Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Hydrometallurgy 215 (2023) 105997

Contents lists available at ScienceDirect

Hydrometallurgy
journal homepage: www.elsevier.com/locate/hydromet

In-situ leaching of copper from spent heaps


Hector M. Lizama
Teck Metals Ltd., Technical Services, Trail, P.O. Box 2000, Trail, BC V1R 4S4, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: In-situ leaching assisted by geophysical techniques were used to exploit previously leached and long-abandoned
In-situ leaching copper sulphide heaps. The technology was first piloted for five months at the 59 kt scale with up to three in­
Heap leaching jection wells. This was followed by a 14-month industrial pilot at the 1.2 Mt. scale with 97 injection wells. In-situ
Geophysics
leaching was then fully implemented, drilling more than one thousand injection wells on 16 abandoned heaps
Electro resistivity tomography
over a period of almost three years. Geophysical surveys were used to visualize the movement of the injected
leaching solutions in-situ. Geophysical imaging was also used to avoid flooding and prevent geotechnical
instability. The imaging indicated the injection wells had a radius of influence of 11 to 13 m, and that well
injection flows should not exceed 1.5 L/s. Oxidative conditions were facilitated in-situ by encouraging air entry
from the heap drainage system and alternating periods of injection with rest periods that allowed air entry
through the injection wells. This work introduces some concepts and terms related to in-situ leaching operations.

1. Introduction decreases with time. At the same time, the already-mentioned dimin­
ishing copper extractions towards the end of the leach cycle compel the
In copper heap leaching, ore is stacked in leach pads and irrigated operation to terminate leaching well before all the available copper is
with acidic solution. As the leaching solution percolates down through recovered to maximize the economic return of the heap. These two
the ore bed, chemical reactions consume acid and release copper ions factors combine to create the situation where the lower-grade ores being
from the contained copper minerals. In the case where the contained processed at the end of mine life may not be too much different in value
copper minerals are secondary sulphides such as chalcocite (Cu2S) and from the abandoned residues generated in earlier times when ore grades
covellite (CuS), the stacked ore is aerated from the bottom as well as were much higher.
irrigated from the top. This is because the copper ions are liberated from This study pertains to the abandoned residues of the Quebrada
those sulphide minerals by oxidative reactions mediated by bacteria. Blanca operation in northern Chile. Originally, starting in 1994, crushed
Leaching of copper sulphide ore in heaps typically takes several months, and agglomerated ore was leached in 6 m lifts stacked one on top of
with diminishing amounts of copper being extracted over the passage of another, in static pad configuration. In 1998, after the second lift, the
time. On reaching a target extraction value, a heap is taken off-line, operation changed to a dynamic pad configuration, leaching the ore in a
allowed to drain down, and then decommissioned. Decommissioning single lift, and removing the residues to make way for fresh ore (Yáñez
can entail removing the material from the pad to a waste dump or using et al., 2007). Those first two static lifts were left in place and served as a
the decommissioned heap as a platform for the stacking of fresh ore as foundation for the dynamic lifts above. The residues from the two buried
part of a new heap for subsequent leaching. In either case, the residues, lifts of leached residues total 16.8 million tonnes and cover an area of
with their remaining copper value, are abandoned. Abandoning the approximately 710,000 m2. In 2016, an opportunity was identified to
residues by removing them is referred to as a dynamic pad configuration recover the residual copper in the buried lifts at low cost by in-situ
and abandoning them by placing fresh ore on top is referred to as a static leaching techniques aided by wireline geophysics (Sinclair and
pad configuration. The individual layers in a static heap are referred to Thompson, 2015). That approach was tested at Asarco's Ray Mine
as lifts. (Williams et al., 2015) and was applied successfully at KGHM's Carlota
As with any commercial venture, heap leach operations maximize operations (Rucker, 2015). However, those applications involved re-
their economic returns by processing the most valuable ores first and leaching of acid-soluble copper oxide minerals rather than secondary
deferring the lesser valuable ores until later. This results in ores with the sulphide minerals. The bacterially assisted oxidative reactions involved
highest copper grades being mined first, so that the copper grade in leaching copper sulphides require air or dissolved ferric ions. The in-

E-mail address: hector.lizama@teck.com.

https://doi.org/10.1016/j.hydromet.2022.105997
Received 23 August 2022; Received in revised form 7 November 2022; Accepted 9 November 2022
Available online 12 November 2022
0304-386X/© 2022 Elsevier B.V. All rights reserved.
H.M. Lizama Hydrometallurgy 215 (2023) 105997

situ leaching techniques described in this study complied with that Table 1
requirement (Lizama and Arrué, 2020). Description of copper resource in the buried lifts (A and B) of the Quebrada
Blanca heap.
2. Materials and methods Property Value

