You are on page 1of 13

InTERnIITIO01# JOURIllaLOF

mlnERHL PRO[ESSln6
ELSEVIER Int. J. Miner. Process. 49 (1997) 171-183

Study of Merrill-Crowe processing. Part I: S,olubility of zinc in alkaline cyanide solution


Gexla Chi *, Maurice C. Fuerstenau, John O. Marsden
Universi~ of Nevada-Reno, Mackay School of Mines, Reno, NV 89557, USA

Abstract

A comprehensive study of zinc solubility was carded out in de-aerated alkaline cyanide solutions. Parameters investigated were: dissolved oxygen, cyanide and lead nitrate concentrations and pH Zinc solubility increases significantly with increased dissolved oxygen and cyanide concentrations. Eh-pH diagrams have been developed and used to predict regions of stability of zinc hydroxide, which may inhibit gold precipitation. The interaction between cyanide and hydroxyl ion activities to prevent the formation of zinc hydroxide is presented. With lead nitrate additions, maximum solubility of zinc occurs in the range of 10-15 mg/l lead nitrate. SEM photomicrographs of zinc particles in the presence of lead nitrate are presented.
Keyword~: zinc solubility; Eh-pH diagrams; Merrill-Crowe processing

1. Introduction

Precious metal recovery from aqueous solution by zinc dust precipitation has been in use since the late 1800s [1]. This technology, originally patented by Salman and Pichard, was applied in 1897 to the Homestake operation in Lead, South Dakota by C.W. Merrill. In 1916, the process was refined with the introduction of T.B. Crowe's vacuum deaerator with which considerable improvement in efficiency was effected [1]. Since then, the Merrill-Crowe process has seen extensive application in the precipitation of precious metals from cyanide solution. Merrill-Crowe systems operate effectively most of the time, but at times and for unexplained reasons, precious metals are not precipitated efficiently. In other cases, well-de, signed systems do not achieve the expected low, barren-solution grades. When these phenomena occur, large quantities of precious metals can be retained in solution

* Corresponding author. 0301-75 L6/97/$17.00 Copyright 1997 Elsevier Science B.V. All rights reserved. Pll S0301-7516(96)00043-9

172

G. Chiet al. l i n t . J. Miner. Process. 49 (1997) 171-183

Au(CN) 2 C'-i " AulCN) 2 CN"

ON-//

~\

I "11~--~Depositgold ed

particle

"~"------~ ] ~ " ~

CN-

ZnlCN)4 2Fig. 1. Schematicof mechanismof gold precipitationon zinc [14].


hampering gold production efficiency. It was decided to investigate the influence of some of the factors which affect the efficiency of precious metals precipitation from cyanide solutions by zinc dust under controlled conditions and with pure solutions. Although the Merrill-Crowe process is widely used in the precious metals industry, most of the published literature on the cementation of gold and silver prior to 1960 dealt largely with plant practice [2]. Detailed laboratory investigations have only been reported during the past two decades. Studies have been mostly confined to the kinetics of gold and silver cementation on rotating discs or cylinders [3-6]. In the late 1980s, comprehensive investigations were conducted by Parga et al. [2,7,8] to study the mechanisms of the cementation reactions. The cementation of gold and silver by zinc dust is an electrochemical process, proceeding by localized anodic and cathodic reactions. A schematic of this process is shown in Fig. 1. The main reactions for zinc dissolution are: Zn + 4 C N - + 1/202 + H20 = Zn(CN)~- + 20H 2Au(CN)2- + Zn = 2Au + Zn(CN)4 Zn + H20 + 2Au(CN)2- = 2Au + HZnO 2 + 3 H + + 4 C N (1) (2) (3)

To better understand the phenomena controlling zinc efficiency in these systems, the solubility of zinc in alkaline cyanide solution was established under various conditions in the absence of precious metals. The effects of cyanide, oxygen and lead nitrate concentrations and pH on zinc solubility were investigated.

2. Experimental methods and materials


In order to simulate the Merrill-Crowe process, a laboratory procedure was established which allowed de-aeration of the solution to achieve a desired dissolved oxygen concentration. Zinc concentration was measured by Atomic Absorption Spectroscopy.

