You are on page 1of 31

Hydrodynamic Behavior of Heap Leach Piles:

Influence of Testing Scale and Material Properties

Sylvie C. Bouffard1 and Paul G. West-Sells

Submitted to Hydrometallurgy Journal

For review and publication

January 6, 2009

Sylvie C. Bouffard, Ph.D., Sr. Metallurgist, Barrick Technology Centre, 323


Alexander St., Vancouver, British Columbia, Canada, V6A 1C4, Tel. 604-682-
1730 ext. 222, Fax 604-682-1767, sbouffard@barrick.com.

Paul G. West-Sells, Ph.D., Formerly Sr. Metallurgist with Barrick Technology


Centre, now Sr. Metallurgist with Western Copper Corporation, 2050-1111 W.
Georgia St., Vancouver, British Columbia, V6E 2M3,
pwest-sells@westerncoppercorp.com.

1
Corresponding author
1
1. ABSTRACT

The dynamics of solution, soluble species, and gas in various heap leaching systems were

quantified with tracer tests to explain differences amongst their leaching behaviors. The tracer

methods included tracer injection in the feed solution and monitoring soluble species during

rinsing. The tracer data was interpreted by the Levenspiel compartment flow model that

considers the packed bed as a black box filled with ore, air, and moisture, the latter occurring as

active flow, by-pass flow, plug flow volume, well-mixed volume, and stagnant/dead volume.

This model was useful as a diagnostic tool but would not be in a predictive fashion.

Hydrodynamics of the heap systems were evaluated during (1) the scale-up from a 350 kg

column to a 55 tonne crib to a 6,300 tonne heap, and (2) the change in ore properties from run-

of-mine to crushed to agglomerated in 55 tonne cribs.

It was established that the gold recovery could be correlated to the amount of stagnant solution in

the system. Columns and cribs that achieved excellent gold recovery had as little as 0% to 18%

of the solution retained in their packed beds in a stagnant form. The 6,300 tonne heap had poor

gold recovery and contained a significant stagnant volume (57% of the total solution retained in

the heap). The proportion of stagnant solution was proportional to the proportion of fines and

the lack of structural support in the heap, the latter translating into a higher bed density.

In contrast to the wide distribution of particle size in the run-of-mine and crushed ores, the

agglomerated ore was more uniformly-sized, which led to a lower proportion of stagnant

solution. The agglomerated ore had a lower bulk density of 1.7 t/m3 (vs 2 t/m3) and a lower

proportion of stagnant solution 0% (vs 88%). Gold extraction from the agglomerated ore was

2
aided by the air available in the pores, while good solution transport out of the packed bed was

attributed to the absence of stagnant zones.

Tracers proved to be simple and effective to quantitatively diagnose the inner workings of heap

leaching systems of various scales. A narrow particle size distribution, a low fines content, and a

low bed bulk density were key to achieving good solution, solute, and gas dynamics, hence high

gold recovery.

2. KEYWORDS

Hydrodynamics; tracer; column; heap; scale-up; modeling, ore properties

3. INTRODUCTION

Roughly fifteen percent of the world’s gold and copper are produced by heap leaching, a simple

process in appearance which consists of stacking coarse ore particles into a pile, irrigating the

pile with a leach solution for anywhere from 60 days for oxidized gold ores to 500 days for

copper sulfide ores, and recovering the dissolved metal value in the leach solution at the bottom

of the pile. Metallurgy and hydrodynamics come face to face in heap leaching processes; the

former is more extensively studied and much better understood than the latter.

During the development of the design criteria for heap leaching a particular ore, the metallurgical

response of the ore is typically measured in systems of increasing size. Initially, column tests are

performed to roughly outline the optimum leaching conditions, and then larger scale tests are

performed. Cribs, large silos at heights equal to anticipated heap lift heights, are often employed

at the conclusion of successful column tests. Demonstration heaps, isolated or embedded piles

that can range from a 1,000 to 100,000 tonnes, are also used to confirm the metallurgical

response before commercial implementation.

3
Insights into the hydrodynamic behavior of the ore emerge only at the crib and heap scale. The

hydrodynamic response is usually not well defined, the methods to quantify it can be expensive,

and the interpretation of the data can be complex. It is important that the solution that is irrigated

on the top of the heap is well distributed as it travels by gravity to the bottom of the heap so that

all of the ore is contacted and can be leached. It is also important that this solution only fills up

part of the void spaces so that the air blown into the bottom of the heap also has a well

distributed path through the heap. The quantity, size and distribution of void spaces and the

percentage that they are filled by air or solution in a heap can be affected by several aspects of

the heap system. The percentage of void space in a heap is set by the bulk density of the heap,

which in turn is affected by the crush size and agglomeration of the ore. The percentage of void

space that is filled by solution can be partially controlled by changing the irrigation rate, but is

also controlled by the bulk density of the ore, and the size distribution of the ore being stacked.

The following are some examples of the parameters defining the hydrodynamic response and the

methods employed to measure them. Lane et al. (2006) used ultraviolet sensitive dye to

determine the pore volume of coarse ore, which is a hydrodynamic parameter of importance in

the modeling of the diffusion of solute species in porous ores. Videla et al. (2005) used a CT

scan image of a column filled with ore particles to measure the moisture holdup and the air

volume – two parameters defining the hydrodynamic response of the column leach. Guzman et

al. (2006) also measured the moisture holdup, in this instance, in large test cells embedded

within a large heap using electrical resistivity tomography (ERT). The data pointed to some

areas partially wet while others were too wet to permit air movement vertically through the heap.