Number of heaps targeted 16


2.1. Site description Area 710,800 m2
Resource tonnage 16,780,000 t
The Quebrada Blanca heap leach pad area, shown in Fig. 1, was Grade (sequential Cu) 0.22% (average)
Recoverable metal (sequential Cu) 36,900 t
divided into North and South Pads. The two pads comprised 23 heaps, of
which 14 were in the North Pad and nine in the South Pad. Of these, 16 Note that the averaged Cu grade was highly variable, according to location.
were targeted for in-situ leaching: eight in the North Pad and eight in the
South Pad. The pads were also divided into lifts. The first two lifts were Arrué, 2020).
inactive and were part of the old static pad configuration. On top of
these two abandoned lifts was a compacted impermeable layer that
2.3. Injection system
served as a base for a single dynamic lift on top. Underneath the
bottommost static lift was the heap primary liner. The static lifts
The injection wells were arranged in either a square or hexagonal
comprised 16.8 million tonnes and held an estimated 37 thousand
injection pattern. Schematics of both injection patterns are shown in
tonnes of sequential copper. Sequential copper is defined as copper that
Fig. 3. Within a given injection pattern, the wells were spaced to ensure
can be solubilized by digestion of finely-ground ore with sulphuric acid
maximum sweep efficiency; sweep efficiency defined as the degree of
and cyanide (Parkison and Bhappu, 1995). Note that sequential copper
wetting within the production zone during the periods of continuous
is understood to include copper oxides and secondary sulphides of
injection. The well spacing was determined by the radius of influence,
copper; primary sulphides of copper such as chalcopyrite are not
which was the extension of an injection plume from the wellbore into
detected by the sequential digestion. The potential resource is detailed
the production zone. The extensions of injection plumes were measured
in Table 1.
by geophysical surveying.
A typical injection well included a wellhead and a downhole pipe
2.2. Technology description string, shown in Figs. 4 and 5, respectively. The wellheads were
configured to be able to introduce a baler (portable grab sampler) or
The heaps were surveyed with topographical and geophysical tech­ level sensor downhole. This was achieved by the injection pipe string
niques to define the boundaries of a production zone. Injection wells being connected by a flange to a tee junction that ended in an air vent
were drilled within the production zone, to allow the introduction of valve. When solution flow was interrupted and the vent valve was
leaching solution, comprising of raffinate solution from the solvent opened, a continuous duct was established from atmosphere to the well
extraction (SX) plant. The production zone was operated with periods of bottom as the solution drained from the well. The side branch of the tee
continuous injection alternating with rest periods where no solution was junction was connected to the raffinate inlet line, which included a
injected. During injection, leaching solution seeped into the pore spaces pressure gauge, a flowmeter, and a feed valve. The pipe string consisted
of the production zone, forming a plume. Within the plume, the leaching of a 4-in. diameter Schedule 40 PVC pipe that extended down to 2 m
solution reacted with the copper-bearing minerals and leached the above the heap's primary liner. A slotted pipe section extending within
copper. During rest periods, air was allowed to enter the wellbores and the boundaries of the production zone allowed introduction of raffinate
diffuse into the production zone below. Pregnant leach solution (PLS) solution. The annular space between the pipe string and the production
was drained from the bottom of the production zone by way of the zone was filled with pea gravel. Towards the top of the well there were
abandoned heap's original drainage system. Fig. 2 illustrates the basic two seals to prevent seepage up from the production zone. The first was
concept of in-situ leaching of spent heaps. a 1 m thick layer of bentonite pellets. The second, on top of the first, was
The concept took advantage of the spent heap's original drainage a cement/bentonite seal that extended all the way to the surface.
system to encourage air entry to the production zone from below. This
was done by managing solution flows in a way that maintained the
drainage lines only partially filled, allowing the flow of air (Lizama and 2.4. Geophysics

Electro resistivity tomography (ERT) wireline geophysical surveys


were carried out to delineate production zones and to monitor sweep
efficiency. The resistivity measurements were collected and analyzed
using the RES2DINV algorithm, with an electrode spacing of either 1 m
or 2 m, depending on the profile used. These surveys were instrumental
in locating the heap HDPE primary liner prior to drilling any wells.
During operation of the injection system, the ERT surveys were used to
visualize the changes in resistivity to infer changes in moisture content.
That gave an indication of solution movement and plume extensions
about the injection wells (Lang et al., 2013).

3. Results

3.1. Initial piloting

As a proof-of-concept test, in-situ leaching was piloted on a small


area at the south end of Heap No. 9. This test involved a production zone
roughly trapezoidal in shape, covering an area of 3220 m2 and having an
Fig. 1. Map of the Quebrada Blanca heap leach pad area. The numbered heaps average depth of 9.7 m (Lizama et al., 2018). Overlaid material from the
shown are dynamic lifts, overlaying static lifts below. dynamic pad had been removed so that the production zone was exposed

2
H.M. Lizama Hydrometallurgy 215 (2023) 105997

Fig. 2. Concept of in-situ leaching of copper from a spent heap. Leaching solution is introduced to a production zone by way of injection wells. The injected leach
solutions form a plume wherein leaching occurs. Solution within the plume drains downwards and is collected by the spent heap's drainage system.

Fig. 3. Injection patterns: square and hexagonal. Note how the well spacing is
Fig. 4. Typical wellhead assembly for an injector well used for in-situ leaching
the same in both patterns. However, each injector in a square pattern has four
of a spent heap.
nearest neighbours, while each injector in a hexagonal pattern has six near­
est neighbours.
drainage system of the original static leach pad.
to atmosphere. On the north and east the test site was bounded by The boreholes drilled for well installation also served to obtain ma­
stacked agglomerate on the dynamic pad; on the south and west, by a terial samples from the production zone prior to leaching. The samples
roadway; and on the bottom by the heap leach pad primary liner. Within were used for analyses and initial moisture contents. According to the
the test site, there were 21 injection and monitoring wells. In the pe­ drill sample results, the production zone had an overall grade of 0.43%
riphery of the test site, there were geotechnical wells fitted with pie­ sequential copper. Mineralogical analyses indicated that >90% of the
zometers to measure the phreatic levels. This was to avoid sequential copper in the pilot test production zone was present as
compromising the geotechnical stability of the static leach pad material. covellite (CuS). A second drilling campaign at the end of the test pro­
Fig. 6 shows a schematic of the test site and well locations. PLS from the vided post-test samples for analyses and residual moisture. Both cam­
production zone exited at two discharge points; one located 100 m north paigns utilized a six-inch rotary vibratory (SONIC) drill (Barrow, 1994).
of the test area and the other 300 m north. Both were part of the During piloting, the solution flows and line pressures were measured
approximately every six hours. Solution sampling and phreatic level