G. Chiet al. lint. J. Miner. Process. 49 (1997) 171-183

173

Solution

filter

sample bottle

[3

Fig. 2. Schematic of experimental apparatus. 1 = transfer tube for adding zinc; 2 = transfer tube for adding Pb(NO3):; 3 = oxygen meter probe; 4= tube for air escaping; 5= glass tube for solution withdrawal; 6 = magnetic stir bar; 7 = magnetic stirrer; 8 = gas dispersion tube. A11 of the experiments were conducted in a four-necked 500-ml glass reaction vessel. A dissolved oxygen probe, nitrogen dispersion tube, sampling device and zinc and lead nitrate addition tubes were placed into the reactor through the openings in the cover, as shown in Fig. 2. In all of the studies, purified nitrogen was passed through the solution via a dispersion tube prior to conducting the experiment. A YSI Model 58 dissolved oxygen meter was used to measure the concentration of dissolved oxygen in the solution. The test solution was prepared with reagent-grade NaCN. Reagent-grade NaOH was used to adjust the pH of the solution. Distilled water was used for all solutions. Before and after every batch experiment, the solution was titrated for cyanide concentration. In each test, 18 m g / l of Merrillite zinc (Pasco) was added with various amounts of Pb(NO3) 2. Solution was withdrawn by means of a peristaltic pump for analysis of zinc concentration when oxygen concentration did not decrease with time.

3. Thermodynamic considerations of the Zn--CN--H20 system


In order to understand in more detail the dissolution of zinc in cyanide solution, it is useful to consider the thermodynamics of the system utilizing Eh-pH diagrams. These diagrams delineate the stability regions of various specious at equilibrium and, consequently, are useful in the study of the solubility of zinc in cyanide solution. It has been suggested that the following reactions occur when zinc dissolves in cyanide solution: Zn + 4 C N - = Z n ( C N ) ] - + 2e Zn + 3 O H - = HZnO2- + H 2 0 + 2e (4) (5)

These half reactions can be coupled with the water discharge half cell to produce hydrogen: Zn + 4 C N - + 2 H 2 0 = Z n ( C N ) ] - + 2 O H - + H 2 Zn + 2 H 2 0 = HZnO 2 + H + + H 2 (6) (7)

174

G. Chi et al./lnt. J. Miner. Process. 49 (1997) 171-183

All of these reactions are electrochemical processes and may be resolved into their appropriate half cells for the construction of pertinent Eh-pH diagrams. The essential half-cell reaction for zinc dissolution in aqueous solution is: Zn 2+ + 2e = Zn E = - 0 . 7 6 2 + 0.0295 log [Zn 2+ ]

(8)

Zinc is so strongly reducing that even at relatively high Zn 2+ concentrations, the potential lies well below that required to reduce water. In alkaline cyanide solutions, zinc forms numerous complexes by reaction with hydroxyl and cyanide ions. Of these, one, Zn(OH) 2, is sparingly soluble. The rest are soluble. The relevant equilibria involving metallic zinc are: Zn(CN)4z- + 2e = Zn + 4 C N E = - 1 . 2 5 + 0.0295 log [Zn(CN)4-] + 0.118pCN Zn(OH)2 + 2e = Zn + 2 O H E = - 0 . 4 1 9 - 0.059pH HZnO] + 3H + + 2e = Zn + 2H20 E = 0.054 - 0.0886 pH + 0.0295 log [HZnO 2 ] ZnO 2- + 2H20 + 2e = 4 O H - + Zn E = - 1.216 + 0.0295 log [ZnO~- ] + 0.118 pOH (12) (11)
r _ _~ 1

(9)

(10)

In addition, hydrocyanic acid acts as a weak acid, and its hydrolysis can be represented as: H + + C N - = HCN pH + pCN = 9.4 - log[HCN] (13)

Eq. (13) indicates a considerable reduction in the cyanide activity when the pH falls below the value of 9.4. The total cyanide content, A, as referred to in practice, is defined by the following equation: A = [HCN] + [ C N - ] and Eq. (13) becomes: pH + pCN = 9.4 + log(1 + 10 pH-94)
_ _

(14)

loga

(15)

By combining Eq. (9) and Eq. (15), one obtains: Zn(CN)4- + 2e = Zn + 4 C N E = - 1 . 2 5 + 0.0295 log[Zn(CN) 2- ] + 0 . 1 1 8 1 9 . 4 - l o g A + log(1 + 10 pH-9"4) - p H ] The Eh-pH diagrams describing these reactions are shown in Fig. 3a-c. The calculations are based on the thermodynamic data taken from references [9-11]. (16)

G. Chi et al. / Int. J. Miner. Process. 49 (1997) 171-183

175

(a)

1.0 0.0
-I .0 -2.0

.............
Zn ~

[
" .....
i

2+u,oG~-_-i~r ......