ERT was also applied by Water Management Consultants at Minera Escondida in Northern

4
Chile, Cripple Creek heap leach in Colorado, and Phelps Dodge’s Tyron property in New

Mexico (Water Management Consultants website).

Many others have used tracer tests (radioactive iodine, radioactive bromide, ZnSO4, Mn, NaBr,

LiCl, NaNO3, or HCl) in combination with flow models inspired from trickle bed reactors to

deduce many time-dependent hydrodynamic parameters, such as the plug-flow velocity, the axial

dispersion, the rate of exchange between mobile and immobile phases, etc. In most tests,

whether performed in columns (Villanueva et al. 1990; Dix et al., 1992; Comba and McGill,

1992; Decker and Tyler, 1999; Bouffard and Dixon, 2001; de Andrade Lima, 2006), cribs (Murr,

1979; Murr, 1980; Schlitt, 1984), or heaps (Howard, 1968; Armstrong et al., 1971; Villanueva et

al., 1990; Diaz and Duran, 1997), the profile of the concentration of the tracer in the solution

draining from the pile was an asymmetric curve with a long tail, when the tracer was injected in

a single shot (as a pulse). The tracer concentration profiles were fitted with either a piston-

dispersion model (Yusef, 1984; Villanueva et al., 1990), a piston-dispersion-exchange model

(Decker and Tyler, 1999; de Andrade Lima, 2006), or a piston-exchange-diffusion model

(Bouffard and Dixon, 2001; Bujalski et al., 2003). Menacho et al. (2007) were inspired by the

flow model for partially-saturated soils that commonly use the van Genuchten parameters

characterizing soil wetting and drainage. Using this model, the authors evaluated the impact of

the proportion of -74 µm (-200 mesh) fines on the heap height and moisture content with depth.

The majority of the model fits indicated the presence of a significant volume of stagnant solution

wetting the particles. Miller (2003) represented the stagnant solution as tortuous pores existing

inside particles and between particles. Diffusion inside the internal pores was shown to be much

slower than diffusion outside the particles. While Bouffard and Dixon (2001) found that the

stagnant solution volume was marginally affected by any change in crush size, agglomeration, or

5
operating conditions, de Andrade Lima (2006) found a correlation between the stagnant solution

volume and the Reynolds number, as well as correlated the mass transfer coefficient to the

Reynolds number.

In this paper, simple tracer tests using sodium bromide and sodium chloride were conducted to

characterize the hydrodynamic behavior of heap leach systems. The normalized tracer curves

were fitted with a mixed flow model where the total volume of solution held in the pile of ore

particles is split between plug flow, well-mixed, and stagnant/dead volumes and where the total

solution flow is split between a by-pass stream and an active flow.

This paper investigates the hydrodynamics of heap systems in three ways. First, we present the

results of tracer tests in columns, cribs, and heaps, three systems of widely different scale, and all

three loaded with the same material (a gold ore from Nevada). Second, we investigate the

hydrodynamic behavior of agglomerated, crushed, and run-of-mine ore, the three most typical

materials stacked in heap leach piles, with all three prepared with the same ore. Third, we relate

the metallurgical performance of the gold ore tested to the hydrodynamic behavior of the heap

leach pile stacked with the ore.

4. EXPERIMENTAL METHODS

The hydrodynamics of soluble species (solutes) were measured in columns, cribs and a

demonstration heap using a combination of methods.

4.1 COLUMN

A PVC pipe 6 m long and 25 cm in diameter was equipped with flanges to become a column

apparatus. The column was filled with 350 kg (dry) of gold ore particles that were previously

agglomerated with the leaching solution and with 0.3 kg/t of a commercially available binder.

6
These were the same agglomerates as loaded in Crib #3, which is described in the next section.

The moist agglomerates were loaded in the column to an initial height of 6.1 m, giving an initial

bulk density of 1.14 t/m3. The surface of the ore bed was evenly wetted by a perforated

distributor plate that released every 15 min the equivalent volume of leach solution that would

otherwise have been applied during 15 min at a constant rate of 3 L/m2/h. Throughout the leach,

the concentration of reagent in the feed and discharge solution was measured, as well as the bed

height. At the end of the leach, the feed solution was replaced with water, producing a step-

down tracer test with respect to the reagent input. Measurement of the reagent concentration in

the discharge solution continued until the concentration was close to nil. Table 1 summarizes

key physical conditions of the column.

4.2 CRIBS

Three concrete cribs, each 2.4 m wide × 2.4 m deep × 6.7 m tall, were constructed to test the

heap leaching of a gold ore. The cribs were built outdoors, and because of that, a roof was built

to avoid rain infiltration. Outdoor temperatures ranged from -5°C to 25°C during the 12 months

of testing, but the temperature inside the cribs was larger than 10°C during the colder months of

winter.

The ore used in all cribs was the same gold ore, only that it was prepared differently. Crib #1

was loaded with 63 tonnes (dry) of run-of-mine ore to an initial height of 5.8 m. Crib #2

contained 62 tonnes of -38 mm crushed ore stacked to 6.3 m. The crushed ore had an average

P80 of 19 mm (3/4 inch) and a fines content (-74 µm (200 mesh)) of 10-20%. Of the unused

crushed ore, 54 tonnes was agglomerated with a solution and with 0.3 kg/t of a commercially

available binder. The agglomerates were then loaded to a height of 6.1 m in Crib #3. Table 1

presents other physical characteristics of the cribs.