3
H.M. Lizama Hydrometallurgy 215 (2023) 105997

Fig. 7. Injection flows and copper production from the initial pilot test. Copper
production was calculated by multiplying the daily injection flow times the
change in copper concentration.

Fig. 8.
Upon termination of the pilot test, the production zone was allowed
to drain naturally, with continuous measurement of effluent flows. A
Fig. 5. Typical wellbore and pipe string assembly for an injector well used for solution balance was then constructed by reconciling: the moisture
in-situ leaching of a spent heap. contents from the pre-test and post-test drilling campaigns, the
measured cumulative influent and effluent volumes, and the drain-down
curves observed at the end of the test. The reconciled solution balance
allowed calculation of the production zone's physical properties before
and during operation, listed in Table 2. The calculations used a 2.6 t/m3
density for the dry rock within the production zone.
The metallurgical balance for the pilot test is summarized in Table 3.
Copper production and acid consumption were calculated by tracking
the daily change in concentration from raffinate to PLS, multiplied by
the daily injection flow. The net acid consumption value of 0.43 kg/kg
Cu considers that for every kg of copper produced, 1.54 kg of sulphuric
acid is returned to the leaching circuit by the SX raffinate solution.
The metallurgical trends observed were deemed indicative of
oxidative leaching of copper by ferric ions and bacterial activity. These
are shown in Fig. 9, which plots the changes in concentration of copper,
total dissolved iron, ferric ion, and acid: ΔCu, ΔFe total, ΔFe3+, and
ΔH+. Those differential values were obtained by subtracting the raffi­
Fig. 6. Map of the pilot test site showing locations of the injection and moni­
toring within the production zone, and the geotechnical wells that bounded it. nate solution concentrations from the corresponding PLS concentra­
tions. Although not shown in the figure, the injected raffinate solution
had a pH that ranged from 1.6 to 2.0 and an oxidation-reduction po­
measurement was done daily. Injection pressure never exceeded 145
tential that ranged from 700 to 750 mV by Ag/AgCl electrode. In turn,
kPa (gauge) and most of the time was approximately 0 kPa (gauge).
the effluent PLS solution had a pH that ranged from 1.7 to 2.2 and an
Injection was tested in continuous mode as well as on-off mode: periods
oxidation-reduction potential that ranged from 640 to 730 mV by Ag/
of injection interrupted with rest periods where the wellhead vent valves
AgCl electrode.
were open to atmosphere. The pilot test started with one injection well
Covellite, the predominant source of sequential copper in the pro­
but then expanded to two and eventually three injection wells operating
duction zone, oxidizes according to the reactions (Petersen and Dixon,
simultaneously. Operating more than three injection wells was observed
2007; Watling, 2006):
to result in flooding of the production zone surface. Likewise, injection
flows higher than 1.5 L/s per well also resulted in flooding. Daily in­ CuS + Fe₂(SO₄)₃→CuSO₄ + 2FeSO₄ + S0 (1)
jection flows and cumulative copper production are shown in Fig. 7.
The pilot test lasted 155 days. Over that time, 30,230 m3 of raffinate 2FeSO₄ + ½O₂ + H₂SO₄→Fe₂(SO₄)₃ + H₂O (2)
solution were injected into a production zone that contained 59,020 t of
dry solids, equivalent to a 0.51 m3/t injection ratio. The injection ratio is Overall : CuS + ½O₂ + H₂SO₄→CuSO₄ + H₂O + S0 (3)
a term analogous to irrigation ratio, a term used in heap leaching
Covellite is oxidized by ferric ion (Eq. 1), generating copper ions and
defined as the total volume of applied leaching solution divided by the
ferrous ions. To maintain the reaction, the ferrous ions must be oxidized
tonnage of ore under leach (Avendaño, 2004). The raffinate solution
back to ferric ions (Eq. 2). Moreover, sufficiently acidic conditions must
injected contained approximately 4 g/L sulphuric acid (pH 1.8) and 1.2
be present to maintain the ferric ions in solution, otherwise they will
g/L ferric ion. In the five months of operation, the pilot test produced 34
precipitate according to the general reaction (Leahy and Schwarz,
t of copper. This was equivalent to a recovery of 13%, based on the pre-
2009):
test resource drilling. Over the course of the test, at least ten ERT surveys
were completed. These surveys indicated that the injection wells had a K+ + 3Fe3+ + 2SO₄2− + 6H₂O⇌KFe₃(SO₄)₂(OH)₆ + 6H+ (4)
radius of influence of approximately 11 m; an example is shown in
Over the course of the pilot test, the change in total dissolved iron

4
H.M. Lizama Hydrometallurgy 215 (2023) 105997

Fig. 8. Plane view ERT profile at 3.4 m depth during injection from a single well. The region of lower resistivity surrounding the injection point relates to higher
moisture content, indicative of the injection plume. The injection plume suggests a radius of influence of 11 m.