--1

] 11
.... -1__
~

i
I I

>

.............

=,~=

6 pH

10

12

14

...............

;
I

z.-

......

o.o . . . . . .

.
-2.1

2.2o+~e~o.-~_.

.......

. .....
....

,,

z,c~4, / ~~, , , ~ ,
/~/-~
I

-I---t---~ ~ |
,

---- ]
=

~- . . . . . . . .

', ~/
Co)

Zn 2 4 6 8 10 12

14
%

pH

/
(C) 1.0-

. _. .++ . .O 2 + H 2 0 + 4 e = 4 O H .
~"
2

--I

'. . . . . . . . . . . . . . .
'

I 1 ' 1l
][--;--1 II o~' ', ~ I1~ ',~=
/

0.0 . . . . . . . .
-1.0 "
-2.0 0

.~o+2o=~o,-,-+.~
~
2 4

..........

:
.... . .......
I

z"~cro,2
n

z
6 pH

10

12

14

Fig. 3. Eh-pH diagrams of the Zn-CN-H20 system at three cyanide concentrations,[Zn(II)]= 1.5 ><10-4 M (10 mg/1), 25C; NaCN: (a) 0.1 g/l, (b) 0.3 g/l, (c) 0.5 g/l.

From these figures it can be seen that zinc will be oxidized at all pH values, while dissolved oxygen will be reduced. In addition, zinc can also reduce water to hydrogen at all pH values. For convenience, activities and concentrations have been assumed to be equal, which is a reasonable assumption considering the relatively low concentration of cyanide and zinc species in solution,

176

G. Chi et al. / Int. J. Miner. Process. 49 (1997) 171-183

4. Results and discussion

Experiments were conducted first to establish the reproducibility of these measurements. Two sets of experiments were conducted, the results of which are presented in Table 1. 4.1. Dissolved oxygen concentration It has been suggested that inefficient removal of oxygen in the vacuum deaeration tower is a serious problem in many Merrill-Crowe systems [12]. Excessive consumption of zinc occurs under these conditions. As shown in Fig. 4, as the level of oxygen is increased, especially at higher concentrations of cyanide, the solubility of zinc is increased substantially. In fact, about 50% of the zinc is dissolved when 0.5 m g / l 02 and 0.5 g / l NaCN are present. The relevant chemical reaction is: Zn+4CN-+1/202+H20=Zn(CN) ] +2OH (17)

If one only considers the effectiveness of the zinc dust, Fig. 4 suggests that the lower the dissolved oxygen concentration, the better. But evidence has been put forward to show that the completeness of gold cementation is increased by low concentrations of oxygen (less than 1 m g / l ) in solution [9]. In this context, it is perhaps of interest that solutions proceeding to precipitation after de-aeration by the Crowe vacuum process typically have oxygen concentrations in the range of 0.5 to 1.3 m g / l [13]. Halbe [12] also found that the amount of oxygen dissolved in a typical pregnant solution is commonly reduced to about 0.5 ppm by the Crowe vacuum. In view of these facts, the rest of the discussion is based on experimental data obtained in solutions with 0.5 mg/1 initial dissolved oxygen. 4.2. Cyanide concentration Insufficient free cyanide in solution is also a common problem encountered in Merrill-Crowe precipitation, which may result in the formation of Zn(OH) 2 and,

Table 1 Experimental reproducibility Experiment


1

Solubility of zinc (mg/1)


2

Conditions: 0.3 g/I NaCN, pH 11.2, 0.5 mg/l 02, 22C: 2 3 4.4 4.7

Conditions: 0.5 g/1 NaCN, pH 11.2, 0.5 mg/1 02, 22C: 1 8.6 2 8.7

G. Chi et al. l i n t . J. Miner. Process. 49 (1997) 171-183


12,

177

10,

2
0 1 Oxygen, mg/I

Fig. 4. Effect of initial dissolved oxygen concentration on zinc solubility; pH 10.9, 0.1 g/1 NaCN, 22C.

consequently, produce high barren solutions. This can be explained by inspection of the Eh-pH diagrams in Fig. 3a-c. These show that with increasing cyanide concentration, the phi range for Zn(OH) 2 precipitation is decreased. When the sodium cyanide concentration exceeds 0.5 g / l , Zn(OH) 2 precipitation does not occur at this zinc concentration at any pH value. Fig. 5 expresses the relationship between zinc solubility and cyanide concentration. From these results it can be seen that, in alkaline cyanide solution, the solubility of zinc

12-

10.

a.
E
~,
,

o o N

2.