7
The irrigation mode of the column could obviously not be repeated in the cribs because of the

large surface area to irrigate. Instead, the irrigation mode chosen was that of real commercial

heap, i.e. six industrial drip tubes spaced every 46 cm with emitters every 46 cm were laid on the

surface of the crib. The cribs were irrigated with the same leach solution continuously. Cribs #1

and #2 were the first to be loaded and irrigated. The irrigation rate was set at 6 L/m2/h in both

cribs. Noticing the significant slump of both beds, a change to the irrigation rate of crib #3 was

thought necessary to control slumping. The irrigation rate of Crib #3 was set at 3 L/m2/h.

During the 12 months of testing, the concentration of gold, reagents, and leach products in the

feed and discharge solutions was measured, as well as the flowrate of these streams. The crib

bulk density was calculated based on the ore volume and tonnage. On occasion, drill cores were

extracted using a Pionjar split-spoon drill from the surface to the bottom of the cribs.

At the end of the leach, a step-down tracer test using only water was performed. In a step-down

tracer test, water is applied to replace the moisture retained in the crib. The water does not

contain any tracer. The reagents contained in the moisture retained in the crib become the tracer.

As the reagents were pushed out of the crib by the incoming water, their concentration in the

discharge solution dropped. A regular monitoring of the concentration of reagents in the

discharge solution generated a tracer curve.

4.3 HEAP

A heap leach pad was built with the same ore as in the columns and cribs. The principal

difference between this ore and the ore used in the columns and in the crib was the higher fines

content (25% vs 17% of -74 µm (200 mesh) fines). Before stacking, the ore was agglomerated in

2 ways. In one half of the heap, 3,150 tonnes of ore was agglomerated with the leach solution in

8
a drum agglomerator and then stacked with a radial conveyor. In the other half of the heap,

3,150 tonnes of ore was agglomerated in the same drum with the same leach solution but also

with 0.3 kg per tonne of a commercially available binder, as previously done in Crib #3. The

size distribution of agglomerates can be found in this publication (Bouffard, 2008).

The whole heap was 55 m long by 55 m wide by 3.4 m tall. The 2 zones were adjacent to one

another and there were no physical barrier between them. The physical characteristics of the

heap are found in Table 1.

The top and sides of the heap were irrigated with a leach solution applied with drip tubes, using

the same drip tubes and the same drip layout as the cribs. The irrigation rate was set to 3 L/m2/h.

During the course of the leach, solution samples were taken from the feed solution, discharge

solution, and from within the heap using six porous ceramic cups that were buried within the

heap. By creating a vacuum in the cup, the solution in the vicinity of the cup was sucked into it,

pumped through a tube extending to the surface, and then assayed. The heap bulk density was

calculated from knowledge of the GPS coordinates of the heap volume and the tonnage stacked.

The heap was sampled with a Pionjar drill for cores during the leach, and with a shovel for grab

samples after leaching.

Also at the end of the leach, the feed solution was replaced with water containing 100 mg/L of

sodium bromide. Simultaneously, the heap underwent two tracer tests, the first corresponding to

the step-up injection of bromide, the second corresponding to the step-down release of species

present in the heap.

9
4.4 MODELING METHODS

The tracer curves that were produced from each testing apparatus were modeled to diagnose the

flow and solute dynamics in order to better explain the gold recovery in each apparatus. In this

study, the testing apparatus were very large, which potentially could have given rise to 3D fluid

and solute dynamics. It was not the purpose of this study to develop a 3D model, but rather to

diagnose, rather than predict, the fluid and solute dynamics using a simple model having few but

meaningful parameters. This study uses the concepts of Levenspiel to model a packed bed

volume as a combination of one or more flow volumes. These volumes are shown at the bottom

of Figure 1. These volumes include: plug flow, by-passing flow, well-mixed volume, and

stagnant/dead volume. The by-passing flow differs from the plug flow only by its much shorter

residence time. In the plug volume, the concentration at the front of the plug could be different

than the concentration at the back of the plug. In the by-passing flow, the concentration at the

front of the stream is nearly the same as the concentration at the back of the stream. The

stagnant/dead volume is a sink volume capturing solutes. It is not the purpose of this model to

understand how solutes transfer into the stagnant/dead volumes or migrate into this volume.

The well-mixed volume is exactly equivalent to a stirred tank.

The relative importance of each volume in relation to the total volume of solution held up in the

ore bed can be determined from a tracer curve. By definition, a tracer curve is performed under

non-steady state condition. Using the non-steady data from the tracer test, it is possible to

determine the respective volumes under steady-state conditions. Therefore, while the model has

time as an independent variable, it is very much a steady-state model.

Another interesting feature of the model is the absence of X, Y, or Z coordinates, making the

model apparently dimensionless, but in actual facts, the model has some dimensionality. The

10
existence of plug flow, by-passing, well-mixed, stagnant, and dead flow in the same packed bed

necessarily implies that the latter is a heterogeneous system. A rigorous model of a

heterogeneous system would have more than one dimension. But it is not the purpose of this

study to develop a rigorous model, but rather to develop a diagnostic model. The Levenspiel

model is in fact a precursor that helps in the development of a rigorous multi-dimensional model

that could be developed at later times.

