You are on page 1of 7

Journal of Colloid and Interface Science 282 (2005) 408–414

www.elsevier.com/locate/jcis

Fundamental considerations on the mechanisms of silver cementation


onto zinc particles in the Merril–Crowe process
G. Viramontes Gamboa a , M. Medina Noyola a , A. López Valdivieso b,∗
a Instituto de Física, Universidad Autónoma de San Luis Potosí, Av. Manuel Nava No. 6, San Luis Potosí, S.L.P., C.P. 78210, Mexico
b Instituto de Metalurgia, Universidad Autónoma de San Luis Potosí, Av. Sierra Leona No. 550, San Luis Potosí, S.L.P., C.P. 78210, Mexico

Received 8 June 2004; accepted 17 August 2004


Available online 11 November 2004

Abstract
Studies on the Merrill–Crowe process as applied to silver recovery have shown that one half of the used zinc powder is wasted in water
reduction at high cyanide concentrations, while the other half reduces silver ions from the cyanide solution. However, the cementation
mechanisms as an electrochemical process taking place on the zinc surface do not explain the split of the electric current resulting from
the anodic dissolution of zinc into two equal values. This study demonstrates that the mechanism for silver precipitation at high and low
cyanide concentrations differs considerably. At low cyanide concentrations cementation is essentially an electrochemically-controlled process
following a shrinking-core behavior. At high cyanide concentrations, the process seems not to be electrochemically controlled. The areas for
zinc dissolution and silver deposition are not connected by an electrical-conducting medium and reduction of silver–cyano complex ions
takes place by hydrogen adsorbed onto silver growing outward from the cementing zinc particles. The results are based on scanning electron
microscopy of solids recovered from cementations in stirred reactors and in situ observations by optical microscopy of the cementation
process on the edge of thin zinc disks in cyanide solutions.
 2004 Elsevier Inc. All rights reserved.

Keywords: Silver; Metal cementation; Cyanide leaching; Zinc; Mineral processing

1. Introduction theoretically has a maximum value ε = 2, since one mole


of zinc reduces two of Ag(CN)− 2 . However, in silver recov-
The Merril–Crowe process for recovering gold and silver ery plants ε has a value around or lower than 1 [7,10,11].
from cyanide-leached ores has been widely used throughout In laboratory experiments the effects of impurities may be
the entire world since its introduction at the end of 19th cen- controlled, and it has been shown that ε = 1 for high free
tury. The acquisition of zinc powder represents one of the cyanide concentrations, like those commonly used in plant
main investments for the plants that use this process, a suit- practice [1,12]. This means that water reduction accounts for
able optimization of its consumption being clearly desirable. one half of the used zinc, which is true under various exper-
Undesirable waste of zinc has its origin mainly in three imental conditions, like different silver and cyanide concen-
trations and lead ions in solution. Understanding the origin
causes: (a) unconsumed zinc left in the precipitate, (b) im-
of the ε = 1 relation has fundamental importance and con-
purities precipitation, and (c) other zinc-consuming side re-
stitutes the aim of this paper.
actions. Together, these effects diminish zinc efficiency to
It is generally accepted that the main chemical reactions
recover silver drastically. The cementation efficiency ε, de-
involved in the cementation of silver from cyanide solu-
fined as the molar ratio of cemented silver to dissolved zinc,
tions using metallic zinc powder are of an electrochemical
nature. This means that the respective half reactions are
* Corresponding author. connected by an electrically-conducting medium, through
E-mail address: alopez@uaslp.mx (A. López Valdivieso). which an electric current is established transporting elec-
0021-9797/$ – see front matter  2004 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2004.08.151
G. Viramontes Gamboa et al. / Journal of Colloid and Interface Science 282 (2005) 408–414 409

the rates for both reactions are the same, instead of (3), the
overall silver cementation reaction should be
Zn0 + Ag(CN)−
2 + H2 O + 2CN

