Вы находитесь на странице: 1из 156

1

THE CHEMICAL AND KINETIC MECHANISM

FOR LEACHING OF CHRYSOCOLLA


BY SULFURIC ACID

by

Mofen Jiang

A Thesis Submitted to the Faculty of the

DEPARTMENT OF HYDROLOGY AND WATER RESOURCES


In Partial Fulfillment of Requirements
For the Degree of
MASTER OF SCIENCE
WITH A MAJOR IN HYDROLOGY

In the Graduate College


THE UNIVERSITY OF ARIZONA
1992

STATEMENT BY AUTHOR

The thesis has been submitted in partial fulfillment of requirements for an


advanced degree at the university of Arizona and is deposited in the University
Library to be made available to bollowers under rules of the library.

Brief quotations from this thesis are allowable without special permission,
provided that accurate acknowledgement of source is made. Requests for permission
for extended quotation from or reproduction this manuscript in whole or in part may

be granted by the head of the major department of the Dean of the Graduate
College when in his or her judgement the proposed use of the material is in the
interests of scholarship. In all other instances, however, permission must be obtained
from the author.
SIGNED:

APPROVAL BY THESIS DIRECTOR


This thesis has been approved on the data shown below.

R. L. Bassett
Associate Professor
Hydrology and Water Resources

Date

ACKNOWLEDGEMENTS

This research was performed at the University of Arizona under the


supervision of Dr. Randy L. Bassett. Financial support was provided by the Copper

Research Center of the University of Arizona.

First of all, I greatly appreciate Dr. Randy L. Bassett for his constant
guidance, attention, patience, and encouragement; without him, this work would not

have been completed.


I also wish to acknowledge Dr. Brent Hiskey, a co-supervisor of this project

and committee member, for his technical guidance, support in equipment, field
sampling, and academic advising. I also extend my thanks to Dr. Charles W. Kreitler

for his review and critique of this work.

I would like to thank all of my research group members, especially Gregg


Davidson, a close friend who gave me a lot of help through it all.

Finally, I express my deepest thanks to my husband, Daquan Tian, for his


unending encouragement, thoughtful comments, and constant support.

TABLE OF CONTENTS
Page

LIST OF ILLUSTRATIONS

LIST OF TABLES

11

ABS IRACT

13

INTRODUCTION

14

1.1 Dissolution Studies

15

1.2 Field Study

16

1.3 Computer Modeling

18

LITERATURE REVIEW

21

2.1 Copper Solution Mining

21

2.2 Mineral Characterization of Chrysocolla

26

2.3 Chemical and Kinetic Mechanism of Chrysocolla Dissolution

28

METHODS

37

3.1 Chrysocolla Sample Preparation

38

3.2 Batch Dissolution Experiment

39

3.3 Field Sampling

42

3.4 Chemical Analysis

44

3.5 Surface Analysis by Scanning Electron Microscope

48

3.6 Calculation of Copper Extraction

48

RESULTS AND DISCUSSION

50

TABLE OF CONTENTS - Continued


Page

4.1 Experimental Results and Discussion

50

4.1.1 Chemical Composition of Chrysocolla

50

4.1.2 Stoichiometry of Proton Consumption

50

4.1.3 Copper Extraction and Rate of Dissolution

54

4.1.4 Morphology

62

4.2 Aqueous Chemistry Results for Field Samples

5. GEOCHEMICAL COMPUTER MODELING


5.1 Background

80
84
84

5.1.1 Ion Association Theory

86

5.1.2 Ion Interaction Theory

90

5.2 Calibration of the PHRQPITZ Code in a Cu-Al-SO4 System

92

5.3 Saturation Analysis for Batch Dissolution Experiments by


PHREEQE and PHRQPITZ

98

5.3.1 Model Input


5.3.2 Model Results

5.4 Saturation Analysis for Field Samples by Geochemical


Computer Models

98

100

107

5.4.1 Estimation of Eh

107

5.4.2 Model Input

112

5.4.3 Model Results

115

TABLE OF CONTENTS - Continued


Page
6. CONCLUSIONS

130

7. APPENDICES

134

pH-Stat computer program, written in Basic

135

The update data file for geochemical computer model


PHRQPITZ: PITZER.DATA

138

The update data file for geochemical computer model


PHRQPITZ: PHRQPITZ.DATA

144

The analytical results from acid mine drainage in West Squaw


Creek and its tributaries (Filipek et al, 1987)