When performing a step change in concentration in the feed solution, either up or down, the

tracer concentration in the discharge solution is measured over time until a new steady state is

established. The discharge tracer concentration can be normalized with respect to the feed

concentration in a step-up tracer test, or with respect to the initial concentration in the bed in a

step-down tracer test. Normalizing the tracer concentration bounds it between 0 and 1. The

normalized concentration is denoted (1-F). When plotted, the data (t, 1-F) generate a curve. It is

this curve that the Levenspiel model attempts to fit using some parameters. The fit is performed

by minimizing the sum of squared differences between measured and fitted (1-F) values, using

the variable t as an independent variable to calculate the fitted (1-F) value using the above

equations. The fitted (1-F) values are calculated using Eq. 1, which is the most general form of

the Levenspiel model.

va  v Vp 
1− F = exp − a t +  Eq. 1
v  Vm Vm 

In Eq. 1, (1-F) is the normalized concentration from a step change, v is the total solution flowrate

(e.g. L/h), va is the active flow (e.g. L/h), Vm is the well-mixed volume (e.g. L), Vp is the plug

flow volume (e.g. L), and t is the time (e.g. h). The independent variable t and dependent

variable (1-F) are obtained directly from the tracer tests. Two equations must also be satisfied:

11
V = Vm + V p + V d and v = v a + vb Eq. 2

where the subscripts d and b refer to dead/stagnant volume and by-pass flow, respectively. The

total volume V and solution flow v are known before the start of the tracer tests. The least-square

minimization scheme attempts to determine the values of va, Vm, and Vp, the values of vb and Vd

being calculated from Eq. 2 by difference.

Because of the different irrigation rate and moisture content in the various setups, it was suitable

to normalize the test conditions using the parameter “ratio of applied solution/moisture content”

defined as:

Applied solution v
= t Eq. 3
Moisture content V

This parameter is dimensionless, has a unit of 1 when exactly 1 pore volume of solution has been

applied, and modifies Eq. 1 as follows:

va  v V  v  Vp 
1− F = exp − a  t +  Eq. 4
v  v Vm  V  Vm 

When the X-axis of Figure 1 has units of time, Eq. 1 is employed, but when the units of the X-

axis corresponds to the dimensionless pore volume ratio (Eq. 3), Eq. 4 is used.

Let us now examine some special cases that arise from simplifying either Eqs.1 or 4. If all four

types of flows exist within a system, the normalized concentration curve that is produced from a

step change will take the form shown in Figure 1. If there is no by-pass flow, the straight line

will begin at 1 on the Y-axis instead of below 1. Note that the shape of the curve will be exactly

the same whether one envisions the plug flow before or after the well-mixed volume. If there is

no plug flow, the straight line will disappear, leaving only the parabolic shape. If there is no

12
stagnant/dead volume, the parabolic portion of the curve will stretch further to the right. If there

is no well-mixed flow, the parabolic portion will be compressed to a straight vertical line.

For this modeling purpose, it was satisfactory to treat each apparatus as a black box containing

multiple volumes, but it was not critical to find out the actual position of each volume relative to

the others. What the Levenspiel model lacks in predictive capability and rigorous treatment of

flow and solute movement, it makes up in its simplicity of use and its interpretation of the

diagnosis.

5. RESULTS AND DISCUSSION

This section quantifies the solute hydrodynamics as a function of the testing scale and ore

properties. To study the influence of the testing scale, ore agglomerated under similar conditions

was placed in columns containing 0.35 tonne of material, in a crib containing 54 tonnes of

material, and in a heap containing 6,300 tonnes. To study the influence of the material

properties, run-of-mine ore, crushed ore, and agglomerated crushed ore were stacked into

separate cribs holding up to 63 tonnes of material. The following two sections analyze the

hydrodynamic behavior in these systems.

5.1 INFLUENCE OF TESTING SCALE

5.1.1 Column

After 2 weeks of irrigation, the column bulk density increased from 1.14 to 1.37 t/m3 and the

moisture retained increased from 8% to 13%. Despite the compaction of the bed, 26% of the

void volume was still filled with air. At the end of the leach, a step-down tracer test was

conducted, i.e. the leach solution was replaced by water. The normalized concentration of the

tracer is plotted as a function of the dimensionless time in Figure 2. The measured concentration

13
dropped very rapidly before 1 pore volume of rinse water was applied, but the drop was not so

instantaneous that the model could be reduced to pure plug flow. The model fits the higher

concentrations well but not the lower concentrations. The model parameters were found to be:

va = v, Vm = 0.41V, Vd = 0, Vp = 0.59V, when inputting the measured value of 47 L for the total

solution volume V. Under these conditions, there is a relatively equal amount of well-mixed and

plug flow volumes.

The same model was fitted again to the same data, but with the distinction that the total solution

volume V, which should be strictly known from the ore moisture content and the total ore mass,

was considered an unknown parameter to be fitted simultaneously with the other parameters.

The fit obtained is greatly improved in the lower range of concentrations. The new model

parameters were found to be: va = v, Vm = 0.29V, Vd = 0, Vp = 0.71V, where V = 40 L. The

difference between the fitted volume of 40 L and the measured volume of 47 L can easily be

explained by changes in the moisture content of the ore. The change is equivalent to a 1.9%

decrease in the moisture content of the ore.

The presence of a plug flow volume could be associated to the mode of solution application,

which consisted of discharging at once every 15 min the volume of solution that would otherwise

have been applied continuously for 15 min. Every 15 min, a pulse of solution was applied

evenly over the column surface, which resulted in very good solution distribution.

5.1.2 Crib

Crib #3 was loaded with the same agglomerates at the column, but was irrigated with drips rather

than a pulsating distributor. After two weeks of irrigation, the bulk density increased from 1.53

14
to 1.73 t/m3. The crib quickly retained from 10 to 13 wt% of solution, leaving 10 vol% of the

total crib volume for air transport.

The moisture content of the crib was examined by drilling and testing the core. Surface

evaporation reduced the moisture content at the surface of the crib, but elsewhere in the crib, the

moisture content could not be related to the depth, to the position between drip emitters, or to the

time of leaching.