0.2

0.4

0.6

NaCN, g/I

Fig. 5. Effectof cyanideconcentrationon zinc solubility,pH 10.5,0.5 mg/l 02 and 22C.

178

G. Chi et al. l i n t . J. Miner. Process. 49 (1997) 171-183 22. 21


o~ 20

c oJ 19 tll c 18 N
17 y = -0.0948x + 87495 R2 = 08967

16 690 700 710 720 N a C N , g/t 730 740 750

Fig. 6. Regression analysis o f plant operating data.

increases with increasing cyanide concentration due to the increased stability of zinc cyanide complexes. At low cyanide concentration, the equilibrium between Zn(CN)J- and Zn(OH) 2 is considered when zinc hydroxide is stable [14], and the relevant reaction is: Zn(CN)42- + 2 O H - = Zn(OH)2(~) + 4CNK = 3.50 X
10 . 3

(18)

It is thus obvious that the increase in cyanide concentration results in an increase in Zn(CN) j - concentration and, consequently, reduces the amount of Zn(OH) 2 formed, which is beneficial to Merrill-Crowe precipitation. High cyanide concentration is not always advantageous, however. When the cyanide concentration is in excess of the value required to avoid any significant formation of zinc hydroxide and sufficient free cyanide is present to maximize the rate of precious metals precipitation, then excess zinc will dissolve. Regression analysis of plant operating data corroborates this observation. The regression model, presented in Fig. 6, shows that at high cyanide concentration (above 690 g / t NaCN), an increase in cyanide concentration results in a decrease in zinc efficiency due to excessive dissolution of zinc, while zinc efficiency is defined as the stoichiometric zinc requirement for precious metals precipitation divided by the total amount of zinc added.

4.3. Effect of pH
The formation of Zn(OH) 2 is a major concern in Merrill-Crowe precipitation, since its formation retards the cementation of precious metals. Theoretically, Eh-pH diagrams can be used to predict the pH region of Zn(OH) 2 formation and stability. But Eh-pH diagrams are constructed on the basis of known thermodynamic data and have certain inherent limitations [14]. The use of solubility versus pH diagrams in conjunction with

G. Chi et al. / lnt, J. Miner. Process. 49 (1997) 171-183


10,

179

8,

%
E

c~

4,

"-O--0.1 g/I NaCN -1l--0.3 gll NaCN ,,-.k--0.5 g/I NaCN

10

11
pH

12

13

Fig. 7. Effect of pH on zinc solubility, 0.5 mg/1 02, 22C.

Eh-pH diagrams is a practical approach to determine the pH range of Zn(OH) 2 formation. That is, when insoluble Zn(OH) 2 is formed, the activity of zinc in solution will be strictly constrained by the low value of the solubility product of Zn(OH) 2. This results in an obvious decrease in zinc solubility. Zinc solubility as a function of pH is presented in Fig. 7. It can be seen that: 1. for cyanide concentrations of 0.1 and 0.3 g / l NaCN, zinc solubility reaches a minimum at a pH range of 10.5-11.2, which implies the formation of some insoluble zinc hydroxide; and 2. for a cyanide concentration of 0.5 g/1 NaCN, zinc solubility reaches a minimum at pH e f about 12.0, which also implies the formation of insoluble zinc hydroxide. Comparison of the above results with those in Fig. 3a-c indicates good agreement betweerL the thermodynamic predictions and experimental results. At cyanide concentrations ot' 0.1 and 0.3 g / l NaCN, Zn(OH) 2 formation is predicted in the pH ranges 10.4-12.0 and 11.0-12.0, respectively. This is supported by the experimental data (Fig. 7) which indicates that some Zn(OH) 2 is being formed over the pH range 10.3-11.4 at the 0.3 g/1 NaCN concentration and that more Zn(OH) 2, as indicated by its effect on zinc so]Lubility, is being formed over the pH range 10.3-11.2 at the lower (0.1 g / l ) NaCN concentration. The increased solubility of zinc observed experimentally over the pH range 11.2-12.0 is presumably due to the formation of increasing concentrations of HZnO~-, as predicted by thermodynamics. Above pH 12.0 the soluble HZnO 2 species predominates. Some mixed hydroxy-cyano complexes, such as Zn(OH)x(CN)~-, may also be formed under these conditions which would contribute to the increased zinc solubility [9].
4.4. Lead concentration