− 1
→ Ag0 + Zn(CN)2−
4 + OH + H2 . (6)
2
For this reaction ε = 1, since one mole of zinc precipi-
tates one of silver. This ε value and Reaction (6) have been
verified not only in controlled laboratory experiments at
high cyanide concentrations [1,12], but also in plant oper-
ations [7,9,11].
According to the traditional view of silver cementation
Fig. 1. Schematic representation of the electrochemical behavior of silver as an electrochemical process, the decrease in ε from its
cementation onto zinc in the Merril–Crowe process. maximum value of 2 to 1 seems to be a subject easy to be
understood. It would appear that the cause is a gradual in-
crease in water reduction as cyanide concentration increases,
trons from the zinc oxidation sites up to the silver reduc- taking place simultaneously in the electrochemical process.
tion sites. A schematic picture of this process, as shown in Despite this apparent simplicity the issue is much more com-
Fig. 1, is included in most scientific reports dealing with the plex, since it is necessary to point out that silver and water
Merril–Crowe method to illustrate the cementation process. reduction rates should be of the same order for Reaction (6)
Accordingly, electrochemical techniques have been exten- to be valid. Taking into account that this is true for conditions
sively used to delineate the mechanisms of precious metal as varied as those in plant operations, there is no reason for
cementation in order to optimize plant operations [2–6,8]. both rates to have similar values. In fact, both rates may be
In the Merril–Crowe process as applied to silver recovery completely different, as shown by the studies on the cemen-
from cyanide solutions, the zinc oxidation and silver reduc- tation of precious metals from cyanide solutions using Evans
tion reactions are diagrams by Parga [6]. This researcher did not detect hydro-
gen evolution in the cementation of silver, which occurred
Zn0 + 4CN− → Zn(CN)2− −
4 + 2e , (1)
in gold cementation as reported elsewhere [2,3,5,6]. It can
Ag(CN)− − −
2 + e → Ag + 2CN ,
0 (2) be stated that equality between both rates is not the rule,
but the exception, as deduced from electrochemical stud-
respectively. If side reactions are absent the final stoichiom- ies [2,3,5,6]. A congruent explanation of the origin of ε = 1
etry is is really absent in the literature.
This work aims at thoroughly understanding the mecha-
Zn0 + 2Ag(CN)− 0 2−
2 → 2Ag + Zn(CN)4 . (3) nism of the transition between Reactions (3) and (6) through
The cementation efficiency for Reaction (3) is ε = 2, the the- two different kinds of experiments. The first one was devised
oretical maximum. This value has been found to be true only to model the Merril–Crowe process through cementation
at low free cyanide concentrations, however [1]. As pointed tests in stirred reactors. The second was aimed at explicitly
out above, ε is always closer to 1 than 2 in plant operation. analyzing the validity of an electrochemical behavior for sil-
Besides impurities cementation, the decrease in ε has been ver precipitation onto zinc at high cyanide concentrations.
assigned to the simultaneous and wasteful consumption of
electrons by the following side reactions [9]:
2. Experimental
O2 + 2H2 O + 4e− → 4OH− , (4)
For all experiments, deionized water and analytical grade
1
H2 O + e− → OH− + H2 , (5) AgNO3 , NaCN, and NaOH reagents were used to prepare
2 alkaline silver–cyanide solution at initial pH 10.8. The so-
Reaction (4) is minimized in Merril–Crowe plants by lutions were purged with oxygen-free highly pure N2 to
deaeration of the pregnant solution prior to the zinc powder remove O2 prior to the zinc addition and along all the ce-
addition. Reaction (5) is less manageable since it depends on mentation period.
the free cyanide concentration, which is determined by the In the first set of experiments, zinc powder cementation
leaching conditions. Water reduction was recognized as the tests were carried out in a 2.5 L glass reaction vessel im-
most electron consuming side reaction soon after the imple- mersed in a temperature-controlled water bath. The zinc was
mentation of the Merril–Crowe method [10,11]. Ever since it added when the O2 concentration was lower than 0.2 mg/L.
has been the main explanation for the loss in zinc cementa- After 1 h the cementation products were collected through
tion efficiency. In adequately deaerated pregnant solutions, filtration with 0.22-µm pore size filters and observed and an-
Eqs. (2) and (5) are the main reduction reactions. Thus, if alyzed using a Phillips XL 30 scanning electron microscope
410 G. Viramontes Gamboa et al. / Journal of Colloid and Interface Science 282 (2005) 408–414

Fig. 2. (a) Schematic view of the zinc disk. (b) Two-dimension-like system
with epoxy resin to avoid the precipitation of silver on the zinc disk faces.