Table 2
Physical properties of the production zone, before and during injection.
Physical properties Before injection During operation
a 3
Bulk volume 31,110 m 31,110 m3
Dry solidsa 59,020 t 59,020 t
80% passing size of 6.8 ± 1.5 mm 6.8 ± 1.5 mm
solidsb
Fines in solids (<150 μm) 25% ± 9 25% ± 9
Total voids 8410 m3 (27% by 8410 m3 (27% by
volume) volume)
Moisture content 5060 t (7.9% by weight) 7570 t (11.4% by weight)
Bulk density 2.06 t/m3 2.14 t/m3
Saturation 60% (of voids by volume) 90% (of voids by volume)
Air-filled voids 3350 m3 (40% of voids) 840 m3 (10% of voids)
a
Bulk density was equivalent to 1,90 t/m3, although historical data indicates
an original bulk density of about 1.5 t/m3, prior to the static lifts being com­
pacted upon decommissioning.
b
Historical data indicates an original P₈₀ of 12 mm before leaching.
Fig. 9. Pilot test concentration differences between PLS and raffinate solution
for copper, total iron, ferric ion, and acid. The differences (Δ) were calculated
Table 3 by subtracting the concentration in the raffinate from the corresponding con­
Copper and acid material balance of the initial pilot test. centration in PLS.

Dry solids 59,020 t


acid consumption observed could have been due to reaction with gangue
Test duration 155 d
minerals cannot be discounted. The gangue minerals identified within
Volume of raffinate solution injecteda 30,230 m3
Injection ratio 0.51 m3/t the pilot test production zone are listed in Table 6.
Pre-test sequential copper content 253.8 t (0.43% grade)
Copper leached 33.9 t (13.3% recovery)
Total acid consumption 66.9 t 3.2. Industrial pilot test
Gross acid consumption per tonne dry solid 1.13 kg/t
Gross acid consumption per kg Cu leached 1.97 kg/kg To demonstrate the economic feasibility of in-situ leaching, a much
Net acid consumption per kg Cu leached 0.43 kg/kg
larger industrial pilot test was carried out on an area covering 59,930 m2
a
Injected raffinate solution had, on average, a pH of 1.77, an oxidation- and having an average thickness of 10 m. This area corresponded to
reduction potential of 724 mV, and contained, also on average, 1.65 g/L total heaps No. 1 and No. 2, shown in Fig. 1, and was overlayed by the dy­
Fe and 1.49 g/L Fe3+. namic lift, except for a small strip (6% of the area) at its north end. The
overlaying dynamic lift had an average thickness of 7 m. A layout of the
was small but positive, indicating that no precipitation of ferric ion industrial pilot area is shown in Fig. 10, indicating the locations of the
occurred. If ferric ion precipitation had occurred, the change in total wells. PLS from the production zone exited at two discharge points, both
dissolved iron would have been negative. The change in ferric ion was located at the Southwest toe of Heap No. 1. Both discharge points tied
negative and relatively constant, indicating ferric ion was being reduced into the drainage system of the original static pad.
to ferrous concomitant with copper leaching. The change in sulphuric The industrial pilot site had 120 wells in total: 97 injection wells,
acid was relatively large and negative: opposite to the change in copper. eleven monitoring wells, nine geotechnical wells fitted with piezometers
All this suggested that copper sulphides were reacting with ferric ions to (to measure underground water pressure), and two production wells
produce dissolved copper ions and ferrous ions, with the ferrous ions fitted with downhole pumps. The production wells served to control
likely being oxidized back to ferric ions by bioleaching bacteria through phreatic levels at the west toe of Heap No. 1. The injection wells were
the consumption of oxygen and acid. However, the possibility that the arranged in a hexagonal pattern, where each injector was surrounded by

5
H.M. Lizama Hydrometallurgy 215 (2023) 105997

strategy and well design were effective in confining the plume to the
production zone with a high sweep efficiency. An example is shown in
Fig. 11, which depicts an east-west row of injection wells. Three months
after starting injection of raffinate solution, there was a significant
change in the resistivity within the production zone and this resistivity
change was distributed relatively evenly, indicating a uniform plume.
There had been very little resistivity change in the dynamic lift above
the production zone.
The industrial pilot ran for 457 days. Over that time, 1,696,200 m3of
raffinate solution were injected into a production zone that contained an
estimated 1,234,700 t of dry solids, equivalent to a 1.37 m3/t injection
ratio. The raffinate solution injected contained, on average, 4 g/L sul­
phuric acid (pH 1.7) and 1.3 g/L ferric ion. In the 14 months of opera­
tion, the pilot test produced 950 t of copper. This was equivalent to a
recovery of 30%, based on resource drilling that identified a sequential
copper grade of 0.26%. Over the course of the test four ERT survey were
completed. The daily injection flows and the cumulative copper pro­
duction are shown in Fig. 12. The figure shows how the average total
injection flow improved once the phreatic level issues in Heap No. 1
were addressed. Nevertheless, the injection flows at the individual wells
did not exceed 1.5 L/s.
The net acid consumption in the industrial pilot was 1.51 kg/kg Cu.
Again, the net acid consumption considers the 1.54 kg of sulphuric acid
Fig. 10. Map of the industrial pilot site, showing the injection pattern and
direction of drainage. Also shown, are the dispositions of monitoring,
geotechnical wells, and production wells.