150

8. LIST OF REFERENCES

152

LIST OF ILLUSTRATIONS
Page

Figure

Layout of a typical copper solution mining system (Magma


Copper Company, 1991)

17

2.1.1

General layout of a copper heap leaching system (Hiskey, 1986)

22

2.1.2

Cross section of a copper waste dump showing one lift under


active leaching and another in a rest period (Hiskey, 1986)

22

2.1.3

Three types of in-situ leaching (After Wadsowrth, 1979)

24

2.2.1

Chrysocolla structure (After Van Oosterwyck-Gastruche, 1970)

27

2.3.1

Shrinking core model for dissolution of chrysocolla

30

2.3.2

Reaction zone model (After Wadsworth, 1972)

33

3.3.1

Schematic diagram of batch experimental apparatus for


chiysocolla dissolution experiments

40

1.2.1

4.1.1(a-h) The relationship between the production of copper to the


consumption of proton from batch dissolution experiments
at different pH values and solution compositions
4.1. 1(i-k)

The relationship between the production of copper to the


consumption of proton from batch dissolution experiments
at different pH values and solution compositions (continual)

52

53

4.1.2(a-d) Copper extraction versus time for batch dissolution experiments

57

4.1.3(a-d) Diffusion mode correlating data from the batch dissolution


experiments

59

4.1.4(a)

SEM micrograph for plug 1 at a magnification of 3,000X

63

4.1.4(b,c)

Non-dispersive X-ray scan of b) the bright, and c) the dark


surface area of plug 1 indicating qualitatively the composition

64

LIST OF ILLUSTRATIONS - Continued


Page

Figure

4.1.5(a,b) SEM micrographs for a) grain 1, and b) grain 2 of plug 2 at a


magnification of 65X

66

4.1.5(c,d) SEM micrographs for bright phases of c) grain 1, and d) grain 2


of plug 2 at a magnification of 7,000X

67

SEM micrographs for dark phases of e) grain 1, and


f) grain 2 of plug 2 at a magnification of 7,000X

68

Non-dispersive X-ray scan of the bright surface area of g) grain 1,


and h) grain 2 of plug 2 indicating qualitatively the composition

69

Non-dispersive X-ray scan of the dark surface area of i) grain 1,


and j) grain 2 of plug 2 indicating qualitatively the composition

70

4.1.6(a)

SEM micrograph for a grain of plug 3 at a magnification 50X

71

4.1.6(b)

SEM micrograph for the bright phase of plug 3 at


a magnification 7,000X

72

Non-dispersive X-ray scan of the bright surface area of plug 3


indicating qualitatively the composition

72

4. 1.5(e,f)

4.1.5(g,h)

4.1.5(i,j)

4.1.6(c)

4.1.6(d)

4.1.6(e)

4.1.6(f)

4.1.6(g)

4.1.6(h)

SEM micrograph for the dark phase of plug 3 at a


magnification 7,000X

74

Non-dispersive X-ray scan of the dark surface area of


plug 3 indicating qualitatively the composition

74

SEM micrograph for the white spot of plug 3 at a


magnification 7,000X

75

Non-dispersive X-ray scan of the white spot surface area


of plug 3 indicating qualitatively the composition

75

Non-dispersive X-ray scan of the white powder of plug 3


indicating qualitatively the composition

76

LIST OF ILLUSTRATIONS - Continued


Page

Figure
4.1.7(a)

SEM micrograph for a grain of plug 4 at a magnification of 65X

77

4.1.7(b,c) SEM micrographs for the bright phase of plug 4 at


magnifications of b) 7,000X, and c) 20,000X

78

4.1.7(d,e) Non-dispersive X-ray scan of the bright surface area of d) 7,000X,


and e) 20,000X of plug 4 indicating qualitatively the composition..

79

Non-dispersive X-ray scan of the dark surface area of


plug 4 indicating qualitatively the composition

81

4.1.7(f)

4.1.7(g)

5.2.1

SEM micrograph for the dark phase of plug 4 at a


magnification 7,000X

81

Saturation index diagram of alunogen and chalcanthite

97

5.3.1(a,b) Ionic strength versus time from a) PHRQPITZ, and

b) PHREEQE for batch dissolution experiments

101

5.3.2

Saturation index of mirabilite versus time

102

5.3.3

Saturation index of jurbanite versus time

103

5.3.4

Three dimensional equilibrium diagram for antlerite

104

5.3.5

Three dimensional equilibrium diagram for brochantite

105

5.3.6

Three dimensional equilibrium diagram for langite

106

5.4.1

Eh-pH relationship for Eh estimated from the Sato equation


(Eh-s) and from WATEQ4F simulating a set of analytical data
from acid mine drainage in west Squaw Creek with similar
conditions as those of Magma samples (Eh-w)