with accessories for chemical analysis. The solutions were


Fig. 3. Silver cementation efficiency ε as a function of cyanide concentra-
assayed for zinc and silver by atomic absorption spectropho- tion.
tometry.
The second set of experiments consisted of the cemen-
tations of silver in a two-dimension-like system. They were diate cyanide concentration regime at [NaCN]0 /[Ag]0 ratios
devised to elucidate if the areas of zinc dissolution and sil- between the two previous values, where ε decreases gradu-
ver precipitation are or are not continuously connected by an ally from 2 to 1. The ε behavior in the figure does not depend
electrically-conducting medium. Fig. 2 shows schematically on the zinc particle size and remains unaltered if lead ions
a thin disk of metallic zinc whose faces were covered with an are present in the solution. According to this result, around
insulating transparent epoxy resin. Only the zinc disk edge one half of the electrons supplied by Reaction (1) is wasted
was exposed to the silver–cyanide solution. Solutions were in water reduction in the high cyanide regime. The other half
prepared with initial AgNO3 and NaCN concentrations of reduces silver according to Reaction (2).
3.7 mM and 0.111 M, respectively. After O2 removal the Fig. 4 shows the cementation patterns at different cyanide
silver–cyanide solution was continuously pumped, through concentrations. In the low cyanide regime the silver cemen-
a hose system, to fill a sealed Petri box containing the tation follows a shrinking core behavior around the zinc
zinc disk. The cementation evolution was directly observed particles (Fig. 4a). In contrast, in the high cyanide regime,
with a transmission optical microscope and recorded with where ε ≈ 1, the cemented silver always grows outward, into
a coupled video-microscopy system. For silver and cyanide the aqueous solution (Fig. 4b). In addition, if free cyanide
concentrations used in these experiments, the cemented sil- concentration is sufficiently high, cemented silver develops
ver grows radially away from the disk edge while the zinc branched morphologies independently of zinc size (Fig. 4c).
disk shrinks. The formation of hollow spaces separating sil- In the high cyanide regimen zinc particles progressively re-
ver and zinc was manifested as bright areas in this two- duce their size inside a hollow space, which has exactly the
dimension-like system. The formation and subsequent evo- same shape as that of the original zinc particle. In the cyanide
lution of a continuous bright hoop demonstrates that silver regime where ε changes from 2 to 1 mixed behavior may be
can be cemented in the absence of physical contact with detected.
metallic zinc. A BUEHLER EPO-THIN low viscosity epoxy To demonstrate that silver outward growth began instan-
resin No. 20-8140-032 and hardener No. 20-8142-016 were taneously upon zinc addition at high cyanide conditions,
used at a weight proportion of 100 parts resin to 36 parts cementation tests with zinc particles having a rough tex-
hardener, cured under room conditions. The electrical resis- ture on its surface were carried out (Fig. 5a). Fig. 5b shows
tivity of the cured product is ρ = 1.0–1.2 × 1016  cm−1 . the characteristic texture of the inner face of the silver shell
cemented with this zinc. Comparison of both images demon-
strates that the initial texture and shape of the zinc particles
3. Results and discussions is printed on the inner side of the silver shells. All analyzed
particles follow this pattern. Because the shape of the in-
Fig. 3 shows the ε dependence on the cyanide concentra- ner cemented silver–solution interface remains intact up to
tion, which is indicated as [NaCN]0 /[Ag]0 ; where [NaCN]0 the end of the cementation test, it can be stated that silver
and [Ag]0 are the total initial cyanide and silver concentra- cementation takes place only on the outer cemented silver–
tions in mol/L, respectively. The tests were carried out with solution interface.
[Ag]0 = 0.927 mM and [NaCN]0 was varied. The curve ex- The cementation of silver in the high cyanide regime can
hibits three cyanide concentration regimes: the low cyanide then be pictured as a silver shell growing around a hollow
concentration regime given by [NaCN]0 /[Ag]0 ratios slightly space and having an inner shape and size of the original zinc
larger than 2, where the concentration of free cyanide ions particle. Inside this hollow space the zinc particle progres-
in solution starts to be significant and ε ≈ 2; the high sively reduces its size. Additionally, the hollow space is free
cyanide concentration regime occurring at [NaCN]0 /[Ag]0 of Ag(CN)− 2 complexes. Otherwise they would be cemented
ratios larger than 8, where ε ≈ 1; and finally, the interme- either on the zinc particle or on the inner face of the silver
G. Viramontes Gamboa et al. / Journal of Colloid and Interface Science 282 (2005) 408–414 411