six nearest neighbours with a 26 m spacing. The radius of influence was


anticipated to be 13 m.
The operating philosophy consisted of alternating injection in every
other East-West row of wells. Hence, all injection wells had rest periods
interspersed with active periods, even though the entire test area
remained under continuous injection. During its rest period, an injection
well's vent valve was left open to atmosphere to allow ingress of air to
the production zone. Initially, raffinate solution was injected for seven
days followed by a seven-day rest period. Soon, however, the alternating
injection/rest periods were extended to two weeks on, two weeks off.
Injection flows and wellhead pressures were measured daily. Raffinate
and PLS were sampled daily. A data logger captured phreatic levels at
the geotechnical wells every 12 h.
On commencement of the industrial pilot, there was a steady rise of
the phreatic level at the west toe of Heap No. 1, prompting the restric­
tion of injection flow in that region. Eventually that was mitigated by Fig. 12. Injection flows and copper production from the industrial pilot. Cop­
reinforcing that area with a buttress and sinking two production wells to per production was calculated by multiplying the daily injection flow by the
pump out solution (indicated on the map in Fig. 10). The injection change in copper concentration.

Fig. 11. Change in ERT profile along a row of injection wells running East to West across heaps No. 1 and No. 2. The profile shows how the resistivity changed from
before any solution was injected to three months after start-up of the industrial pilot. Here resistivity is the reciprocal of electrical conductivity.

6
H.M. Lizama Hydrometallurgy 215 (2023) 105997

returned to the leaching circuit for every kg of copper extracted by the


SX circuit. Although the net acid consumption of the industrial pilot was
three times that of the initial pilot test, the acid consumptions were
comparable when based on production zone tonnage. The initial pilot
test had consumed 1.13 kg of acid per tonne of dry solids after injecting
0.51 m3/t of raffinate. In comparison, the industrial pilot had consumed
0.9 kg/t dry solids at the same injection ratio of 0.51 m3/t. Hence, the
discrepancy in net acid consumption can be attributed to the difference
in copper grade of the two production zones: 0.43% in the initial pilot
test and 0.26% in the industrial pilot. The metallurgical balance for the
industrial pilot is summarized in Table 4. As before, copper production
and acid consumption were calculated by multiplying the concentration
changes with the injection flows.
As with the previous trial, the metallurgical trends of the industrial
pilot suggested oxidative leaching of copper. This was apparent from the
metallurgical trends, shown in Fig. 13. First, the ΔFe total trend was
positive, indicating no precipitation of ferric ion. Second, the ΔFe3+ Fig. 13. Industrial pilot test concentration differences between PLS and raffi­
trend was negative, indicating the reduction of ferric ions to ferrous ions. nate solution for copper, total iron, ferric ion, and acid. The differences (Δ)
And finally, the ΔH+ trend was opposite to the ΔCu trend. were calculated by subtracting the concentration in the raffinate from the
The industrial pilot was successful in demonstrating the economic corresponding concentration in PLS.
feasibility of in-situ leaching of the spent heaps. Although the industrial
pilot ran for 14 months, payback for all the incurred costs was obtained
within eleven months. As a result, the decision to proceed with in-situ Table 5
leaching of the other spent heaps was taken well before the industrial Sequence of heaps placed under injection for in-situ leaching.
pilot test was completed. Heap Injection wells Placed on injection

1 42 March 2017
3.3. Full-scale implementation 2 55 March 2017
6 56 November 2017
Implementation of in-situ leaching began well before the industrial 7 99 February 2018
8 87 March 2018
pilot campaign finished. The rapid execution was warranted by the high 9 43 May 2018
copper production from the industrial pilot despite initial restrictions on 10 46 June 2018
injection flow. The first spent heap to start injection did so only six 11 46 August 2018
months into the industrial pilot campaign. Eventually, 16 spent heaps 12 54 September 2018
13 85 December 2018
were leached in-situ, including the two that were part of the industrial
14 89 January 2019
pilot. Table 5 details the sequence in which the spent heaps were 16 83 February 2019
brought online. More than one thousand injection wells were installed 15 86 March 2019
and completed. The industrial pilot's 26 m well spacings in hexagonal 5 58 July 2019
patterns were extended to all the spent heaps. Additionally, individual 3 78 August 2019
17 68 September 2019
well injection flows under 1.5 L/s were also applied.
In-situ leaching of spent heaps was practiced for 34 months at
Quebrada Blanca. That period is counted from the start of the industrial
pilot to the time when injection of raffinate solution ceased due to
operational constraints. Over the almost three years of in-situ leaching,
14 million m3 of raffinate solution were injected into various production
zones, and 7600 t of copper were produced. That was equivalent to a
recovery of 21%, based on the available resource detailed in Table 1.
The daily injection flows and the cumulative copper production are
shown in Fig. 14. The figure shows how the total injection flow increased
as additional spent heaps were incorporated to the system. After 15
months, the number of operating injection wells averaged 200 and never
exceeded 230.