109

5.4.2(a,b) Saturation index (SI) versus pH for silica bearing

minerals from a) WATEQ4F and b) SOLMINEQ

118

10

LIST OF ILLUSTRATIONS - Continued


Page

Figure
5.4.3(a-c) Saturation index (SI) versus pH for aluminum bearing minerals

5.4.4

from a) SOLMINEQ, b) WATEQ4F, and c) PHRQPITZ

120

Log K and log TAP versus temperature for alunite


from WATEQ4F and SOLMINEQ

121

5.4.5(a-c) SI versus pH for iron bearing minerals from a) SOLMINEQ,

b) WATEQ4F using Eh-w, and c) WATEQ4F using Eh-s

123

5.4.6(a,b) Saturation index (SI) versus pH for jarosite mineral series


from WATEQ4F calculation, a) using Eh-w,

5.4.7

5.4.8

5.4.9

and b) using Eh-s

125

Saturation index (SI) versus pH for nontronite mineral


series from SOLMINEQ

126

Saturation index (SI) versus pH for copper-iron mineral


series from WATEQ4F by using Eh-w and Eh-s

128

Saturation index (SI) versus pH for copper bearing minerals

129

11

LIST OF TABLES
Page

Table
3.2.1

Solution composition for batch dissolution experiments

41

3.2.2

Solution composition for batch reaction of epoxy plugs

43

3.4.1

Parameters used in the atomic absorption analysis

47

4. Li

Chemical composition of acid solution after dissolution of


chrysocolla

51

Stoichiometry of proton consumption for producting 1 mole of


copper in the batch dissolution experiments

55

4.1.3

Rate constants determined from CGB equation

61

4.2.1

Analytical results for samples collected at the


San Manuel Copper Mine

82

5.1.1

Equations for calculation of activity coefficients

88

5.2.1

Ion interaction parameters for Cu-Al-SO4 system and


solubility products for chalcanthite and alunogen

94

Solubility data in Al2(SO4)3-CuSO4-H20 system


from Occleshaw (1925)

96

Equilibrium constants and reactions for antlerite, brochanite,


and langite (from the data base of WATEQ4F)

99

4.1.2

5.2.2

5.3.1

5.4.1

5.4.2

Summary of Eh values for samples from the San Manuel


Copper Mine

111

Densities from SOLM1NEQ calculation for samples from


the San Manuel Copper Mine

113

12

LIST OF TABLES
Page

Table
5.4.3

5.4.4

Identification numbers, reaction equations,


and equilibrium constants at 25C for minerals of interest
in WATEQ4F and SOLMINEQ

114

Values of ionic strength for field samples


calculated from the geochemical computer models

116

13

ABSTRACT

Batch experiments were conducted to determine the dissolution characteristics

of chrysocolla using a computer controlled p11-Stat titration system. Dissolution can

be described by the diffusion controlled shrinking core model with a 1:2 ratio of
copper to proton; rate was inversely proportional to pH and ionic strength.

Scanning electron microscope was used to investigate the possibility of


secondary mineral precipitation. Secondary precipitation was not sufficient to affect
the dissolution rate, however, the shrinking core model did indicate the formation of
a residual silica layer.

Chemical analysis of aqueous samples collected from the solution mining


operation at the San Manuel Copper Mine indicated that the major ions were Al, Fe,

Cu, Si02 and SO4. Analysis of the saturation index illustrated the potential for
formation of secondary surface coatings which may impact dissolution in the field.

Geochemical computer models, PHRQPITZ, PHREEQE, SOLMINEQ, and


WATEQ4F were selected to model both experimental and field results.

14

CHAPTER 1
INTRODUCTION

The colloquial term "Copper solution mining" can be defined as the


hydrometallurgical extraction of copper by leaching low-grade ore, abandoned mine

workings and waste materials. Presently this is an extremely important source of


copper for the mining industry. Reasons for the popularity of solution mining include

lower operating costs, lower energy and labor requirements, faster start-up time,
advances in solvent extraction electrowining (SX-EW), improved in-situ leaching
techniques, and less environmental impact as compared with the conventional mining,

milling, smelting, and refining process for declining ore grades (Hiskey, 1986).
Recycling spent acid, a process which is desirable to decrease the cost of leaching
production and avoid stream pollution (Malouf, 1972), is employed by most solution

mining operations. Unfortunately, there is little information on the geochemistry of

recycling solutions involved in copper leaching, especially high ionic strength and
chemically complex solutions (Bassett and Hiskey, 1989).