Fig. 5. (a) Characteristic texture of initial zinc particles. (b) Texture of the
inner face of the cemented silver shell at the high cyanide concentration
regime.

high cyanide regimen, where ε = 1, between the zinc parti-


cle and the silver shell there is no physical contact. If this
hypothesis is true, as experimentally shown below, the elec-
trons needed to reduce silver on the outer face of the silver
shell do not come directly from the zinc oxidation areas
located on the zinc surface. Instead, a chemical species is
produced directly on the zinc surface, diffuses through the
solution inside the hollow space and the pores of the silver
shell, and finally reduces silver–cyano complex species at
Fig. 4. Silver cementation patterns. (a) Shrinking-core behavior in the low
cyanide concentration regime. (b) Outward growth at high cyanide concen- the cemented silver–aqueous solution interface.
tration regime. (c) Outward growth with development of dense-branched Direct measurements of the interaction energy between a
silver morphology, also in the high cyanide regime. metallic sphere (zinc) inside another hollow particle (silver
shell) filled with a multiple electrolyte (cyanide solutions)
shell; these possibilities, however, have not been confirmed are really difficult. Confronted with the difficulty of obtain-
experimentally. ing directly this data, the zinc disk experiments were de-
The transition of ε from its maximum value of 2 to 1, signed to explicitly avoid the physical contact between zinc
takes place together with a transition in the cementation and its respective silver shell. The idea was to demonstrate
pattern, from a shrinking core behavior to an outward sil- the truthfulness of the hypothesis by deliberately avoiding
ver growth. This suggests the possibility that the final value that physical contact between cemented silver and the ce-
ε = 1 at high cyanide concentrations is a consequence of a menting zinc particle. If silver cementation stops then the
morphological and stoichiometric transition. From Figs. 4b, physical contact is a necessary condition. Oppositely, if it
4c, and 5, it seems that the zinc particle separates from the continues then the cementation process should take place
silver shell right at the beginning of the cementation reac- following a mechanism distinct from that conventionally ac-
tion. These observations lead to the hypothesis that in the cepted.
412 G. Viramontes Gamboa et al. / Journal of Colloid and Interface Science 282 (2005) 408–414

are H2 bubbles. Because the epoxy resin is an excellent insu-


lator, this experiment demonstrates that the electrochemical
mechanism depicted in Fig. 1 is far from reality in the high
cyanide regime, where ε = 1. The electrochemical view is
only supported for low and intermediate cyanide regimes as
Fig. 4a clearly shows.
Zinc disk experiments demonstrate that silver cementa-
tion in the high cyanide regimen does not necessarily need
the electrical continuity between the areas for zinc dissolu-
tion and silver deposition. The results can only be explained
on the basis of two chemical reactions taking place at dif-
ferent interfaces and coupled only by a transport process
through the hollow space. Even though the chemical identity
of the species that reduces silver on the outer silver–solution
interface has not been determined exactly in this study, there
are reasons to believe that hydrogen is the most likely can-
didate. Hydrogen is produced at the zinc–aqueous solution
interface at such an extent that hydrogen bubbles are directly
observed inside the hollow space of the zinc disk experi-
ments. Thermodynamically, silver reduction with hydrogen
from cyanide solution is not excluded. Gas desorption tests
by pyrolysis, as detected by mass spectrometry, carried out
at high vacuum on the solids recovered from the zinc pow-
der cementations, indicate the codeposition of H, H2 , and
HCN. This possibility should be taken carefully since homo-
geneous reduction of silver from alkaline cyanide solutions
with pressurized H2 has been reported to be unlikely due to
high stability of the hydrogen molecule [13,14]. However, it
is well known that Cu2+ , Ag+ , and Hg2+ can activate mole-
cular hydrogen to make it more reactive as has been pointed
out by Osseo-Asare [15]. Thus, the formation of less alka-
line microenvironments around the cemented silver–cyanide
solution interface may lead to the activation of molecular
hydrogen through the stabilization of low concentrations of
Ag+ . As noted above, the absence of Ag(CN)− 2 molecules
inside the hollow space demonstrates the formation of mi-
croenvironments having properties different from those in
the bulk cyanide solution.
The electrochemical mechanism as shown by Eqs. (1),
(2), and (3) describes adequately silver cementation only in
Fig. 6. (a) Bright hoop separating the zinc disk and the cemented silver. the low cyanide regimen. For the high cyanide regimen, the
(b) Gas bubble detected inside the hollow space. (c) Growth of silver out- mechanism suggested by the above results consists, first of
ward from the bright hoop.
all, of the production of H2 at the zinc–aqueous solution
interface according to the following electrochemical reac-
Silver cementation carried out in two-dimensional sys- tions:
tems like the zinc disk additionally allowed a direct obser-
Zn0 + 4CN− → Zn(CN)2− −
4 + 2e , (7a)
vation of the evolution of the hollow space separating ce-
mented silver and metallic zinc. Fig. 6 shows the result of 2H2 O + 2e− → 2OH− + H2 , (7b)
a cementation experiment on the zinc disk after 48 h. The
Zn0 + 4CN− + 2H2 O → Zn(CN)2−
4 + 2OH− + H2 . (7c)
continuous bright hoop in this 2D system corresponds to
the hollow space in the zinc powder cementations. It is sur- The overall Reaction (7c) increases the pH around the
rounded by silver on the outer side and zinc on the inner side. zinc particles leading to the formation of a high-alkalinity
As the cemented silver grows outward the zinc disk shrinks; microenvironment. Hydrogen is then transported either by
the bright hoop expands separating completely both metals diffusion or through gas microbubbles to the inner silver–
(Figs. 6a and 6c). Additionally, Fig. 6b shows gas bubbles cyanide interface. It then continues its diffusion through the
detected inside the hollow hoop, with high probability they pores of the silver shell until it finally reduces silver–cyano
G. Viramontes Gamboa et al. / Journal of Colloid and Interface Science 282 (2005) 408–414 413