The improvement experienced in Merrill-Crowe processing when lead salts are added was demonstrated in 1894. It has long been considered that this improvement is due to

180

G. Chi et al. ~Int. J. Miner. Process, 49 (1997) 171-183


i 1.0"
2 +H,O+" ~ ...... m -"

oH
pb.~

.... . -L
m --___

0.5" 0.0" 2"H30 + 2e = 20H H,


t,,tU

." = PbO

. . . .

.....
,

l1
i~! e , ,1~ I I i

IIll,=,=l=l=,Wlmlml~,~llllll.lmL~.=,ml=lll

I=1=~ ~U 6 , r o l l ,
I

-0.5 -1.0 -1.5 -2.0 0


i

Pb

n~

I Zn(CNh 2 I

- - -

Zn

6 pH

10

12

14

Fig. 8. Eh-pH diagram of Zn-Pb-CN-H20 system; 0.l g/1 NaCN, [Zn(II)]=1.5X 10 - 4 M, 10 mg/l
Pb(NO3) 2

the formation of a zinc-lead couple [9]. The Eh-pH diagram of Z n - P b - C N - H 2 0 system is given in Fig. 8, which shows thermodynamically that lead can be reduced to the metallic state in alkaline cyanide solution. Pb 2+ is reduced at the zinc surface forming cathodically-charged areas of metallic lead [14]. The negatively-charged gold and silver cyanide species are thus thought to be reduced preferentially in these regions. This helps to localize gold and silver deposition, preventing the entire surface of zinc from becoming coated with gold. The experimental results showing the effects of lead nitrate concentration on zinc solubility are presented in Fig. 9. At 10-15 ppm lead nitrate concentration (0.006-0.009 g/1 Pb2+), zinc solubility reaches a maximum under conditions of 0.1 and 0.3 g/1
10

~8
o
- ' 0 = ' 0 . 1 g/I N a C N ,,--A,--0.3 g/I N a C N ' 4 ' - 0 . 5 g/I N a C N

10

20

30

Lead Nitrate,

mg/I

Fig. 9. Effect of lead salt addition on zinc solubility, 0.5 mg/l

02, pH 10.9, 22C.

G. Chi et a l . / Int. J. Miner. Process. 49 (1997) 171-183

181

Fig. 10. SEM photograph of zinc particles showing lead reduction on the zinc surface; 15 mg/1 Pb(NO3)2, 0.1 g / l NaCN and pH 10.5.

NaCN, which coincides with the observed optimal lead concentration for Merrill-Crowe precipit~ition [14]. The increase in zinc solubility implies that it is the reduction of lead ion on the zinc surface which enhances zinc dissolution in alkaline cyanide solution. SEM photography, as shown in Fig. 10, has been used to demonstrate this phenomenon. The detrimental effect of excess lead nitrate addition was examined utilizing scanning electron microscopy. Fig. 11 shows a zinc particle that was completely coated with lead deposits when 150 mg/1 Pb(N03) 2 was added. Lead has also been shown to reduce the inhibiting effect of zinc hydroxide formation on the zinc surface at low cyanide concentrations. Parga [2] found that the presence of small mrlounts of lead ion shifts the passivation current of the zinc anodic polarization

Fig. 11. SEM photograph of zinc particles, one of which is totally coated with lead deposit; 150 mg/l Pb(NO3) z, 0.1 g/1 NaCN and pH 10.5.

182

G. Chi et al. / Int. J. Minel: Process. 49 (1997) 171- I&

curve to a higher level, in the extreme case by an order of magnitude. This finding can be used to explain Fig. 9. At low cyanide concentrations (0.1 and 0.3 g / l NaCN), zinc hydroxide is thermodynamically stable, while at high cyanide concentration (0.5 g/1 NaCN), zinc hydroxide is thermodynamically unstable. Thus, the beneficial effect of lead salt addition in systems of low cyanide concentration is more apparent.