The purposes of this research are to determine the reaction rates for the
dissolution of chrysocolla, a representative of oxide ore deposits in the Southwestern

United States, under laboratory conditions; and to investigate actual field leach
solutions for solubility controls on solution compositions. A better understanding of

the leaching process will facilitate optimizing the dissolution of chrysocolla by acid

15

leach solutions.

The research is divided into three parts: laboratory dissolution studies,


examination of actual field samples, and geochemical computer modeling.

1.1 Dissolution Studies

The objective of the laboratory dissolution studies was to examine the effect

of pH, ionic strength and secondary mineral precipitation on the rate of leaching of

copper under one set of circumstance in the mining operation.

Chrysocolla

(CuO SiO2 2H20) was chosen for this study because of its dominant presence in
many actual oxide leaching operations. Batch experiments were conducted utilizing

a computer-controlled titration apparatus which maintained a constant pH and


monitored the amount of acid titrant delivered during the experimental operation.
Batch experiments were conducted with different electrolyte solution compositions

and concentrations to investigate the influence on the rate of dissolution. The


parameters of interest were pH, ionic strength, and precipitated surface coatings.
Potential secondary mineral precipitation on the mineral surface was investigated
using a scanning electron microscope.

16

1.2 Field Study

The field study was conducted to determine the effect of a geologic medium

on the composition of the leaching solution. Samples were obtained from the site
of a leaching operation conducted by Magma Copper Co. in San Manuel, Arizona.

Open pit heap leaching and underground in-situ leaching operations are both
conducted at this site. The principal ore mineral being leached at this mine is
chrysocolla, and the lixiviant is sulfuric acid. Figure 1.2.1 shows a schematic layout
of a typical solution mining system, similar to the one at the San Manuel Mine. The

open pit is a medium-sized operation with a daily production of 28,000 tons of ore
and 68,000 tons of waste. The leach heaps, each of which contains 110,000 tons of

ore with a surface area of 125,000 ft2, cover a 230 acre plot. An impervious
polyethylene pad underlies the ore. The leaching rate is 0.8 gallons per minute per
100 ft2 of area.

In-situ leaching sites locate below the open pit area and above the
impoverished section of the underground mine. Injection wells cased with PVC pipe

with slotted bottoms are drilled to a depth of 1000 ft. The copper-bearing solution,
i.e. pregnant leach solution (PLS), is pumped to a 10,000,000 gallon feed pond, along

with PLS from dump leaching. The combined solutions are eventually sent to a SX-

EW plant for copper recovery while raffinate (barren solution) is discharged to the

reffinate pond as a recycling lixiviant (Magma Copper Company, 1991). Samples

17

RAFFUIATE
POND

P15 POND

PLANT FE

POND

SX PLANT

Figure 1.2.1 Layout of a typical copper solution mining system (Magma Copper
Company, 1991).

18

were collected from different location of the operation system for chemical analysis.

1.3 Computer Modeling

The computer simulation of the field data and laboratory experiments was
done in two stages. The first was a modification of the thermodynamic data files of
a mass transfer computer model PHRQPITZ (Plummer et al., 1988). Aluminum and

copper were added to the data base so that minerals containing the elements could

be evaluated. The determination of thermodynamic activity of aqueous species in


PHRQPITZ relies on the ion-interaction theory, which is a semi-empirical extension

to the classic Debty Hckel theory. The required parameters for aluminum and
copper, as well as thermodynamic solubility product constants for minerals of interest

were obtained from the literature. Validation of the modified PHRQP1TZ program
was done by comparing computed solutions to experimentally derived solubility data

from the literature.


Secondly, SOLMINEQ (Kharaka et al., 1988), PHREEQE (Parkhurst et al.,

1980), WATBQ4F (Ball and Nordstrom, 1991), and the modified PHRQP1TZ
computer codes are used to interpret the laboratory and field results. SOLMINEQ

is used to calculate the density values for all samples based on computed total
dissolved solids. Computed results from the different models are then compared.

Two capabilities of the above programs are speciation (most useful in this

19

study were computation of activities) and determination of saturation states of


minerals. Speciation is the distribution of the total elemental concentration among

aqueous species for a given chemical analysis, at a specified pH and temperature.