species on the outer silver–aqueous solution interface. The which is 1.5 × 10−5 cm2 /s [6]. Molecular hydrogen meets
reduction reaction on this interface may be either a simple this requirement as it is smaller in size in comparison with
redox reaction given by that of Ag(CN)− 2.
1 These novel results demonstrate that some electrochem-
Ag(CN)− 2 + 2 H2 → Ag + 2CN + H
0 − + (8) ical methods, such as Evans diagrams, should be used with
a lot of precautions to carry out silver cementation on zinc
or an electrochemical scheme in which hydrogen oxidation studies particularly at high cyanide concentrations.
and silver reduction take place at physically separated sites
as follows
1 4. Conclusions
H2 → H+ + e − , (9a)
2
The chemical routes and morphological patterns present
Ag(CN)− − −
2 + e → Ag + 2CN ,
0 (9b) in the cementation of silver from cyanide solutions with zinc
1 powder depend on the cyanide concentration. At low free
Ag(CN)−
2 + H2 → Ag0 + 2CN− + H+ . (9c) cyanide concentrations an electrochemical process taking
2
place on the zinc particles describes well experimental ob-
Both Reactions (8) and (9c) leads to the formation of servations. For high free cyanide concentrations, like those
a less-alkaline microenvironment. The chemical species in- commonly used in plant operations, zinc oxidation and silver
volved in those reactions are adsorbed at the outer silver– reduction reactions take place on different, non-electrically
aqueous solution interface. Adding Reactions (7c) and either connected, interfaces. Under this last condition, the absence
(8) or (9c) leads to the overall silver cementation stoichiom- of contact between both metals causes the reduction of the
etry for high cyanide concentrations given by Reaction (6). efficiency of zinc to recover silver, exactly by one half of its
Accordingly, the ε = 1 relation originates because the areas theoretical maximum.
for zinc dissolution and silver deposition are not electrically
but electrolytically connected. Maybe the role of hydrogen
microbubbles is to separate the zinc powder particles from Acknowledgments
the silver shell.
The two different mechanisms of silver deposition at This work has been partially supported by the Consejo
low and high cyanide concentration can be accounted for Nacional de Ciencia y Tecnología, México, Grants G29589E
by the stability of Zn(OH)2 and its relationship to the and ER026 Materiales Biomoleculares, and by Centro de
rate-determining steps of the cementation process. At low Investigación y Desarrollo Tecnológico de Servicios Indus-
cyanide concentration and high pH, Zn(OH)2 is highly sta- triales Peñoles. The authors thank Professor Bernardo José
ble and it has been found that it coats the surface of zinc Luis Arauz Lara for the assistance in the use of the equip-
[1,4,6,16]. Under these conditions the rate-determining step ment at the Laboratorio de Fluidos Complejos, Instituto de
is the zinc oxidation [17]. Thus, silver–cyano species diffuse Física, Universidad Autónoma de San Luis Potosí. Thanks
from the bulk solution to the zinc–aqueous solution interface to Dr. Miguel Angel Vidal for use of the mass spectrometer,
where they are cathodically reduced. Then the cementation Instituto de Investigación en Comunicaciones Opticas, from
process follows a shrinking core behavior as observed in the same University, and for helpful discussions.
Fig. 4a, which shows the cemented silver growing toward
the center of the zinc particle.
References
At high cyanide concentration, zinc–cyano complex
species are more stable than Zn(OH)2 [6,16]. Under these [1] G. Viramontes Gamboa, M. Medina Noyola, A. López Valdivieso, Hy-
conditions the cementation process has been reported to be drometallurgy (2004), submitted for publication.
controlled by diffusion through the boundary layer [4,6], [2] I. Barin, H. Barth, A. Yaman, Erzmetall. 33 (7/8) (1980) 399–403.
[3] D.W. Kirk, F.R. Foulkes, J. Electrochem. Soc. 131 (4) (1984) 760–769.
which agrees well with the dense-branching morphology of [4] J.D. Miller, R.Y. Wan, J.R. Parga, in: Proc. of Symp. on Precious and
the silver deposit [1] and a first-order rate for the cemen- Rare Metals Technologies, Albuquerque, NM, USA, 1988, pp. 281–
tation Reaction [6]. When cementation starts, silver–cyano 290.
species are first reduced at the zinc–aqueous solution inter- [5] M.J. Nicol, E. Schalch, P. Balestra, H. Hegedus, J.S. Afr. Inst. Min.
face forming the initial silver layer, which is readily elec- Metall. 79 (1979) 191–197.
[6] J.R. Parga, Analysis of the Zinc Cementation Reaction for Recovery
trically charged driving the silver–cyano species adsorbed of Precious Metals from Cyanide Solutions, Ph.D. dissertation, Dept.
at this interface to be discharged. So the diffusion control of Metallurgy and Metallurgical Engineering, University of Utah, Salt
through the boundary layer is established and the reduction Lake City, 1987.
of silver–cyano species will take place at the silver–aqueous [7] J.R. Parga, R.Y. Wan, J.D. Miller, Miner. Metall. Process. SME (1998)
solution interface making the deposit grow outward. For this 170–176.
[8] E.A. Von Hahn, T.R. Ingraham, Trans. TMS-AIME 239 (1967) 1895–
to occur in the absence of an electrical contact between the 1900.
silver layer and the zinc particle, the reducing species needs [9] C.A. Flemming, Hydrometallurgy 30 (1992) 127–162.
to have a diffusion coefficient higher than that of Ag(CN)− 2, [10] A. James, Cyanide Practice, third ed., 1902.
414 G. Viramontes Gamboa et al. / Journal of Colloid and Interface Science 282 (2005) 408–414

