Você está na página 1de 5

Hydrometallurgy 178 (2018) 7–11

Contents lists available at ScienceDirect

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

The kinetics of the dissolution of chrysocolla in acid solutions T



Michael J. Nicol , Chandrika Akilan
School of Engineering and Information Technology, Murdoch University, Perth, WA 6150, Australia

A R T I C LE I N FO A B S T R A C T

Keywords: The dissolution of liberated chrysocolla particles in dilute (0.033 to 0.15 mol L−1) sulfuric acid solutions has
Chrysocolla been shown to be relatively rapid and the kinetics of dissolution are consistent with rate-determining diffusion of
Leach acid through the silica layer that is formed around the particles from which the copper has been dissolved. The
Acid rate is inversely proportional to the square of the particle size and increases linearly with increasing acidity. The
Ash-layer
effect of the addition of chloride ions on the rate is negligible at concentrations below 70 g/L chloride. Addition
Diffusion
of sulphate ions results in reduced rates of dissolution due to complexation of the proton by sulphate ions. The
effect of temperature is relatively small as expected for a process for which the rate is controlled by diffusion.
The rate of dissolution in dilute acid is slower than malachite but faster than tenorite under the same con-
ditions. These results suggest that heap leaching of chrysocolla in dilute acid may be slow relative to malachite
but should present no chemical problems and that there is no advantage to be gained from the addition of
chloride ions. High sulphate concentrations (100 g/L) in the raffinate solution to the heap could reduce the rate
by up to one-half.

1. Introduction dissolution of relatively pure chrysocolla particles under conditions


similar to those encountered during heap leaching of oxide or mixed
Chrysocolla is a copper silicate mineral with the nominal formula oxide/sulfide ores.
CuSiO3.xH2O although the actual formula is accepted to be
Cu8(Si4O10)2(OH)12.xH2O in which x is normally 8. It is considered to 2. Thermodynamics
be a gel-like substance with a disordered array of microcrystallites
(Pohlman and Olson, 1976). It has a very high surface area of greater The species distribution for the dissolution of chrysocolla in acidic
than 250 m2/g that reduces to below 25 m2/g after acid leaching sulfate solutions is shown in Fig. 1. This was obtained using HSC
(Pohlman and Olson, 1976). Visual observation of a large particle Chemistry Roine (2015) with the free energy of formation of chryso-
during dissolution showed that the pale blue surface changes to a white colla (−1220.8 kJ mol−1 for CuSiO3·2H2O) obtained from the literature
hydrated silica outer layer (similar to silica gel) with the inner blue core (Wei et al., 2010).
gradually decreasing in diameter as leaching progresses. It can be seen that chrysocolla is completely soluble at pH values
It is known to be easily soluble in dilute acids but very few studies below about 3 and completely insoluble at pH values above about 6.
(Hsu and Murr, 1975; Pohlman and Olson, 1974) have focussed on the Thus, there are no thermodynamic restrictions on the dissolution of
kinetics of dissolution process and the dependence of the rate on par- chrysocolla under acidic conditions.
ticle size, temperature and acid and chloride concentrations while most
(Bryce et al., 1968) have used low grade ores with the resulting pro- 3. Experimental
blems of mass transport to the mineral surfaces within larger ore par-
ticles. Natural chrysocolla inevitably occurs as intergrowths with quartz
Dissolution in acid occurs by the following reaction and it is difficult to obtain sufficient material of high grade. Small
lumps of chrysocolla from Arizona were sourced locally and XRD ana-
CuSiO3 . xH2 O + 2H+ = Cu2 + + SiO2 . H2 O + xH2 O (1)
lysis provided little confirmatory evidence as expected from the lit-
in which hydrated silica forms as a product layer on the surface of the erature (Pohlman and Olson, 1976) that reported either none or very
mineral. poor crystallinity or crystallinity of colloidal size. The XRD scan
The following is a summary of a study of the kinetics of the obained with the present sample using Cu radiation is compared with a


Corresponding author.
E-mail address: m.nicol@murdoch.edu.au (M.J. Nicol).

https://doi.org/10.1016/j.hydromet.2018.04.001
Received 9 January 2018; Received in revised form 22 March 2018; Accepted 1 April 2018
Available online 04 April 2018
0304-386X/ © 2018 Elsevier B.V. All rights reserved.
M.J. Nicol, C. Akilan Hydrometallurgy 178 (2018) 7–11

and accurate than AA or ICP analysis at these concentrations.