Application of two times the solution holdup in the crib rinsed out approximately 90% of all

soluble species. The shape of the rinsing profile shown in Figure 2 suggests the presence of a

small plug flow volume and a large well-mixed reservoir. A good fit of the data was obtained

with these parameters: va = v, Vm = 0.88V, Vd = 0, Vp = 0.12V, where V = 6220 L.

The well-mixed conditions that existed in the crib led to fairly uniform gold leaching with depth.

In the middle of the leach cycle, the gold extraction averaged 65% out of a maximum of 90%

and the extraction at the top of the crib was 5% higher than at the bottom. At the end of the

leach cycle, the gap in the extraction between the top and bottom was reduced to about 2%

because of the further leaching at the bottom; the gold extraction averaged 69%. At any time,

the gold extraction was not correlated to the position of the drip emitters.

5.1.3 Heap

Both halves of the heap, one agglomerated without binder and the other with, slumped from 3.4

to 2.3 m within a few days of irrigation. The bulk density increased from 1.45 t/m3 before

irrigation to 2.01 t/m3 after irrigation. The extent of slumping was significant, compared to the

crib which reached a final density of 1.73 t/m3.

15
The heap retained on average 14 wt% moisture. The moisture content was the same everywhere

in the heap, whether in the presence of the binder-agglomerated ore or not, whether at the

beginning or near the end of the leach, or whether under irrigation or drainage. There was a

slight increase in the moisture content with increasing depth of the heap. The most significant

increase in the moisture content was noted at the very bottom of the heap where it reached 20

wt%.

During leaching, the average reagent concentration was the same at the top and bottom of the

plateau of the heap, but larger on the sides of the heap. The larger concentrations on the sides

can be explained by the higher evaporation rate leading to increased concentrations.

At the end of the leach, the heap was rinsed with a 100 mg/L sodium bromide solution to

examine solution flow through the heap. During the rinse, the concentration of reagent (other

than bromide, but the same as in the column or crib) on the sides of the heap dropped

significantly, even though the irrigation to the sides had been turned off. Bromide was detected

at concentrations of up to 70 mg/L on the sides of the heap indicating that lateral diffusion from

the plateau to the sides of the heap occurred.

Soon after the start of the rinse, the reagent concentration dropped very rapidly in the discharge

solution. After 1 pore volume applied, the concentration of reagent in the discharge solution was

reduced to 15% of its initial concentration in the heap (Figure 3, bottom curve). The model with

77% active volume (va = 0.77v), 43% of well-mixed volume (Vm = 0.43V), and 57% of

stagnant/dead volume (Vd = 0.57V) produced the best fit of the data.

16
The top curve in Figure 3 shows that, at the end of the rinse, 70% of the reagent present in the

heap prior to rinsing was not rinsed out. The ratio not-rinsed/rinsed (70%/30%) is fairly

comparable to the fitted ratio (57% stagnant/43% well-mixed).

This analysis can be taken one step further by comparing it to the bromide data. Let us

remember that the bromide tracer test is a step-up tracer test, while the reagent tracer test is a

step-down tracer test. That distinction is very important when comparing which heap volumes

exchange solutes and which don’t. By the time 1 pore volume of bromide solution had been

pumped, 39% of the bromide species was retained in the heap, and 61% passed through the heap

without being retained. The moisture in the heap that picked up the bromide is the same

moisture that released the reagent. This moisture is the well-mixed moisture of 43% according

to the reagent data, and 39% according to the bromide data. Conversely, the moisture in the heap

that did not release the reagent must not have picked up bromide. This moisture is the

stagnant/dead moisture of 57% according to the reagent data, and 61% according to the bromide

data. Overall, there is a very good agreement between the bromide data and the model

prediction: 39% vs 43% and 61% vs 57%.

In his model, Levenspiel considers the stagnant/dead volume as only one volume. We

demonstrated that the stagnant/dead volume in the heap was in fact stagnant. To reach this

conclusion, we monitored over time the gold concentration in the solution retained in the heap by

collecting moisture into 6 porous ceramic cups embedded in the heap. During leaching, 6 out of

6 porous cups indicated that the gold concentration in the moisture around the cups changed.

During rinsing, only 3 of the 6 cups saw a change. If the stagnant/dead volume has been strictly

dead, only the 3 cups that saw a change during rinsing would have also seen a change during

leaching. Since this was not the case, it can be concluded that the solution was not held in dead

17
pore volumes. The solution being predominantly stagnant, it would have taken more than 1 pore

volume to wash out the 60% of reagent still remaining in the heap after rinsing. This conclusion

also implies that soluble gold could have been recovered from any parts of the heap, albeit very

slowly.

5.1.4 Discussion

This section summarizes the changes in density and flow models characterizing the 25 cm

column, the crib, and the heap.

As the bed of agglomerated ore grew wider, the final bulk density after wetting increased from

1.37 t/m3 in the 25 cm column, to 1.73 t/m3 in the crib, and finally to 2.01 t/m3 in the heap. Thus

the difference between the 25 cm column and the heap was as large as 0.64 t/m3. The high

density of the heap can be partially explained by the absence of walls to support the bed of

agglomerates. This led to increased slumping, which increased the bulk density, which brought

agglomerates closer together, increasing the probability of producing stagnant zones between

agglomerates. The binder added in half of the heap did not produce a heap that slumped less.

