You are on page 1of 7

Minerals Engineering 23 (2010) 504510

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Determination of oxygen gasliquid mass transfer rates in heap bioleach reactors


Jochen Petersen *
Centre for Bioprocess Engineering Research, University of Cape Town, Private Bag X3, Rondebosch 7701, South Africa

a r t i c l e

i n f o

a b s t r a c t
A detailed experimental study is described which was conducted to determine the rate of oxygen gas liquid mass transfer within the packed bed of heap bioleach reactors at different temperatures (22 68 C), using the Na2SO3 method. The raw data was analysed using a simplied lm mass transfer model, making corrections for oxygen solubility in concentrated solution and for increased water vapour partial pressure at elevated temperatures. The results compared favourably against two independent experimental leach studies, indicating kLa values between 33 and 46 h1. The value varied with the particle size distribution of the packing, with kLa assuming larger values for those packings that had a higher nes contents. While kLa increases with temperature, the solubility of oxygen decreases simultaneously, resulting in net mass transfer rates being relatively unaffected by temperature in the range studied. This indicates that thermophile heap bioleach reactors are likely to operate under gasliquid mass transfer limitations, especially at high altitude. 2010 Elsevier Ltd. All rights reserved.

Article history: Received 4 October 2009 Accepted 16 January 2010 Available online 13 February 2010 Keywords: Leaching Heap leaching Bio-leaching Oxygen Gasliquid mass transfer

1. Introduction Oxygen is the primary reactant in the oxidative dissolution of sulphide minerals as it occurs in heap (bio)leaching, regardless of whether this is facilitated by micro-organisms, by the ferric/ferrous iron couple, or proceeds directly at the mineral surface. In an operating heap oxygen has to transfer from air in the interstitial spaces between rocks into the aqueous phase where it taken up by microorganisms or reacts with ferrous iron species directly. Oxygen gas liquid mass transfer is a kinetic process, the rate of which is determined by a number of factors, primarily temperature, gas and liquid composition and gasliquid interfacial area. The overall rate of leaching in a heap bioleach process is determined by the rates of a large number of sub-processes, with the slowest being rate limiting (Petersen and Dixon, 2006a). It is postulated that oxygen gasliquid mass transfer becomes rate limiting in heap bio-leaching at temperatures above 60 C, due to the reduced solubility of oxygen in aqueous solution and to increased rates of mineral oxidation at higher temperatures. The solubility of oxygen in aqueous solution has been well studied, and thermodynamic correlations are available in the literature (for example Tromans, 2000). Likewise, mass transfer correlations for many gasliquid contacting systems in the chemical and biochemical industry, such as aerated tanks or packed bed absorbers, are reported, but similar correlations remain largely unexplored for heap systems. The present study outlines the determination of kLa

values (the gasliquid mass transfer rate constant) for heap systems, both on the basis of leach data and gas adsorption experiments in a heap-like system. 2. Estimation of kLa from column leach experiments A good estimate of the rate of oxidation can be made from an analysis of previous column leach data (Petersen and Dixon, 2002, 2006b). On two separate occasions, Geocoat column tests were run to determine the heat generation behaviour of the mineral concentrate under investigation. In both cases the extent of total oxidation was determined from a headtails analysis of the coated concentrate. From an analysis of relevant thermodynamic data it can be stated that for each mol electrons transferred from mineral to oxygen, approximately 100 kJ heat are generated (Petersen and Dixon, 2002). Here this approach is extended to recognize that the total rate of electron transfer within a column over a certain period of time must correlate with the rate of oxygen uptake into solution over the same period (assuming there is no other electron sink in the system). This can in turn be used to determine the oxygen mass transfer coefcient for a given system. The approach of correlating oxygen consumption with heat generation is commonly applied in the analysis of bioprocesses (Doran, 1995). 2.1. Theory

* Tel.: +27 21 650 5766; fax: +27 21 650 5501. E-mail address: jochen.petersen@uct.ac.za 0892-6875/$ - see front matter 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.mineng.2010.01.006

Oxygen gasliquid mass transfer is usually described by Newtons law of cooling:

J. Petersen / Minerals Engineering 23 (2010) 504510

505

r O2 ;ads kL ac cb

1
Oxygen solubility [ppm]

14 12 10 8 6 4 2 0 0 20 40 60 80 100
pure water, dry air 0.15 M salinity, dry air 0.15 M salinity, sat. air high salinity, high altitude

where kLa is the lumped mass transfer coefcient (incorporating specic gasliquid surface area a, hydrodynamics and effective lm thickness), c the equilibrium concentration of oxygen at the gasliquid interface, which may be determined from thermodynamics (Tromans, 2000), and cb the bulk concentration of oxygen in the solution. As bio-systems react rapidly, it is usually assumed that cb is small (10% of saturation c). Eq. (1) can hence be simplied to

r O2 ;ads % 0:9kL a c

Given a rate of total oxidation, rox, per unit mass of ore from experimental analysis, oxygen consumption rate is simply calculated as follows:

Temperature ('C)
Fig. 1. Theoretical equilibrium oxygen solubility in pure water in dry air, mildly saline water in dry and saturated air as a function of temperature. Also shown is the solubility in highly saline solution at low (high altitude) pressure and saturated air.

r O2 ;cons

! ! ! ! mgO2 mol e 1 mol O2 1 d r ox kgore h kgore d 4 mol e 24 h ! mg 3:2 103 mol O2

3
It is also critical to consider the effect of humidity in a column (or heap) situation. Due to the low gas velocities, passing air is likely to become saturated with water vapour at the given temperature. At saturation the partial pressure of water vapour in the gas phase corresponds to the water vapour pressure at that temperature. This can be calculated as follows (Perry and Green, 1984):

The rate of oxygen consumption per unit volume heap is related back to the rate of transfer into the solution by determining the liquid volume fraction within the heap:

r O2 ;ads r O2 ;cons

where e represents the liquid hold-up to solid mass ratio, which must be carefully evaluated for each case separately. Combining Eqs. (2)(4) thus yields for the mass transfer coefcient:

pvap exp57:154 0:3538 T 7:617 104 T 2 6:0 107 T 3 9


where T is again in Kelvin. The presence of water vapour effectively depresses the partial pressure of oxygen in the gas phase:

kL a % 370

r ox

e c

where kLa is measured in h1, rox in (mol e) kg1 d1, e in L kg1 and c in (mg O2) L1. The only variable is the equilibrium concentration of dissolved oxygen c, which needs to be determined from thermodynamics, and the total oxidation rate rox, determined from experiment. Tromans (2000) offers a comprehensive formula for the calculation of oxygen solubility in pure water as well as solutions. The formula for c (in mol kg1) in pure water is simple:

pO2 0:21P pvap

10

c pO2 k

with pO2 being the oxygen partial pressure in the gas phase in atm (which may change considerably with altitude and water vapour saturation of the heap air). The partition coefcient k is calculated from

Thus the solubility of oxygen in the aqueous phase in a heap system is further depressed by water vapour at elevated temperatures. Fig. 1 shows the equilibrium solubility of oxygen, c (in ppm or mg L1), in pure water in dry air, a 0.15 molal sulphate solution in dry air and in a saturated heap, all at 1 atm of total air pressure at different temperatures, clearly indicating that at elevated temperatures above 60 C, oxygen solubility becomes signicantly depressed because of the water vapour effect. It is interesting to note in this context that the oxygen solubility of a highly saline process solution (u = 0.6) at high altitude (P = 0.65 atm), such as is the situation in certain Chilean copper

k exp

( ) T 299:378 0:092 T T 298 2:0591 104 0:046 T 2 203:35 T ln 298 8:3144 T

where T is in Kelvin. For aqueous solutions, Tromans (2000) proposes the introduction of a coefcient as follows:

heap leach operations (see for example Ojumu et al., 2006), would result in oxygen solubilities as much as 2.5 times less than in the equivalent heap with mild salinity at sea level. This is also indicated in Fig. 1. 2.2. Experimental data

c pO2 k u pO2 k 1 jC i c h

where Ci refers to the molality of the salt i in solution. The parameters j, c and h are empirically determined for different types of dissolved salts, some of these are listed by Tromans. For a 0.15 molal sulphate solution approximately what one would expect in a typical heap leach liquor this stipulates a value for u of about 0.95, i.e. a 5% reduction in oxygen solubility.