Table 4
Copper and acid material balance of the industrial pilot.
mid-test end of test

Dry solids (t) 1,234,700 1,234,700


Duration (d) 229 457
Volume of raffinate solution injected (m3) 627,800 1,696,200
Fig. 14. Injection flows and copper production from in-situ leaching of spent
Injection ratio (m3/t) 0.51 1.37
heaps at Quebrada Blanca. The injection flows are monthly averages.
Pre-test sequential copper content (%) 0.26 0.26
Pre-test sequential copper content (t) 3210 3210
Copper leached (t) 338 950 4. Discussion
Copper recovery (%) 10.5 29.6
Acid consumed (t) 1108 2901
Gross acid consumption (kg/t dry solids) 0.90 2.34
4.1. Copper recovery and acid consumption
Gross acid consumption (kg/kg Cu leached) 3.28 3.05
Net acid consumption (kg/kg Cu leached) 1.74 1.51 Copper recovery and acid consumption were found to be related to

7
H.M. Lizama Hydrometallurgy 215 (2023) 105997

the injection ratio. This was apparent from the copper recovery trends
plotted in Fig. 15 and the acid consumption trends plotted in Fig. 16.
Both figures plot data from the initial pilot test, the industrial pilot test,
and from heaps No. 6, 7, and 8. Those three heaps were operated and
monitored as a single entity. After the industrial pilot, heaps No. 6, 7,
and 8, were the first ones to be in-situ leached, under the implementa­
tion schedule shown in Table 5. As such, they were subject to close
monitoring, unlike the heaps that came afterward. Injection did not start
at the same time for heaps 6, 7, and 8. However, the effluent for those
three heaps exited at a common point. Hence, they were monitored as if
they were one heap.
The leach curves shown in Fig. 15 were relatively similar despite the
copper grades being different in the three production zones. Sequential
copper contents in the pilot test, industrial pilot, and the combined
heaps No. 6, 7, and 8, were: 0.43%, 0.26%, and 0.22%, respectively.
Hence, copper recovery was not dependent on grade. Likewise, the acid
consumption was also not related to copper grade; the three production
Fig. 16. Acid consumption as a function of the injection ratio for the initial
zones yielded relatively similar acid plots in Fig. 16, despite their pilot test, the industrial pilot test, and the combined heaps 6, 7, and 8. The acid
different copper contents. Both the recovery plots in Fig. 15, and the acid consumption is expressed as per tonne of material within the production zone.
consumption plots in Fig. 16, approximated linear relationships. Indeed,
combining the data from Figs. 15 and 16 revealed a universal gain of
twelve points of recovery for every one kg of acid consumed per tonne of Table 6
production zone material (data not shown). Mineralogical composition of initial pilot test production zone material.
The universal relationship between percent copper recovery and acid Mineral Chemical formula Weight content (%)
consumption per tonne indicate that the materials in all three produc­
Covellite CuS 0.4
tion zones displayed a common response to in-situ leaching. Hence, it Chalcopyrite CuFeS₂ 0.3
can be postulated that the in-situ leach response of any one of those Pyrite FeS₂ 2.2
production zones was representative of the entire resource comprised by Quartz SiO₂ 40.3
Potassium feldspar KAlSi₃O₈ 11.3
the spent heaps of Quebrada Blanca. The in-situ leach response was most
Albite NaAlSi₃O₈ 8.3
likely governed by the mineralogical composition of the material in the Plagioclase NaAlSi₃O₈-CaAl₂Si₂O₈ 2.3
production zone. Mineralogical analyses were only carried out on pro­ Muscovite KAl₂(AlSi₃O₁₀)(F,OH)₂ 27.3
duction zone material from the initial pilot test, summarized in Table 6. Biotite K(Mg,Fe)₃(AlSi₃O₁₀)(F,OH)₂ 1.4
However, the common in-situ leach response means that the mineral­ Kaolinite Al₂Si₂O₅(OH)₄ 5.1
Chlorite (Mg,Fe)₃(Si,Al)₄O₁₀(OH)₂(Mg,Fe)₃(OH) 0.1
ogical profile in Table 6 likely also represents that of the industrial pilot

and the combined heaps No. 6, 7, and 8, if not the entire resource. Iron hydroxide FeO(OH) 0.1
One consideration regarding recovery is whether the copper ob­ Rutile TiO₂ 0.3
tained by in-situ leaching could have been obtained by simply extending
the leach cycle of the original heaps. When the heaps listed in Table 5
were stacked between 1995 and 1998, the sequential copper grade was than what remained in the spent heaps. It is for that reason that the
>1%. Those heaps were leached, on average, for about 500 days and heaps in Table 5 were abandoned back then and had to wait until in-situ
yielded an overall recovery of 76%. Arguably, those heaps could have leaching made economic sense. As a point of interest, ore mined in 2018
been irrigated for an additional year and recovered the same or more had a grade of 0.3% sequential copper.
copper as 34 months of in-situ leaching 20 years later. However, that
strategy could not compete with leaching the fresh ore being mined at 4.2. Oxidative nature of the in-situ leaching
the time. The heaps that replaced the ones listed in Table 5 had
sequential copper grades higher than 0.9%; they recovered more copper Assuming that the mineralogical profile in Table 6 was representa­
tive of the spent heaps at Quebrada Blanca, then covellite (CuS) was the
secondary sulphide mineral that was leached in-situ. Hence, copper
would have been released according to the reactions in Eqs. 1–3. This
work assumed that the ferric regeneration reaction in Eq. 2 was facili­
tated predominantly through bacterial action that consumed acid and
oxygen. It is possible that covellite oxidation occurred abiotically, or
that the leaching observed was simply the mobilization of copper bound
in ferric precipitates (see Table 6) over many years (Petersen et al.,
2013). Bacterial action was deemed more plausible, however, because
bacterial oxidative activity has been measured in the Quebrada Blanca
heaps (Lizama, 2001) and bioleaching bacteria were detected in solids
and solution samples from the industrial pilot test (data not shown).
Assuming that bacterial action took place, it could have done so
within the production zone or external to it. This study accounted for
internal and external possible bacterial action by using injection flows
and differential ferric ion concentration between PLS and raffinate, the
ΔFe3+. In essence, the daily copper production justified by ferric con­
sumption according to the chemical reaction in Eq. 1 was considered as
Fig. 15. Copper recovery as a function of the injection ratio for the initial pilot chemical leaching of covellite. The ferric consumed by that chemical
test, the industrial pilot test, and the combined heaps 6, 7, and 8. leaching was deemed to be replenished by bacterial action that was