Differences in the flow models were also noted between column, crib, and heap. The plug flow

volume decreased from 65% in the 25 cm column, to 12% in the crib, to 0% in the heap. The

well-mixed volume reached a maximum in the crib (88% in the crib vs 35% in the column and

43% in the heap). The stagnant/dead volume was 0% in the column and crib, compared to 57%

in the heap. The stagnant/dead volume in the heap was shown to be stagnant, not a dead volume.

The particle size distribution of the ore may have contributed to the large proportion of stagnant

pores in the heap. The ore contained 25% of -74 µm (200 mesh) fines, compared to 15% in the

crib. During the wetting phase, fines help distribute the solution throughout the heap. However,

18
such an abundance of fines, having a very large surface area to wet per unit volume ratio,

increased the moisture content from 11.4% in the crib to 14% in the heap, as well as increased

the proportion of stagnant moisture from 0% in the crib to 70% in the heap.

5.2 INFLUENCE OF MATERIAL PROPERTIES

This section describes the hydrodynamic behavior and leach performance of each of the 3 cribs,

and follows a commentary on the similarities and differences observed between cribs. Let us

recall that the main difference between Cribs #1 and #2 was the particle size distribution of the

ore, i.e. run-of-mine vs crushed ore, and that the main differences between Cribs #2 and #3 were

agglomeration and irrigation rate.

5.2.1 Non-Agglomerated Run-of-Mine Ore

Crib #1 saw its bulk density increased from 1.85 to 1.97 t/m3 within 2 weeks after the start of

irrigation. The ore occupied ≈ 82% of the crib volume after slumping. Crib #1 retained 13 wt%

moisture and had 0% air porosity. The compartment model fitted the concentration data well

(Figure 4, lower curve). The fit was obtained with these values: va = v, Vm = 0.12V, Vd = 0.88V,

Vp = 0. Only 12% of the moisture retained was well-mixed; the rest, 88%, was not exchanging

solutes with the well-mixed solution. Gold recovery in this crib was poor because the crib was

too compact, contained no air, and retained a significant fraction of stagnant solution.

5.2.2 Non-Agglomerated Crushed Ore

Crib #2 also saw its bulk density increased from 1.68 to 1.88 t/m3 within 2 weeks of the start of

irrigation. It retained 14% moisture, slightly more than Crib #1 and also had 0% air porosity.

The moisture holdup was partitioned as 12% well-mixed and 88% dead volume (Figure 4, lower

curve). Rinsing of Crib #2 with water was very fast, so fast that the first discharge solution or so

19
had a concentration already only 40% of the initial concentration within the crib. The model

suggested 0% by-passing flow in favor of 100% active flow, but that 100% funneled through

only 12% of the total moisture retained. The flow passing through this small amount of moisture

was so large with respect to the whole moisture retained, that this flow might as well have been

considered to be a by-passing stream. Fitting the data with the by-pass flow did not produce as

good a fit as by considering 100% active flow. But there is physical evidence from Crib #2

suggesting the channeling of the solution, in particular channeling through the corners, which

were rinsed out almost completely, surmising that the rest of the crib was not rinsed out at all,

thus why 88% of the moisture was modeled to be a dead volume. Whether the solution was by-

passing through the corners or whether the well-mixed moisture existed along the same corners

is a pedantic issue. What is undisputable is the very low amount of ore in contact with well-

mixed solution.

For the same reasons as Crib #1, gold recovery in Crib #2 was poor. Most of the gold that was

recovered came from the top third of Crib #2. The extraction in the bottom two-thirds of Crib #2

was close to zero.

5.2.3 Agglomerated Crushed Ore

Crib #3, which contained crushed agglomerated ore, was irrigated 3 L/m2/h, or half the rate of

Cribs #1 and #2. Its final bulk density was only 1.73 t/m3, significantly less than the other two

cribs. At the end of the leach, the gold extraction was 86% in Crib #3, compared to significantly

less in Cribs #1 and #2. The better performance of Crib #3 is attributed to better solution and air

transport. This better transport was likely due to agglomeration, which resulted in less slumping

and thus increased air porosity from 0 to 10%, and reduced the moisture holdup from 14% to

20
11.4%. The model fit indicated that there was 88% of well-mixed volume and 12% of plug flow

(Figure 4, upper curve).

5.2.4 Discussion

The final bulk density of the cribs decreased from 1.97 t/m3 in Crib #1 filled with run-of-mine

ore, to 1.88 t/m3 in Crib #2 filled with crushed ore, to 1.73 t/m3 in Crib #3 with agglomerated

crushed ore. The bulk density was thus inversely proportional to the homogeneity of the

apparent material stacked, the run-of-mine being the least homogenous of all with a wide range

of particle sizes. This result is consistent with hydrology theory. The homogeneity of the

material also affected the moisture holdup; the moisture was 2% higher in Cribs #1 and #2 than

in Crib #3.

The hydrodynamic behavior of Cribs #1 and #2 was identical. 88% of the solution retained filled

up dead volume completely disconnected from the remainder (12%) of the solution considered to

be well-mixed that may have flowed along the crib walls.

There was 0% air porosity in Cribs #1 and #2. The measured air permeability was low, not

permitting upward flow of air at irrigation rates larger than 3 L/m2/h. These conditions resulted

in better gold recovery from the top of the crib where inward diffusion of air supplied oxygen

and from the corners of the cribs where the air blown at the bottom of the cribs escaped.

The hydrodynamic behavior of Crib #3 was more ideal than Cribs #1 and #2. 100% of the

solution was well-mixed. There were no dead zones. Air occupied 10% of the crib volume.