Two sets of experimental data of columns run under Geocoat conditions were evaluated in terms of the total oxidation over time under leach (Petersen and Dixon, 2002, 2006b). These data sets were established from two different minerals evaluated under signicantly different conditions.

506

J. Petersen / Minerals Engineering 23 (2010) 504510

70

Table 1 kLa values and oxygen solubilities as determined from the experimental data. T (C) 20 25 30 35 40 45 50 55 60 65 70 c (ppm O2) 8.70 7.91 7.24 6.66 6.15 5.69 5.26 4.86 4.47 4.07 3.66 kLa (h1) 42.57 46.80 51.14 55.60 60.23 65.11 70.37 76.19 82.88 90.91 101.03

Total oxidation [mol e /kg]

60 50 40 30 20 10 0 0 20 40 60 80 100 y = 0.8174x - 4.1111 R = 0.9921


2

Days since inoculation


Fig. 2. Total oxidation prole for chalcopyrite leach (taken from Petersen and Dixon, 2002).

2.2.1. Data set I Experiments with an Australian chalcopyrite concentrate (26.8% Cu, 27.5% Fe, 28.8% S) were conducted (Petersen and Dixon, 2002). Eight identical short column tests were conducted over 100 days. From time to time a column would be dismantled to assess the degree of mineral leaching on the basis of the recovered tails from the Geocoat leach. The total oxidation in terms of electrons transferred could then be calculated on the basis of the Cu, Fe and S mass balances. When plotting the total oxidation data against time (Fig. 2), it becomes obvious that after some initial warm-up, oxidation within the columns proceeds at a more or less constant rate over most of the run, and tends to level off only towards the end. The rate of copper dissolution followed the same linear trend, which was taken as indicative of a reagent supply limited reaction (i.e. limited by oxygen mass transfer). Data points belonging to the linear phase of the curve have been correlated by linear regression to give an oxidation function with time, as shown in Fig. 2. The slope of the line corresponds to the oxidation rate in this phase, i.e. 0.8174 (mol e) kg1 d1. The liquid hold-up in the experiment was not determined, but has been estimated to be in the order of 0.8 L solution per kg concentrate. 2.2.2. Data set II Experiments conducted with an Australian CuAu concentrate (7.4% Cu, 20.7% Fe, 22.7% S) have been partially reported by Petersen and Dixon (2006b). A number of short column tests were conducted at different xed temperatures with the corresponding micro-organisms. Also conducted was an experiment (RT-1) in which the temperature was gradually ramped from ambient to 70 C. Again, the total oxidation was calculated on the basis of
25
Mes. 22 C

headtails assays of Cu, Fe and S. It was noted that the oxidation proles were very similar, regardless of experimental conditions. All data sets indicated an initial acceleration, then continued on a more or less linear prole with some levelling off towards the end (Fig. 3). As can be seen, there is some degree of spread the longer the experiment continues (shaded area). Data points have been correlated by linear regression to determine the rate of oxidation, which ranges from 0.72 (mol e) kg1 d1 for the slowest (RT-1) and 0.85 (mol e) kg1 d1 for the fastest (moderately thermophiles at 52 C) rate observed. These values are within 15% of each other and correspond remarkably well with the value determined from data set I (0.82 (mol e) kg1 d1). Thus the two data sets can be considered as consistent, and the average oxidation rate rox in Eq. (5) can be taken as approximately 0.8 (mol e) kg1 d1. Fig. 3 indicates that this value is not (or only very mildly) a function of temperature in the range 2270 C. In that context it should be noted that the specic ferrous iron oxidation rates of different microorganisms active in different temperature ranges (mesophiles, such as Acidithiobacillus or Leptopspirillum, moderately thermophiles such as Sulfobacillus and extremely thermophiles, such as Sulfolobus) are all very similar (Petersen and Dixon, 2006b, using data from Hansford and Vargas (2001) and Searby and Hansford (2004)). The liquid hold-up in this experiment has also been estimated to be around 0.8 L kg1 concentrate, although it should be noted that concentrate I was coarser, had a higher specic gravity and was more densely coated than that of data set II. However, in either case the overall liquid hold-up in the Geocoat bed (i.e. solution relative to total solids) is between 7% and 8%, which is very typical of agglomerated materials. 2.3. Evaluation of kLa The determination of kLa from the oxidation rate data has been set out in the theory Section 2.1 above. Table 1 presents values for c, and hence kLa for an 0.15 M aqueous sulphate solution in saturated air at atmospheric pressure at various temperatures, calculated from Eq. (5), for rox = 0.8 mol (mol e) kg1 d1 and e = 0.8 L kg1 concentrate. It is now of interest to develop a simple model for the functional relationship between kLa and T, assuming that the overall rate of oxygen adsorption is not a signicant function of temperature (as conrmed by Fig. 3). Assuming that the temperature relationship of kLa can be expressed in terms of a simple Arrhenius term:

Total Oxidation [mol e /kg conc.]

20 15 10 5 0 0

Mod. 52 C Extr. 68 68 C C RT -1

kL a kL aT 0 exp
10 20 30 40 50

   Ea 1 1 R T0 T

11

Days since inoculation


Fig. 3. Oxidation load for the CuAu Geocoat leaches (taken from Petersen and Dixon, 2006b).

then the data presented in Table 1 can be used to determine the activation energy Ea and a start value for kLa (T0). Fig. 4 shows the data points and regression line.

J. Petersen / Minerals Engineering 23 (2010) 504510

507

4.8 4.6

100% 80% 60% 40% 20% 0%

Kennecot Ore Packing 1 Packing 2 Packing 3 Geocoat (estim.)

4.2

4 3.8 3.6 3.4 3.2 3 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006
y = 1682.4x + 3.7393 R2 = 0.9962

Cumm. Passing

4.4

ln k La

10

100

1000

10000

100000

Particle Size [micron]


(1/T0-1/T)
Fig. 4. Regression analysis of kLa vs. temperature. Fig. 5. Particle size distributions for the various packings used in the column adsorption tests.

O2 adsorption rate [mol/L/min]

The data shows almost perfectly linear behaviour, indicating that the non-linear effects of oxygen solubility (Eq. (7)) and water vapour pressure (Eqs. (8) and (9) largely cancel each other out in the temperature range considered here. Thus, a sufciently accurate value for kLa can be predicted in the range 2070 C with kLa (20 C) = 42.6 h1 and E = 14.0 kJ mol1, giving the nal formula for kLa (T):

0.001 0.0009 0.0008 0.0007 0.0006 0.0005 20 30 40 50 60 70


Temperature [C]

Kennecot Ore Packing 1 Packing 2 Packing 3

   14; 000 1 1 kL a 42:6 exp T 0 20  C 293 K 8:314 293 T

12

3. Determination of kLa from column experiments using Na2SO3

3.1. Experimental An alternative route to the determination of oxygen mass transfer in heap systems is to measure the rate of oxygen removal from the gas phase given a predictable rate of oxygen consumption in the liquid phase. The reaction of sodium sulphite with dissolved oxygen in the presence of a dissolved cobalt as catalyst (Cooper et al., 1944) is ideally suited for this:

Fig. 6. Measured volumetric oxygen adsorption rates in the different columns vs. temperature. The dashed lines indicate the average rates given in Table 3 in each case.

Na2 SO3 1=2 O2 ! Na2 SO4


which is rapid and follows rst order reaction kinetics:

13

r 0O2 k T cO2

14

The intrinsic rate constant k0 has been determined experimentally in a series of batch tests in a stirred Erlenmeyer ask, tted with a dissolved oxygen probe at different temperatures (20 50 C). In each test the solution was rst heated and bubbled with oxygen, before adding the Na2SO3 and catalyst and rapidly sealing the ask. The measured rate data (not shown) was conrmed to be rst order and the rate constant k0 (in s1) has been correlated with temperature as follows:

k T 92:76 exp

  18; 000 RT

15

To measure the rate of oxygen adsorption in packed beds, tests were conducted in short 10 cm diameter columns, packed with three different packings of aggregated inert material (sand and coarse gravel) as well as a sample of crushed heap leach ore from the Kennecot Mine (UT, USA). The particle size distributions for the four packings tested are reected in Fig. 5. The columns were heated from the outside, and hermetically sealed to ensure no gas can enter or exist through channels other than the gas ports. The columns were continuously rinsed with a 30 g L1 sodium sulphite solution at 5 L m2 h1 (i.e. 40 mL h1) and aerated at 1.5 N m3 m2 h1 (i.e. 200 mL min1). For each

experiment the column was heated to the target temperature on the day before, then the feed- and off-gas from each column were sampled into 30 L plastic bags. The gas ow rate was conrmed with a soap bubble gas ow meter each time. The lled bags were then connected to a gas-analyser, which would pump the contained air at the rate required by the instrument, which was substantially higher than the feed rate used in the columns. Extremely consistent reading of the feed air oxygen concentration conrmed consistent operation of the analyser. All experiments were carried out in triplicate. Average gas ow rates in and out were determined for each temperature and an average reading of the gas-analyser was used to determine the oxygen concentration in the off-gas at each temperature. This information was then used to calculate the volumetric rate of oxygen consumption at each temperature. Fig. 6 shows the volumetric rate of oxygen consumption for the four different columns. The average values, together with their relative standard deviations, are also reected in the rst column of Table 2. As can be seen, none of the experiments suggest that oxygen adsorption rate is a signicant function of temperature and the relative standard deviations are in the order of 37%. There does however appear to be a clear correlation between the average oxygen adsorption rate and packing size distribution. Thus the experimental results support the ndings from the leach studies, that leaching progresses at a more or less constant rate, largely independent of temperature (Fig. 3). 3.2. Theoretical evaluation Before determining values of kLa for this system, it must be realised that the sodium sulphite system is not immediately compara-

508

J. Petersen / Minerals Engineering 23 (2010) 504510

Table 2 Results for oxygen gasliquid mass transfer in column experiments. r O2 measured (mol L1 min1) Geocoat experiments Packing 1 Packing 2 Packing 3 Kennecot ore 0.000860 4% 0.000801 3% 0.000742 5% 0.000619 7% r O2 equivalent at 20 C (mg L1 h1) 333.3 (estim.) 366.85 341.82 318.72 264.22 kLa at 20 C (h1) 42.57 46.85 43.65 40.71 33.74

where DiS is the diffusion coefcient of species i in solvent S, and kr the rst order reaction constant describing the consumption of i due to chemical reaction. Above equation is easily solved numerically by discretising as follows and solving by continuous substitution:

 t  t 1 c 2ct ct ct t x c x 1 x x D x 1 kcx 2 Dt Dx

19

1.0

0.8 0.6 0.4 0.2 0.0 0 0.2 0.4 0.6 0.8 1

Norm. distance from gas liquid interface


Fig. 7. Normalised oxygen concentration proles for reacting (curved) and nonreacting lm (linear).

ble to that of an equivalent bio-leaching system. The former case represents gas adsorption into a reacting lm, whereas the latter merely represents diffusion through a lm to a reacting surface. The distinction is illustrated in Fig. 7. Reaction at the surface would result in a linear oxygen concentration prole through the lm, whereas diffusion into a reacting lm would result in depletion of the oxygen some distance into the lm and a non-linear concentration prole. The rate of oxygen transfer into the lm is determined by the oxygen concentration gradient at the gasliquid interface:

r ads f 0

 dc dx x0

16

where the factor f0 relates exclusively to mass transfer properties of the lm uid. As Fig. 7 clearly shows, the surface gradient is much steeper for the reacting lm than for the diffusing lm, and hence one would expect a substantially larger rate of oxygen mass transfer into such a reacting lm, even if all other transport properties are similar. A value for kLa for the bio-leaching system can be determined from the measured gas adsorption rate for the sodium sulphite system only if one has knowledge of the expected ratio between reacting and non-reacting surface gradients as indicated in Fig. 5, i.e.:

r ads;r kL ac cb



 0   dc dc Ha dx r dx nr theor:

17

Eq. (19) was solved using parameters relevant to the current system to mimic (a) diffusion only, (b) diffusion to a reacting surface under the lm, (c) diffusion through a mildly reacting lm to a strongly reacting surface, and (d) diffusion into a reacting lm. In each case the initial concentration of dissolved oxygen in the lm was set to zero except at the gas liquid surface, where it was set to unity, and the system solved until it reached steady state (no change of prole with time). Values chosen for the reaction constant were 0.01 s1 in scenario c and 0.065 s1 (Eq. (15) at 25 C) in scenario d. The reaction constant at the surface was arbitrarily set to a large value (0.5 s1) to ensure rapid consumption. The diffusivity of oxygen in water was taken as 2.5 109 m2 s1 (Perry and Green, 1984). The results are represented in Fig. 8ad. Fig. 8a, representing diffusion only, shows that the lm gradually becomes fully saturated with oxygen as would be expected. Fig. 8b shows the linear diffusion prole through a non-reacting lm to the reacting solid surface as indicated in Fig. 7. By implication, the prole will have to be linear as the rate of diffusion through each layer of the lm will have to be equal. It was on the basis of this scenario that values for kLa had been calculated from mineral leaching data. Fig. 8c, which includes a mildly reacting lm, differs from the previous scenario, with the steady state prole being curved. This would represent a system where some oating bacteria might also consume oxygen, even though the bulk will sit a the mineral surface. Fig. 8d represents diffusion into a rapidly reacting lm as would be the case for the sodium sulphite system, with the value for kr chosen identical to that measured from the batch experiments. Clearly, this system is substantially different from the one represented in Fig. 8b, where bacteria react primarily at the mineral surface. In this case all oxygen is depleted already half way between gasliquid and mineral surfaces. For the rapidly reacting system (Fig. 8d) the slopes at the gas liquid interface have been evaluated for a range of temperature, acknowledging the change with temperature of diffusivity (determined using the WilkeChang relationship, Poling et al., 2001) and reaction constant (Eq. (15)). The results are reected in Table 3. The values for the surface slope essentially correspond to the Hatta number in Eq. (17), as the surface slope for the non-reacting system in this normalised representation would be equal to 1. This analysis shows that the Hatta number is a very weak function of temperature and should thus stipulate a more or less constant rate of oxygen adsorption, regardless of temperature. This again conrms the observations from the leach and gas adsorption experiments (Figs. 3 and 6). 3.3. Determination of kLa in column experiments The volumetric oxygen adsorption rates for the different column experiments shown in Fig. 5 were now divided by 4.50 the Hatta number at 20 C (Table 3) to obtain the equivalent rate of adsorption for a diffusion only system (Eq. (17)). From this a value for kLa is determined using Eq. (2), assuming a lm oxygen gradient of 0.9 c = 7.83 mg L1, similar to what was assumed in the initial analysis. The results of this analysis are included in Table 2. As can be seen, the results are remarkably close to those obtained from the experimental observations in the Geocoat stud-

where the indices r and nr indicate the reacting (sodium sulphite) and non-reacting (bio-oxidation) systems, respectively. This ratio as is dened here is effectively identical to the Hatta number Ha. (Bird et al., 2002). This ratio can be obtained from a theoretical analysis of diffusionreaction through a liquid lm: The linear mass transportreaction equation of a small concentration of a dissolved species diffusing through a solvent and reacting according to rst order kinetics can be described as follows:

Norm. concentration

dci d ci DiS 2 kr ci dt dx

18

J. Petersen / Minerals Engineering 23 (2010) 504510

509

(a)
Norm. gas concentration

1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(b) 1.0
0.9

Norm. gas concentration

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Norm. Distance fom gas-liquid interface

Norm. Distance fom gas-liquid interface

(c) 1.0
Norm. gas concentration
0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(d) 1.0
0.9

Norm. gas concentration

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Norm. Distance fom gas-liquid interface

Norm. Distance fom gas-liquid interface

Fig. 8. Plots of normalised gas concentration proles for the different scenarios described in the text.