8
H.M. Lizama Hydrometallurgy 215 (2023) 105997

outside of the production zone. In-situ leach operations comprised only a Blanca operations.
portion of Quebrada Blanca's leach circuit. The bulk of the leach solu­ Two examples bear mentioning on the use of ERT geophysics for safe
tions were used in conventional heap and dump leach operations where operation of in-situ leaching of the spent heaps. One example concerned
chemical reaction in Eq. 2 would have been prevalent. In-situ copper the start-up of the industrial pilot when the rise in the phreatic level at
production not accounted by chemical reaction in Eq. 1 was deemed to the west toe prompted the restriction of injection flow. Geophysical
have been obtained through bacterial activity within the production surveys showed that the rise in phreatic levels were explainable by the
zone according to reaction in Eq. 3. Fig. 17 compares the degree of topography of the primary liner, which formed underground basins that
copper leaching directly attributable to ferric leaching of covellite for channeled internal solution flows. Fig. 18 shows the flow directions and
the three production zones. magnitudes over the surface of the liner underneath heaps 1 and 2,
The ΔFe3+ data used to construct the plots in Fig. 17 was highly calculated from geophysical data.
variable, as indicated by the large error bars. Indeed, there was no sta­ The flow paths apparent in Fig. 18 showed that the high phreatic
tistically significant difference between the three data sets. On average, levels were located right at the mouth of an internal catchment that
the ferric consumption in the three production zones accounted for covered 20% of the production zone's bottom surface, receiving the flow
approximately 40% of the copper leached, meaning that 60% of the of 28 injection wells. Thanks to that visualization, the area was rein­
copper was produced more likely than not by oxygen consumption (Eq. forced with a buttress and two production wells were installed to pump
3). Any acid-consuming, oxidative leaching, would have required the out solution.
presence and replenishment of oxygen in the production zones. This Another example of using geophysics for safe operation of in-situ
replenishment likely came by air circulation through the drainage sys­ leaching concerned Heap No. 13. The phreatic levels in that heap rose
tem underneath the buried static pad and by the ingress of air by way of steadily during the first six months of operation. Furthermore, seepage
the open wellbores during rest periods. of leach solution was observed on the north slope of that heap. Again,
geophysical surveys showed that the seepage and phreatic levels were
explainable by the topography of the primary liner and the internal
4.3. Injection management
solution flows, or lack of them. Fig. 19 shows the geophysics-derived
flow directions and magnitudes over the surface of the liner under­
Injection management was key to in-situ leaching of the Quebrada
neath the entire South Pad of the Quebrada Blanca heap.
Blanca spent heaps. First, a sufficient volume of solution, applied at a
The flow paths apparent in Fig. 19 showed that approximately 40%
high enough rate, was necessary for the wetting fronts of the injection
of Heap No. 13's production zone bottom surface had no flow paths
plumes to meet and thereby maximize solution-mineral contact. That
available for drainage. Hence, the solution injected into that part of the
injection flow had to be high enough to overcome gravity drainage but
production zone had nowhere to go and was accumulating. Indeed, some
low enough to prevent channeling. Any channeling of leaching solution
seepage was observed at the north toe of Heap No. 13. In response to that
would have caused heterogeneous wetting of the production zone,
visualization, solution flow was shut off to that section of the affected
decreasing copper recovery. Gravity drainage depends on the hydraulic
heap.
conductivity of the production zone. The initial pilot test measured
hydraulic conductivities of up to 10− 4 m/s: between sand and gravel.
5. Conclusions
Second, the saturation or partial saturation of the zone of influence had
to alternate with rest periods that allowed for sufficient oxygenation.
In-situ leaching of spent heaps at Quebrada Blanca was successful in
The rest periods needed to be of sufficient duration to drain or partially
creating a resource from material that had been considered a waste
drain the zone of influence about the injection well and allow seepage of
product. As with any new application of technology, there were tech­
air into the pore spaces of the production zone.
nical and economic risks involved. Hence, in-situ leaching of the spent
Injection management was accomplished by way of ERT geophysics.
heaps was first tested at the pilot scale to establish technical feasibility,
ERT surveys and imaging allowed visualization of relative changes in
then tested at the industrial scale to establish economic feasibility. Full
moisture contents within production zone. That, in turn, allowed esti­
scale implementation produced a significant amount of copper in a
mating the degree of wetting and sweep efficiency. In addition,
manner that was safer and more economical than re-mining, agglom­
geophysics served as an indicator tool for safe operation of the pro­
eration, and stacking of the material in the spent heaps.
duction zone; safety was and is the most important aspect of Quebrada
In-situ leaching of the spent heaps required that a sufficient volume
of leaching solution was applied to ensure uniform wetting of the ma­
terial within an injection well's zone of influence. At the same time, the
injection flow needed to be high enough to overcome gravity drainage
but low enough to prevent heterogeneous wetting due to solution
channeling. The key to achieving that balance was through monitoring
the movement of the injected leach solution through the material being
leached. That was accomplished by wireline geophysics. Geophysics was
also instrumental for mapping out potential areas for in-situ leaching
and identifying potential hazards during operation.
This application of in-situ leaching concerned copper sulphides,
which involve oxidative reactions potentially catalyzed by bacteria.
Hence, air had to be supplied to the system, either directly through
passive aeration or indirectly through dissolved ferric ions. Oxidative
conditions were facilitated by utilizing the original heap drainage sys­
tem to encourage air entry from below, and alternating periods of in­
jection with rest periods where the injection wells were opened to
atmosphere, encouraging air entry.
Fig. 17. Fraction of copper leached directly by ferric ions and indirectly by
(likely) bacterially assisted oxidation in the initial pilot test, the industrial pilot Declaration of Competing Interest
test, and the combined operation of heaps 6, 7, and 8. The error bars represent
one standard deviation. The author declares that they have no known competing financial