SOLMINEQ, PHREEQE and WATEQ4F calculate speciation by solving a set of


mass-action, mass-balance and oxidation-reduction equations using the ion-association

theory. PHRQPITZ calculates thermodynamic activity without speciation by using

the ion-interaction theory. The saturation index or thermodynamic affinity of all

relative minerals is computed for each sample solution. The saturation index
indicates which minerals have the potentials to precipitate or dissolve under the
conditions of the sample. The variables of pH, Eh, temperature, and solution
composition can be adjusted to determine the conditions most efficient for leaching
operation.

Several codes are used to model the lab and field samples because there is
no single model suited to answer the questions proposed in this study. For example,
PHRQPITZ works well for solution ionic strengths up to 20 molal for some ions, but

is limited to 14 elements and their corresponding minerals, and does not have the
capability of adjusting the redox state. The other three codes are suitable for ionic
strength up to near 1 molal but can be adjusted for variable Eh and complex solution

composition. The thermodynamic data bases contain equilibrium constants which


differ among models. The redox function in WATEQ4F is used to estimate the Eh
values for field samples based on iron ratios, as well as to calculate speciation and

20

saturation indexes using input Eh values.

21

CHAPTER 2

LITERATURE REVIEW

2.1 Copper Solution Mining

The art and the practice of leaching as a hydrometallurgical process for


copper recovery has a long history. According to records, heap leaching was applied

at Rio Tinto in Spain more than 300 years ago (Taylor and Whelan, 1942). With the

rising cost of the conventional mining process for declining ore grades, solution
mining is becoming more and more popular in the copper industry. The manifold,

intricate processes involved in leaching have not been well understood, and are
presently being researched.

Hiskey (1986) clearly divided copper solution mining into three categories:
heap leaching, dump leaching, and in-situ leaching.

In heap leaching, as shown in Figure 2.1.1, copper bearing ores are first
crushed and then piled on an impervious pad to heights from 9 to 45 feet. Leaching

solutions are delivered to the surface of heaps through pipe lines and sprayed out.

Leaching solutions then percolate through the ore becoming the pregnant leach
solution (PLS), which is finally sent to a copper recovery plant through a collection
channel (Hiskey, 1986).

Dump leaching, as shown in Figure 2.1.2, is similar to heap leaching, but is

22

Make Up
Reagents

COPPER
RECOVERY

PLANT

Cu

Figure 2.1.1 General layout of a copper heap leaching system (Hiskey, 1986).

r5

Recovery

plant

02 Depleted Air
V

Leach
Solution

REST

Bedrock

and br
Fresh

Soil

Air

LEACH SOLUTION

PERCOLATING DOWNWARD
PREGNANT SOLUTION
COLLECTION DAM
AND POND

Figure 2.1.2 Cross section of a copper waste dump showing one lift under active

leaching and another in a rest period (Hiskey, 1986).

23

built on the original bedrock/soil with run-of-mine size and longer leach cycle. The

leaching material is usually waste rock, containing copper less than 0.2% (Hiskey,
1986).

In-situ leaching is employed on an original mineral deposit and is classified


into three types as shown in Figure 2.1.3 (Wadsworth, 1979).

Type I refers to the leaching of a fractured ore body near the surface and
above the water table, including underground stopes and the remnant subsidence

zones. Similar to dump and heap leaching, lixiviant is applied to the ore through
injection holes or directly at the surface. PLS is collected in sumps or dams built in
the underground workings.

Type Ills represented by supergene deposits which are relatively shallow at

a depth of less than 500 ft either above or below the water table.
Type III is the leaching of deep hypogene deposits, which exist at a depth of

more than 500 ft below the water table.


In the recovery plant, PLS solution is first concentrated and purified using ion

exchange or solvent extraction. In solvent extraction, copper is extracted from PLS

solution by an organic extractant and then released into a rich solution of sulfuric
acid and copper sulfate. The two reactions are:

Extraction
Cu2 + (aq)

+ 2RH(org) = R2Cu(0) + 2H +

(aq)

(2.1.1)

24

- DEPOSrrS OCTENDING UNDER


WATER TABLE IN SUPERGENE

I- DL*IPS AND
DEPOSITS

ABOVE

ZONE

WATER TABLE

___.

m- oup

HYPOGENE DEPOSITS

t_i_.