4. Results

As will be demonstrated below, the rate of the dissolution of chry-


socolla particles can be very well described in terms of rate-determining
diffusion of protons through the amorphous silica product layer.
This well known “ash layer” model (Prosser, 1996) for the dis-
solution of a particle with a product layer results in the following ex-
pression for the extent of dissolution(X) as a function of time(t)
t/(3τ) = 1 − 2X/3 − (1 − X)2/3 (2)
in which τ is the time for complete dissolution.
τ = ρR2/(6bCD) (3)

Fig. 1. Species distribution for chrysocolla in acidic sulfate solutions at 25 °C. in which b is a stoichiometric factor (in this case 1 mol copper dis-
solved/mol sulfuric acid consumed), C is the concentration
(mole cm−3) of the species diffusing into or out of the product layer and
D is the diffusion coefficient (cm2 s−1) in the product layer.
Fig. 3 shows the dissolution curves for various size fractions in
0.1 mol L−1 sulfuric acid.
It is obvious that the initial dissolution rate is rapid but it slows
noticeably after about 60% dissolution and that the rate decreases as
the particle size increases. Also shown (as square points and right axis)
are the corresponding plots according to Eq. (2). The good linearity of
the plots up to 80% extraction suggests that the data is consistent with
the assumed rate-determining step.
As a further test of the rate equation, the values of 3 x slope (=1/τ)
of these lines are plotted in Fig. 4 as a function of the reciprocal of the
square of the geometric mean diameter (d) of the screen sizes defining
each size fraction as per Eq. (3).
Fig. 2. Photograph of a group of particles from the +500–710 μm size fraction
Again, the adequate linearity (except for the point for the smallest
before and after leaching.
size) confirms the applicability of the leaching model. From the slope of
the line in Fig. 4, the diffusion coefficient of the proton in the hydrated
published scan in Appendix A. The only impurity detected by XRD was silica layer (density 2.21 g cm−3) can be estimated as
quartz. The lumps were hand-crushed in a pestle and mortar and sev- 7.8 × 10−6 cm2 s−1 that is remarkably similar to the measured diffu-
eral size fractions from 180 μm to 850 μm obtained by dry screening the sion coefficient of HCl in silica gel of 8.5 × 10−6 cm2 s−1 (Patrick and
crushed material. Fig. 2 shows a number of such particles from the Allan, 1934). These values are about an order of magnitude lower than
+500 to 710 μm size fraction. the value in the solution phase (9.3 × 10−5 cm2 s−1 at infinite dilution)
Examination of the particles shows that the chrysocolla (pale blue/ (Cussler, 2011), confirming that diffusion in the product layer is rate-
green) is generally liberated from the quartz (colourless) and other determining.
minor dark gangue minerals. The composition in terms of chrysocolla The results (in terms of the time for complete dissolution) as ob-
content increased with decreasing particle size due to preferential tained from plots similar to those shown in Fig. 3 for experiments on the
crushing of the softer copper mineral. The copper content decreased effect of the acid concentration on the rate of dissolution of the
from 17.8% to 11.7% from the finest to coarsest fraction. +250–355 μm size fraction at 25 °C are summarized in Fig. 5.
After leaching, the converted chrysocolla appears as amorphous It can be seen that the rate increases linearly with increasing acidity
cloudy, white particles easily distinguished visually from the original as would be required for diffusion (a first-order process) as the rate
crystalline quartz. There does not appear to be a noticeable size change controlling step.
during dissolution. The effect of the addition of chloride (as sodium chloride) on the
The leach tests were conducted in a stirred, temperature-controlled rate was studied using the +355–500 μm size fraction in a solution of
baffled reactor as described previously (L Velásquez-Yévenes et al., 0.1 mol L−1 acid at 25 °C. The results are summarized in Table 1.
2010). A small glass tube heat exchanger was fitted to the reactor It is apparent that the rate is independent of the chloride con-
through which chilled water from a low-temperature water bath was centration in the range studied as could be expected if the rate is con-
circulated. Most experiments were conducted at 25 °C at a stirring rate trolled by the diffusion of acid within the silica product layer.
of 580 rev min−1. Agitation at 400 rev min−1 was found to be required Addition of excess sulfate ions (as magnesium sulfate) could be
to maintain the particles in suspension. Unless stated otherwise, 600 mL expected to reduce the rate due to the formation of hydrogen sulfate
of a leach solution of the appropriate composition was added to the ions thereby reducing the proton concentration. The results of such
reactor and, after the temperature had stabilized, 1 g of the desired size experiments with the +250–355 μm size fraction in 0.1 mol L−1 acid at
fraction was added to the reactor. 2 mL samples were periodically re- 25 °C are shown in Table 2.
moved after allowing the suspension to settle for 20 s before sampling These results confirm the above expectation and also shown in
the solution. Table 2 are the calculated proton concentrations using thermodynamic
The solution samples were added to 5 mL of 2 M ammonium data (Martell and Smith, 2004) for the protonation of sulfate and the
chloride.ammonium hydroxide at pH 10 and the volume made up to formation of magesium sulfate ion-pairs at various ionic strengths. The
10 mL. The absorbance of the solution at 615 nm was measure in a approximate correlation between the rate of dissolution and the proton
20 mm quartz cell using a Shimadzu spectrophotometer. A standard concentration is consistent with the rate-determining step being the
solution of copper was used for calibration. This method is more rapid diffusion of protons through the product layer. It is likely that

