche minerals: Phenomenological and analytical models-Some comparisons,
MA
Minerals Engineering 33, 46-53.

Garcia, C., Arias, H., Campos, J., Roco, J., San Martin, O., Whittaker, J., 1999. Bioleaching
at Zaldivar. Biomine ’99 and Water Management in Metallurgical Operations – Conference
ED

Proceedings. 1999 August 23-24; Perth, Australia; Glenside, Australia: Australian Mineral
Foundation Inc., 72-81.
PT

Gericke, M., Seyedbagheri, A., Neale, J., van Staden, P.J., 2011. Advancements in the
approach to research and design of heap bioleaching processes, the 19th International
Biohydrometallurgy Symposium (IBS2011)., September 18-22, 2011, Changsha-China,
CE

Volume I, 698-705.

Ghorbani, Y., Petersen, J., Harrison, S.T.L., Tupikina, O.V., Becker, M., Mainza, A.N. &
AC

Franzidis, J. 2012. An experimental study of the long-term bioleaching of large sphalerite ore
particles in a circulating fluid fixed-bed reactor. Hydrometallurgy. 129–130: 161-171.

Ghorbani, Y., Petersen, J., Becker, M., Mainza, A.N., Franzidis, J-P., 2013. Investigation and
Modelling of the Progression of Zinc Leaching from Large Sphalerite Ore Particles,
Hydrometallurgy (131-132), 8-23.

Govender, E., Bryan, C.G., Harrison, S.T.L., 2013. Quantification of growth and colonisation
of low-grade sulphidic ores by acidophilic chemoautotrophs using a novel experimental
system, Minerals Engineering 48,108-115.

Grosse, A.C., Dicinoski, G.W., Shaw, M.J., Haddad, P.R, 2003. leaching and recovery of
gold using ammoniacal thiosulphate leach liquors (a review). Hydrometallurgy 69 (1-3), 1-
21.

Harvey, T.J., Bath, M.D., 2003. Development of the first commercial GEOCOAT® heap
leach for refractory gold at the Agnes Mine, Barberton South Africa. In: Tsezos M,

16
ACCEPTED MANUSCRIPT

Hatzikioseyian A, Remoundaki E (eds) Biohydrometallurgy: a sustainable technology in


evolution (part I). National Technical University of Athens, 387-398

Ilyas, S., Lee, J-C., Chi, R-A.., 2013. Bioleaching of metals from electronic scrap and its
potential for commercial exploitation, Hydrometallurgy (131-132), 138-143.

T
John, L. 2011. The Art of Heap Leaching - The Fundamentals. In International Conference:

P
Percolation leaching: The status globally and in Southern Africa. Misty Hills, Muldersdrift,
South Africa, 8-9 Nov; SAIMM Symposium Series S69, ISBN 978-1-920410-24-7, 17-42.

RI
Kappes, D.W., 2002. Precious metal heap leach design and practice, in Proceedings - Mineral

SC
Processing Plant Design, Practice and Control, SME, Colorado, USA, ISBN 0-87335-223-8,
Vol.2, 1606-1630.

NU
Lahtinen, M., Svens, K., Haakana, T. & Lehtinen, L. 2008. Zinc plant expansion by Outotec
direct leaching process. 47th Annual Conference of Metallurgists of CIM, Winnipeg,
Manitoba Canada, Zinc and Lead Metallurgy. 167.
MA
Lizama, H.M., Harlamovs, J.R., Belanger, S. & Brienne, S.H. 2003. The Teck Cominco
Hydrozinc™ process. In Hydrometallurgy 2003 - Firth International Conference in Honour of
Professor Ian M. Ritchie - Volume 2: Electrometallurgy and Environmental
ED

Hydrometallurgy. C.A. Young, and others, Eds. volume 2 ed. TMS (The Minerals, Metals &
Material Society). 1503 -1516.

Logan, T.C., Seal, T. and Brierley, J.A., 2007. Whole-Ore Heap Biooxidation of Sulfidic
PT

Gold-Bearing Ores. In Biomining, D.E. Rawlings, D.B. Johnson (Eds.), Springer Verlag,
Berlin, ISBN 978-3-540-34909-9, pp 113-138.
CE

Ojumu, T.V., Petersen J. and Hansford, G.S., 2008. The effect of dissolved cations on
microbial ferrous-iron oxidation by Leptospirillum ferriphilum in continuous culture.
Hydrometallurgy 94, 69-76.
AC

Paul H., Johnson, P.H. "Thin layer leaching method" U.S. Patent 05,562962. 28 March 1975.

Petersen, J. and Dixon, D.G., 2007a. Modeling and Optimisation of Heap Bioleach Processes.
In Biomining, D.E. Rawlings, D.B. Johnson (Eds.), Springer Verlag, Berlin, ISBN 978-3-
540-34909-9, pp 153-176.

Petersen J. and Dixon, D.G, 2007b. Principles, Mechanisms and Dynamics of Chalcocite
Heap Bioleaching. In Microbial Processing of Metal Sulfides, E.Donati, W. Sand (eds.),
Springer Verlag, Berlin, ISBN 978-1-4020-5588-1, pp 193-218.

Petersen, J., Minnaar, S.H., du Plessis, C.A, 2010. Carbon dioxide and oxygen consumption
during the bioleaching of a copper ore in a large isothermal column, Hydrometallurgy 104,
356-362.

17
ACCEPTED MANUSCRIPT

Readett and Fox, 2010. Commercialization of Nickel Heap leaching at Murrin Murrin
operations, Proceedings of XXV International Mineral Processing Congress (IMPC),
Brisbane, Queensland, Australia, 6-10 September, 3611-3616.

Saari, P., Riekkola-Vanhanen, M., 2011. Talvivaara bio-heap leaching process, Percolation
Leaching: Southern African Institute of Mining and Metallurgy Percolation Leaching: The

T
status globally and in Southern Africa 2011, 53-66.

P
Sand W., Gehrke T., Josza P-G., Schippers A., 2001. (Bio)chemistry of bacterial leaching –

RI
direct vs indirect bioleaching. Hydrometallurgy 51:115-129.

SC
Scheffel, R., 2002. Copper Heap Leach Design and Practice, Mineral processing plant design,
practice, and control: proceedings, vol. 2, Chapter 12, edited by Andrew L. Mular, Doug N.
Halbe, Derek J. Barratt, 1571-1605.

NU
Sánchez-Chacón, A.E, Lapidus, G.T., 1997. Model for heap leaching of gold ores by
cyanidation, Hydrometallurgy 44, (1-2), 1-20
MA
Watling, H.R., 2006. The bioleaching of sulphide minerals with emphasis on copper
sulphides-A review, Hydrometallurgy 84, 81-102.

Welham, N.J. "Method for ammoniacal leaching" A.U. Patent 07,903815. 13 July 2007.
ED
PT
CE
AC

18
ACCEPTED MANUSCRIPT

Highlights
- Heap leach technology can be more profitable when compared to conventional process
routes
- Successful heap leach operation requires a thorough understanding of the underlying
principles for optimal operation
- Heap leaching technology offers a promising route for the processing of many different low-

T
grade ores and urban ore type materials

P
RI
SC
NU
MA
ED
PT
CE
AC

19

You might also like