[11] J.V.N. Dorr, F.L. Bosqui, Cyanidation of Gold and Silver Ores, [15] K. Osseo-Assare, in: C.A. Young, A.M. Alfontazi, C.G. Ander-
McGraw–Hill, New York, 1950, pp. 184–187. son, D.B. Dreisinger, B. Harris, A. James (Eds.), Hydrometallurgy
[12] S. Vilchis-Carbajal, Modelación matemática de la cinética de ce- 2003—Fifth International Congress in Honor of Professor Ian Ritchie,
mentación de plata con polvo de zinc a partir de soluciones cianuradas, vol. 2; Electrometallurgy and Environmental Hydromatallurgy, Miner-
M.S. thesis, Instituto Politénico Nacional, Mexico, D.F., 1992. als, Metals, and Materials Society, 2003, pp. 1151–1165.
[13] A.H. Webster, J. Halpern, J. Phys. Chem. 61 (1957) 1245–1248. [16] G. Chi, M.C. Fuerstenau, J.O. Marsden, Int. J. Miner. Process. 49
[14] G. Deshenes, G. Ritcey, G.M. Ghali, in: M.C. Jha, S.D. Hill (Eds.), (1997) 171–183.
Precious Metals ’89, Minerals, Metals, and Materials Society, 1988, [17] S. Vilchis Carbajal, I. González, G.T. Lapidus, J. Appl. Elec-
pp. 341–357. trochem. 30 (2000) 217–229.

You might also like