Gold leaching occurred at practically the same rate and to the same extent (69%) everywhere in

Crib #3. Because of improved solute dynamics due to agglomeration creating a more porous bed

of lower bulk density (< 1.73 t/m3), significantly more gold was recovered from Crib #3 than

21
Cribs #1 or #2. The better performance of Crib #3 compared to Cribs #1 and #2 could be

attributed to agglomeration, with little to no influence of the crib walls, irrigation rate, or use of

binder for agglomeration.

6. CONCLUSIONS

This study showed that the compartment flow model developed by Levenspiel effectively

modeled solution transport in various unsaturated packed beds, from columns of various

diameters and heights, to cribs, and to a demonstration heap. All packed beds were stacked with

gold ore particles undergoing leaching. Under all conditions tested, a simplified or generalized

form of the compartment model fitted well the normalized concentration data obtained from a

step-up or step-down tracer test in the packed bed.

None of the systems tested were either 100% plug flow, 100% well-mixed, or 100% stagnant

volume. The closest to 100% plug flow behavior was a column 25 cm in diameter where 61% of

the volume was plug flow. The closest to 100% well-mixed behavior was a crib of 2.4 m × 2.4 m

surface area having 88% well-mixed flow. The closest to 100% stagnant volume was the

demonstration heap of 6,300 tonnes where 57% of the volume was stagnant. In short, increasing

the cross sectional area of the packed bed shifted from flow ideality to non-ideality, increasing

the stagnant volume at the expense of the plug flow volume.

In tests where good gold recovery was obtained, there existed significant well-mixed and plug

flow, the stagnant/dead volume was low or zero, and there was good air porosity. The bulk

density after slumping appeared to be the best predictor of whether there was significant

stagnant/dead volume. For this ore in these systems, a bulk density of 1.75 – 1.80 t/m3 appeared

22
to be the cut-off above which a significant portion of the solution retained in the bed became

stagnant, air porosity was minimal and gold recovery was poor.

This work demonstrated that column testwork provides useful information solely about the

metallurgical response of the ore; the hydrodynamic behavior was not scalable to cribs or heaps.

Valuable insights into the metallurgical and hydrodynamic response of the ore can be obtained

from crib tests, but little to moderate confidence should be assigned to the hydrodynamic

response from a crib when dealing with problematic ores (high clay content, high fines content,

significant decrepitation, need for agglomeration).

Agglomeration of the ore appeared to be the most effective way to keep the bulk density low,

avoiding the presence of stagnant/dead volume and low air porosity. Unfortunately,

agglomeration did not work in all cases as the demonstration heap, stacked with agglomerated

ore, had one of the highest initial bulk densities tested, then slumped significantly to reach the

highest bulk densities tested. To reduce the bulk density, screening the fines prior to

agglomeration seems to be the most sensible approach to heap leach this ore.

7. ACKNOWLEDGEMENTS

The authors express sincere thanks to the staff of the Barrick Technology Centre for assistance in

the execution of the testwork and to Barrick Gold Corporation for authorization to publish this

paper.

8. REFERENCES

F.E. Armstrong, G.C. Evans, G.E. Fletcher, 1971. Tritiated water as a tracer in the dump

leaching of copper, US Bureau of Mines, RI 7510

23
S.C. Bouffard, 2008. Agglomeration for heap leaching: equipment design, agglomerate quality

control, and impact on the heap leach process, Minerals Engineering, 21: 1115-1125.

S.C. Bouffard, D.G. Dixon, 2001. Investigative study into the hydrodynamics of heap leaching

processes. Metallurgical and Materials Transactions B, 32B, 763–776

J. Bujalski, R. Tiller-Jeffery, H. Watling, M.P. Schwarz, 2003. CFD modeling and comparison

with experimental residence time distributions in single and two phase porous flows,

Proceedings of the Third International Conference on CFD in the Minerals and Process

Industries, CSIRO, Melbourne, Australia, 463-468

P.G. Comba, S.L. McGill, 1992. Rinsing of spent precious metal ores for heap leach

decommissioning. In: J.P., Hager (Ed.), Proceedings of Symposium on TMS Annual Meeting,

EPD Congress. Minerals and Metals Material Society, 17-29

L.R.P. de Andrade Lima, 2006. Liquid axial dispersion and holdup in column leaching, Minerals

Engineering, 19: 37-47

D.L. Decker, S.W. Tyler, 1999. Evaluation of flow and solute transport parameters for heap

leach recovery materials. J. Environ. Quality 18, 543-555

F. Diaz, O. Duran, 1997. Characterization of the solution runoff from copper pile leaching

process using tracers. In: Mineral Engineering ‫ۥ‬97 Conference Abstracts, 16-18

Guzman, R.E. Scheffel, S.M. Flaherty, 2006, Geochemical profiling of a sulfide leaching

operation: A case study. Proceedings of the 2006 SME Annual Meeting.

E.V. Howard, 1968. Chino uses radiation logging for studying dump leaching processes. Min.

Eng. 4, 70-74

24
G.R. Lane, M. Jahraus, M.D. Melker, 2006. Dye penetration techniques used to determine heap

leach potential of a telluride bearing Cripple Creek Breccia Ore. Proceedings of the 38th Annual

Meeting of the Canadian Mineral Processors, C. Hardie (Ed.), The Canadian Institute of Mining,

Metallurgy and Petroleum, Ottawa, Ontario, pp. 71-82

O. Levenspiel, 1999. Chemical Reaction Engineering – Third Edition, John Wiley & Sons:

Toronto, 668 p.

J.M. Menacho, L. E. Gutierrez, P.A. Chavez, Matching theory and practice in heap leaching

processes, Cu 2007, Volume IV (Book 1), Proceedings of the John E. Dutrizac International

Symposium on Copper Hydrometallurgy, P.A. Riveros, D.G. Dixon, D.B. Dreisinger, M.J.