9
H.M. Lizama Hydrometallurgy 215 (2023) 105997

Fig. 18. Map of calculated preferential flows over the primary liner surface based on ERT geophysical imaging. The visualized portion of the primary liner surface
corresponds to the lower boundary of the industrial pilot production zone. Note how the location of high phreatic levels coincides with a single outlet for 20% of the
production zone drainage.

Fig. 19. Map of calculated preferential flows over the primary liner surface based on ERT geophysical imaging. The visualized portion of the primary liner surface
corresponds to the lower boundaries of the production zones corresponding to heaps No. 9 through 17. Note how the locations of high phreatic levels and seepage
coincided with the lack of available flow paths to drain away injected solution.

10
H.M. Lizama Hydrometallurgy 215 (2023) 105997

interests or personal relationships that could have appeared to influence Lizama, H., Arrué, D., 2020. Recovery of Copper from Heap Leach Residues. United
States Patent Application US2020/0340077A1, October 29, 2020.
the work reported in this paper. The author declares that all funding
Lizama, H., Arrué, D., Hurtado, R., Pizarro, F., Armenti, M., Witzel, M., 2018. Piloting of
related to this paper was provided by Teck Metals Ltd. and Teck Que­ Heap Secondary Leaching at Quebrada Blanca. HydroProcess 2018 Proceedings,
brada Blanca S.A. The author is employed by Teck Metals Ltd. Santiago, Chile.
Lizama, H.M., 2001. Copper bioleaching behaviour in an aerated heap. Int. J. Miner.
Process. 62, 257–269.
Acknowledgements Parkison, G.A., Bhappu, R.B., 1995. The Sequential Copper Analysis Method –
Geological, Mineralogical, and Metallurgical Implications. SME Preprint No. 95–90,
The author wishes to thank Teck Metals Ltd. and Teck Quebrada Littleton, CO, U.S.A.
Petersen, J., Dixon, D.G., 2007. Principles, Mechanisms and Dynamics of Chalcocite
Blanca S.A. for permission to publish this work; to D. Arrué, M. Armenti, Heap Bioleaching. Microbial Processing of Metal Sulfides. Springer, Dordrecht,
R. Hurtado, D. Zarchikoff, F. Pizarro, A. van Staden, S. Herlitz for their Netherlands, pp. 193–218.
technical assistance; and to M. Witzel, R. Roco, P. Wan, and B. Benson Petersen, J., Muzawazi, C., Mwase, J., Jones, G., 2013. Ammonia Heap Leaching of
Chalcopyrite in Low Grade Ores - a Fresh Perspective? Copper 2013 Proceedings,
for their support. Santiago, Chile.
Rucker, D., 2015. Deep well rinsing of a copper oxide heap. Hydrometallurgy 153,
References 145–153.
Sinclair, L., Thompson, J., 2015. In situ leaching of copper: challenges and future
prospects. Hydrometallurgy 157, 306–324.
Avendaño, C., 2004. Revisión de la lixiviación en pilas de minerales de cobre. LX Users
Watling, H.R., 2006. The bioleaching of sulphide minerals with emphasis on copper
Conference, La Serena, Chile.
sulphides - a review. Hydrometallurgy 84, 81–108.
Barrow, J.C., 1994. The resonant sonic drilling method: an innovative technology for
Williams, G., Clayton, N., Lang, K., 2015. Integrated Approach to Secondary Copper
environmental restoration programs. Ground Water Monitor. Remed. 14 (2),
Recovery in Historic Stockpiles at the Asarco Ray Mine – a case study. Proceedings of
153–160.
the Heap Leach Solutions 2015 Conference, Reno, NV, USA.
Lang, K., Banas, R., Ruano, R., 2013. The Use of Fiber Optics and Advanced Wireline
Yáñez, H., Alhborn, G., Ramos, S., 2007. Quebrada Blanca, 12 years of bioleaching in
Geophysics to Determine Pregnant Liquor Solution Front, Depth and Velocity.
altitude. In: Menacho, J.M., Casas, J.M. (Eds.), HydroCopper 2007. Gecamin,
Proceedings of the Heap Leach Solutions 2013 Conference, Vancouver, BC, Canada.
Santiago, pp. 149–155.
Leahy, M.J., Schwarz, M.P., 2009. Modelling jarosite precipitation in isothermal
chalcopyrite bioleaching columns. Hydrometallurgy 98, 181–191.

11

You might also like