8
M.J. Nicol, C. Akilan Hydrometallurgy 178 (2018) 7–11

Fig. 3. Effect of particle size on the dissolution of chrysocolla in 0.1 mol L−1 acid at 25 °C.

Fig. 5. Effect of acidity on the rate of dissolution of the +250–355 μm size


Fig. 4. Effect of particle size on the slopes of the lines in Fig. 2.
fraction of chrysocolla at 25 °C.

9
M.J. Nicol, C. Akilan Hydrometallurgy 178 (2018) 7–11

Table 1
Effect of chloride on the rate of dissolution.
[NaCl] M 1/τ min−1

0 0.0049
0.2 0.0055
0.5 0.0051
2 0.0051

Table 2
Effect of sulfate on the rate of dissolution.
[Sulfate] mol L−1 1/τ min−1 [H+] calc mol L−1

0.1 0.0102 0.046


0.3 0.0086 0.034 Fig. 6. Effect of temperature on the rate of dissolution of the +355–500 um
0.5 0.0063 0.028 fraction in 0.1 mol L−1 H2SO4.

Table 4
Table 3 Initial rates of dissolution in 0.1 mol L−1 H2SO4 at 25 °C.
Effect of temperature on the rate of dissolution. Mineral Initial rate mol cm−2 min−1
° −1
Temp C 1/τ min
Malachite 2.4 × 10−6
20 0.0042 Chrysocolla 1.3 × 10−6
25 0.0049 Tenoritea 2.5 × 10−7
30 0.0056
35 0.0068
a
0.2 M H2SO4 – rotating disk.

These results confirm unpublished reports that the rate of chrysocolla


dissolution will occur by reaction with the HSO4− ion as well as the leaching in heaps is slower than that of malachite. The absence of
proton. However, the diffusion coefficient of the bisulfate ion is an significant quantities of tenorite in most copper ores is fortunate in that
order of magnitude less than that of the proton and the contribution of the rate of leaching could be expected to be an order of magnitude
bisulfate to the dissolution will be comparable to that of the proton only slower than the other copper “oxide” minerals.
at low proton concentrations.
The effect of temperature was studied using the +355–500 μm size
fraction in 0.1 mol L−1 acid and the results shown in Table 3 and as an 5. Conclusions
Arrhenius plot in Fig. 6.
The activation energy calculated from the slope of the plot is The dissolution of screened samples of liberated chrysocolla parti-
23.4 kJ mol−1. This low value is as expected for a diffusion controlled cles in dilute (0.033 to 0.15 mol L−1) sulfuric acid solutions has been
reaction (Cussler, 2011). shown to be relatively rapid and the kinetics of dissolution are con-
It is interesting to compare the rate of dissolution of chrysocolla sistent with rate-determining diffusion of acid through the silica layer
with that of malachite (Cu2(OH)2CO3) (Nicol, 2018) and tenorite (CuO) that is formed around the particles from which the copper has been
(Majima et al., 1980) for which the rate-determining step is the rate of dissolved. The rate is inversely proportional to the square of the particle
the chemical reaction on the surface of the minerals (no product layer is size and increases linearly with increasing acidity. The effect of the
formed in the case of malachite or tenorite). Previously obtained data addition of chloride ions on the rate is negligible at concentrations
for the dissolution of malachite under identical conditions showed that below 70 g/L chloride. Addition of sulphate ions (as magnesium sul-
the time for complete dissolution is less for malachite than chrysocolla. phate) results in reduced rates of dissolution due to complexation of the
The relative rates are compared in Table 4. proton by sulphate ions. The effect of temperature is relatively small as
Note that, in the case of chrysocolla, the initial rate using Eq. (3) is expected for a process for which the rate is controlled by diffusion.
infinite but will actually be the rate of the chemical reaction in the The rate of dissolution in dilute acid is slower than malachite but
absence of diffusion limitations. An arbitrary choice of X = 0.1 has faster than tenorite under the same conditions. These results suggest
been chosen to estimate this rate of the chemical reaction. that heap leaching of chrysocolla in dilute acid may be slow relative to
As a result of the difference in the form of the rate equations, the malachite but should present no chemical problems and that there is no
rate of chrysocolla dissolution decreases more rapidly at high extents of advantage to be gained from the addition of chloride ions. High sul-
reaction than does that for malachite or tenorite. Thus, the time for 80% phate concentrations (100 g/L) in the raffinate solution to the heap will
dissolution under these conditions is 3.5 times longer for chrysocolla reduce the rate by up to one-half.
than malachite although it is only 2.1 times longer for 50% dissolution.