5. Summary and conclusions


The following conclusions can be drawn from this study: 1. The solubility of zinc increases with increasing dissolved oxygen concentration in solution. Thus, any failure to maintain low dissolved oxygen concentration (i,e., < 0.5 rag/l) in the Crowe vacuum tower will cause excessive consumption of zinc dust and poor zinc efficiency. 2. The solubility of zinc increases with increasing cyanide concentration in alkaline cyanide solution. High cyanide concentration is beneficial to gold and silver cementation, but lower zinc efficiency is experienced due to excessive zinc dissolution. 3. Regression analysis of plant operating data shows that high cyanide concentrations will result in a decrease of zinc efficiency for Merrill-Crowe precipitation. This is in good agreement with the laboratory results of this investigation. 4. There is no common pH value in alkaline cyanide solution at which precipitation of zinc hydroxide will occur. The pH for Zn(OH) 2 precipitation depends on the cyanide concentration. 5. Zinc solubility reaches a maximum at 10-15 m g / l lead nitrate addition, which coincides with the optimal lead concentration for Merrill-Crowe precipitation that is observed in practice. 6. SEM photography demonstrates the role of the zinc-lead couple when a lead salt is added to a simulated precipitation solution. This phenomenon supports the long suggested assumption that the improvement of Merrill-Crowe precipitation with lead salt addition is due to the reduction of lead on the zinc surface. Evidence was also obtained from SEM photography that zinc particles may become totally coated with lead deposits when excess lead salt is added.

Acknowledgements
The authors wish to acknowledge the financial support provided by the Center for Precious Metals of the Mackay School of Mines, University of Nevada, Reno.

References
[1] Leblanc,R., 1942. Precipitation of gold from cyanide solution by zinc dust. Can. Min. J., April, May and June Issues, pp. 213-219, 297-306, 371-379. [2] Parga, J.R., 1987. Analysis of the Zinc Cementation Reaction for Recovery of Precious Metals from Cyanide Solutions. Ph.D. Dissertation, Dept. of Metallurgy and Metallurgical Engineering, Universityof Utah, Salt Lake City, April, 1987.

G. Chi et al. l i n t . J. Miner. Process. 49 (1997) 171-183

183

[3] Strickland, P.H. and Lawson, F., 1971. The cementation of metals from dilute aqueous solution. Proc. Australas. Inst. Min. Metall., March: 71-79. [4] Miller, J.D.,1981. Solution concentration and purification, metallurgical treatises. In: K.T. John and J.F. Elliot (Editors), AIME, pp. 95-116. [5] Annamalai, V. and Murr, L.E. 1979. Influence of deposit morphology on the kinetics of copper cementation on pure iron. Hydrometallurgy, 4: 57. [6] Riddiford, A., 1966. The rotating disc system. Adv. Electrochem. Eng., 47: 4-87. [7] Parga, J.R., Wan, R.Y. and Miller, J.D. 1988. Zinc dust cementation of silver from alkaline cyanide solutions - Analysis of Merrill-Crowe Plant data. Min. Metall. Process. August: 170-176. [8] Miller, J.D., Wan, R.Y. and Parga, J.R., 1990. Characterization and electrochemical analysis of gold cementation from alkaline cyanide solution by suspended zinc particles. Hydrometallurgy, 24: 373-392. [9] Finkelstein, N.P., 1972. The chemistry of the extraction of gold from its ores. In: R.J. Adamson (Editor), Gold Metallurgy in South Africa. Chamber of Mines of South Africa, Johannesburg, pp. 284-351. [10] Pou[baix, M., 1966. Atlas of electrochemical equilibria in aqueous solution. Pergamon Press, London, pp. 406--413. [11] Latimer, W.M., 1952. Oxidation Potentials. Prentice Hall, pp. 168-182. [12] Halbe, D., 1985. Recovery of gold and silver from leach solution. Heap and Dump Leaching Newsletter, May DHL Company, Colorado, USA, June, 1985, pp. 5-9. [13] Wartenweiler, F., Precipitation of gold from cyanide solution. In: A. King (Editor), Gold Metallurgy on the Witwatersrand. Johannesburg, Transvaal and Orange Free State, Chamber of Mines, 1949, Chapter 12. [14] Marsden, J.O. and House, C.I., 1992. The Chemistry of Gold Extraction. Ellis Horwood, Chichester. UK and Prentice-Hall, Englewood Cliffs, NJ, 1992.

You might also like