Table 3 Values of the theoretical surface slope in the normalised model system as a function of temperature. The surface slope effectively equals the Hatta number. T (C) 20 25 30 35 40 45 50 55 60 65 70 DO2 (109 m2 s1) 2.18 2.50 2.84 3.20 3.59 4.00 4.43 4.89 5.37 5.87 6.40 k0 (s1) 0.057 0.065 0.073 0.082 0.092 0.102 0.114 0.126 0.139 0.153 0.168 dc
dx x0

()

4.500 4.490 4.468 4.462 4.462 4.452 4.470 4.472 4.482 4.496 4.510

ies (42.6 h1, Eq. (12)). There appears to be an effect of particle size distribution of the packing material on the value of kLa in the present study: there is a tendency for materials with a larger portion of ne (<1000 lm) material to exhibit a higher overall kLa. This is not unexpected as one would assume the ner material to provide a larger gasliquid surface area per unit volume (the parameter a in kLa) across which mass transfer can take place. 4. Conclusions The present study has shown, through both experiment and theoretical evaluation, that gasliquid mass transfer of oxygen in heap leach scenarios is remarkably constant between different systems and is not a signicant function of temperature in the range 2268 C. The latter is surprising, given the rapidly declining solubility of oxygen with increased temperature in this range. The

declining solubility is offset by a proportional increase in the value of the gasliquid mass transfer coefcient kLa, primarily because of the increased diffusivity of oxygen in water. Experimentation with different size distributions in the packed material indicated a tendency towards increased mass transfer with materials with a higher nes content, probably due to the likely increase in exposed surface area within the bed. The results indicate that heap bio-leaching as a process is determined by the rate of gasliquid mass transfer of oxygen into solution in all the cases studied, which cover different materials, temperatures and microbial consortia. While operating heaps at low temperatures may be limited even more by slow mineral kinetics, thermophilic heap bio-leaching certainly cannot progress any faster than the rate of oxygen gasliquid mass transfer allows, even though it may be the preferred route for chalcopyrite leaching. Operating heaps at high altitude and in highly saline solutions further limit the rate of oxygen mass transfer and hence the overall rate of metal recovery. Acknowledgements This work was conducted in the context of the AMIRA P768A project and the author gratefully acknowledges their sponsorship. The author would also like to acknowledge the contributions to the experimental portion of this work made by Emmanuel Ngoma, Thierry Kamunga Kazadi and Tracy Reddy. Geocoat is a trade mark process of GeoBiotics LLC of Lakewood, CO, USA. References
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2002. Transport Phenomena, second ed. John Wiley & Sons.

510

J. Petersen / Minerals Engineering 23 (2010) 504510 Perry, R.H., Green, D. (Eds.), 1984. Perrys Chemical Engineers Handbook, sixth ed. McGraw-Hill. Poling, B., Prausnitz, J., OConnell, J., 2001. The Properties of Gases and Liquids. McGraw-Hill, New York. Ojumu, T.V., Petersen, J., Searby, G.E., Hansford, G.S., 2006. A review of rate equations proposed for microbial ferrous-iron oxidation with a view to application to heap bioleaching. Hydrometallurgy 83 (14), 2128. Searby, G., Hansford, G.S., 2004. In: Tzesos, M., Hatzikioseyian, A., Remoudaki, E. (Eds.), Biohydrometallurgy: A Sustainable Technology in Evolution Part II. National Technical University of Athens, Greece, pp. 12271236. Tromans, D., 2000. Modeling oxygen solubility in water and electrolyte solutions. Ind. Eng. Chem. Res. 39 (3), 805812.

Cooper, C.M., Fernstrom, G.A., Miller, S.A., 1944. Performance of agitated gasliquid contactors. Ind. Eng. Chem. 36 (6), 504509. Doran, P.M., 1995. Bioprocess Engineering Principles. Academic Press. Hansford, G.S., Vargas, T., 2001. Chemical and electrochemical basis of bioleaching processes. Hydrometallurgy 59, 135145. Petersen, J., Dixon, D.G., 2002. Thermophilic heap leaching of a chalcopyrite concentrate. Miner. Eng. 15 (11), 777785. Petersen, J., Dixon, D.G., 2006a. Modeling and optimisation of heap bioleach processes. In: Rawlings, D.E., Johnson, D.B. (Eds.), Biomining. Springer Verlag, Berlin, pp. 153176. Petersen, J., Dixon, D.G., 2006b. Competitive bioleaching of pyrite and chalcopyrite. Hydrometallurgy 83 (14), 4049.

You might also like