Appendix A. Appendix

A published (University of Arizona, 2018) XRD scan of chrysocolla (Fig. A.1) is compared to that obtained with the sample used in this study (Fig.
A.2). The former was characterized as having broad, low peaks with weak lines often diffuse. The pattern is fairly characteristic. The similarity in
terms of very broad lines for the present sample provides confirmatory identification of the present sample as chrysocolla. Note the sharp peaks in the
present sample (Fig. A.2) due to quartz as an impurity.

10
M.J. Nicol, C. Akilan Hydrometallurgy 178 (2018) 7–11

Fig. A.1. XRD scan of chrysocolla courtesy of RRUFF project at University of Arizona (2018)

Fig. A.2. XRD scan for the sample of chrysocolla used in the present study.

References 38, 771–778.


Pohlman, S.L., Olson, F.A., 1974. Kinetic study of acid leaching of chrysocolla using a
weight loss technique. In: Aplan, F.F., McKinney, W.A., Pernichele, A.D. (Eds.),
Bryce, D.M., Cerigo, D.G., Jennings, P.H., 1968. Percolation and agitation leaching of an Solution Min. Symp., Proc. pp. 446–460.
oxidized copper ore. CIM Bull. 61 (673), 641–645. Pohlman, S.L., Olson, F.A., 1976. Characteristics of chrysocolla pertinent to acid leaching.
Cussler, E.L., 2011. Diffusion, 3rd ed. Cambridge University Press. In: Yannopoulos, J.C., Agarwal, J.C. (Eds.), Proc. Int. Symp. on Copper. A.I.M.E., Las
Hsu, P.C., Murr, L.E., 1975. Simple kinetic model for sulfuric acid leaching of copper from Vegas (Chapter 48).
chrysocolla. Metall. Trans. B 6B (3), 435–440. Prosser, A.P., 1996. Review of uncertainty in the collection and interpretation of leaching
L Velásquez-Yévenes, L., Nicol, M.J., Miki, H., 2010. The dissolution of chalcopyrite in data. Hydrometallurgy 41, 119–153.
chloride solutions. part 1. effect of solution potential. Hydrometallurgy 103, Roine, A., 2015. Outokumpu HSC Chemistry for Windows, Version 7.1. Outokumpu
108–113. Research of Finland.
Majima, H., Awakura, Y., Yazaki, T., Chikamori, Y., 1980. Acid dissolution of cupric University of Arizona, 2018. RRUFF Database, Chrysocolla R050053. uamineralmuseum.
oxide. Metall. Trans. B 11B, 209–214. org.
Martell, A.E., Smith, R.M., 2004. NIST Standard Reference Database 46, Ver 8. National Wei, L., Motang, T., Zhaobo, T., Jing, H., Jianguang, Y., Shenghai, Y., 2010.
Institute of Standards and Technology, Gaithersburg, MD, USA. Thermodynamic Research of the Dissolving of Chrysocolla (CuSiO3·H2O) in the
Nicol, M.J., 2018. The kinetics of the dissolution of malachite in acid solutions. Ammonia-Ammonium Chloride-Water System. EPD Congress 2010, Miner. Met.
Hydrometallurgy 177, 214–217. Mater. Soc. pp. 327–334.
Patrick, W.A., Allan, B.W., 1934. The diffusion of electrolytes in silica gel. J. Phys. Chem.

11

Você também pode gostar