Collins (Eds). Toronto, Canada, . 413-426

G. Miller, 2003, Ore geotechnical effects on copper heap leach kinetics, Hydrometallurgy 2003 –

Proceedings of the Fifth International Conference in Honor of Professor Ian Ritchie – Volume 1:

Leaching and Solution Purification, C.A. Young, A.M. Alfantazi, C.G. Anderson, D.B.

Dreisinger, B. Harris, A. James (Eds.), The Minerals, Metals and Materials Society, Vancouver,

Canada, 329-342

L.E. Murr, 1979. Observations of solution transport, permeability, and leaching reactions in

large, controlled, copper-bearing waste bodies. Hydrometallurgy 5, 67-93

L.E. Murr, 1980. Theory and practice of copper sulphide leaching in dumps an in situ. Miner.

Sci. Eng. 5, 121-189

W.J. Schlitt, 1984. The role of solution management in heap and dump leaching. In: J.B. Hiskey

(Ed.), Au and Ag Heap and Dump Leaching Practice. SME-AIME, 71-83

25
Videla, C.L. Lin, J. Miller, 2005. 3D characterization and analysis of heap leaching systems

using X-ray microtomography (XMT). Hydrocopper 2005, J.M. Menacho and J. M. Casas de

Prada (Eds.), Santiago, Chile, 363-372

I. Villanueva, N Heresi, P. Henriquez, P. Vega, 1990. Tracer techniques in the study of flow

patterns in leaching tanks. Appl. Radiation Isotopes 41 (10-11), 1159-1163

R. Yusuf, 1984. Liquid flow characteristics in heap and dump leaching, M.Sc. Thesis,

University of New south Wales, Australia

26
1

Normalized concentration during step


0.9 vb
va  v Vp 
0.8 1− F = exp− a t + 
v  Vm Vm 
0.7
0.6
change

0.5
V p + Vm
0.4 Area =
ν
0.3
0.2
0.1
0
0 0.5 1 1.5
Time

vb

v va
Vp Vm

Vd

Figure 1: Compartment model where the total volume V is split between plug flow (Vp), well-

mixed (Vm), and stagnant or dead volume (Vd), and where the total solution flow (v) is

split between a by-pass stream (vb) and an active flow (va).

27
1.0
5.1 m column
0.9
5.5 m crib

Normalized concentration
0.8
0.7 Crib model fit

0.6 Column model fit - Unknown V

C/Co
0.5 Column model fit - Known V

0.4
0.3
0.2
0.1
0.0
0.0 1.0 2.0 3.0 4.0 5.0
Solution applied/moisture content
(1 = 1 pore volume applied)

Figure 2: Normalized concentration profiles of tracers during the rinse of a 5 m tall column of

25 cm in diameter loaded with agglomerates and uniformly irrigated at 3 L/m2/h,

compared to the rinsing of a 5.5 m tall crib, loaded with the same agglomerates and

irrigated at the same rate as the column.

28
1.2 100%

Normalized concentration, C/Co


90%

Percentage left in the heap (%)


1.0
80%
70%
0.8
60%
0.6 50%
40%
0.4
30%
20%
0.2
10%
0.0 0%
0.00 0.50 1.00 1.50

Ratio of applied solution/moisture content


(1 = 1 pore volume)

Figure 3: Normalized concentration and recovery profiles of tracers during the rinsing of the

demonstration heap.

29
1.0
0.9 Crib 1

Normalized concentration, C/Co


Crib 2
0.8
Crib 3
0.7 Crib 3 model fit
0.6 Crib 2 model fit
0.5
0.4
0.3
0.2
0.1
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Applied solution/moisture content
(1 = 1 pore volume)

Figure 4: Normalized concentration profile of the reagent during the rinse of Crib #1 containing

63 t of run-of-mine ore, Crib #2 containing 62 t of crushed ore, and Crib #3

containing 54 t of agglomerated ore.

30
Table 1: Physical characteristics of the various testing equipment

Column Heap Crib #1 Crib #2 Crib #3


Material Same ore throughout
Particle size Crushed Crushed ROM Crushed Crushed
Agglomeration Yes Yes No No Yes
Dry tonnage (t) 0.35 6300 62.8 62.2 54.0
2
Irrigation rate (L/m /h) 3 3 6 6 3
Irrigation mode Pulse Drip Drip Drip Drip
Initial (wt%) 7.5-9 7.5-10 3.5 3.5 7.5-9
Moisture
Final (wt%) 12 14-20 13 14 10-13
Initial (m) 6.1 3.4 5.82 6.31 6.10
Final (m) 5.1 2.3 5.46 5.64 5.46
Height
Difference (m) 1.0 1.1 0.36 0.67 0.64
Slump (%) 16 32 6.2 10.6 10.5
Initial (t/m3) 1.14 1.45 1.85 1.68 1.53
3
Bulk density Final (t/m ) 1.37 2.01 1.97 1.88 1.73
Difference (t/m3) 0.23 0.56 0.12 0.20 0.20
Solid (%) 57% --- 82% 79% 72%
Volume
Liquid (%) 17% --- 26% 27% 18%
proportion
Air (%) 26% --- 0% 0% 10%
Plug flow (%) 59-71 0 0 0 12
Well-mixed (%) 29-41 43 12 12 88
Hydrodynamics
Stagnant volume (%) 0 57 88 88 0
Active flow (%) 100 77 100 100 100

31

You might also like