IILII. Ito 5a

at

Id

-t-y/#I/A

it
-

I'

I'

U,

aC

(N

* '"ArIthcmI
aatar
taNs
* Rubbcsd (EWoIasa o Mlunq)
i4otraCtSd. 0 CflsmicOIIy
inicSd porosity

aC
0.

Rubbkd

jNydiotroct.d

Sr

chsicNty iidud

Figure 2.1.3 Three types of in-situ leaching (After Wadsowrth, 1979).

25

Stripping

R2Cu(0) + 2H(aq) = 2RH(0) +

2+

(aq)

(2.1.2)

The concentrated and purified copper solutions are recovered using


electrowinning method.

Electrowinning is an electrochemical process in which

copper is plated on a cathode from solution. The electrochemical reactions are:

+ 2e =

Cathode

(s)

H20 = 2H + 0.5 02 + 2e

Anode
Overall

Cu2 + H20

= CU(s) +

2H + 0.5 02

(2.1.3)
(2.1.4)
(2.1.5)

The organic extractant and reagents are recycled, and the raffinate is also
recycled as lixiviant.

Leaching is a complicated process, including hydrology, geology, surface


chemistry, thermodynamics, and heterogeneous kinetics (Bassett and Hiskey, 1989).

The previous work in terms of the mineral characterization and the chemical and
kinetic mechanism of the leaching of chrysocolla will be reviewed in Chapter 2.2 and
2.3.

26

2.2 Mineral Characterization of Chrysocolla

Chrysocolla is a secondary copper mineral commonly associated with


malachite, azurite, limonite, tenorite, native copper, geothite, manganese oxides, and

feldspar. Its properties depend on the origin of the sample. The color varies from

green to greenish-blue, sky blue, brown, and black, depending on the impurities.
(Dana, 1878; Windrell, 1942; Martinez, 1963; Raghavan, 1977).

The chemical formulas proposed vary, with CuO Si02 .2H2() being the most
commonly used, as well as Cu3 5(OH)2(A1Si3)010 -nH2O (Chukhnov and Anosou,
1950), 2CuO Si02 3H20 (Martinez, 1963), and Cu8 .(8i4010)2 -(OH)12 nH2O as a more

recently used stoichiometries, in which n equals 0, 4, and 8 (Van OosterwyckGastuche, 1970).

The structure was considered an amorphous gel until the late sixties. In 1970,

Van Oosterwyck-Gastruche proposed that chrysocolla is a definite microcrystalline


with a characteristic fibrous structure (0.5 to 2-3 u m in length, 6.5-7.0 nm in width,
and an inner striation 2 nm in diameter), an orthorhombic unit cell (a= 0.52, b = 1.77,

and c = 0.78-0.88 nm), and a combination of sheet and chain silicate structure. The

octahedral layer is filled with copper and to a limited extent is replaced by


aluminum, protons or divalent cations (Figure 2.2.1).

Advanced techniques, including X-ray diffraction, infrared spectroscopy,


differential thermal analysis and electron microprobe analysis, have been used to

27

C.
C

0
0
0

o
OH
H20
Si

0 above the plain

ointheplain
0 under the plain
Si above the plain
Si under th plain

Figure 2.2.1 Chrysocolla structure (After Van Oosterwyck-Gastruche, 1970).

28

characterize chrysocolla. (Martinez,1963; Newbery 1967; and Raghavan 1977). The

results revealed that chrysocolla is chemically heterogeneous, containing a large


amount of both chemically and physically adsorbed water.

Prosser et at. (1965)

obtained 40% of pore volume in the chrysocolla particle using adsorption-desorption

isotherms of water vapor. Pohiman (1976) concluded that chrysocolla has a large
surface area, an average pore size of 1.3 to 1.6 nm, and a specific gravity of 2.4, using

adsorption-desorption of nitrogen gas.


The copper content of chrysocolla varies from 23% to 37%, depending on the

source and impurities. Pohlman (1974) stated that the chrysocolla has a copper
content of 34.2% by weight, or chrysocolla of 94% purity with only trace amounts of

Al, Ca, Fe, and Mg. Raghavan (1977) reported the copper contents from 23% to

27% depend on the different sizes of chrysocolla. Sullivan reported 36.14% of


copper in chrysocolla.

2.3 Chemical and Kinetic Mechanism of Chrysocolla Dissolution

The general leaching reaction of chrysocolla by sulfuric acid is:

CuO'Si02 '2H20(s) +2H(q) =CU) +Si02 .n1120(s) +(3 -n)H2O(liQ)

(2.3.1)

In a stoichiometric reaction, two moles of protons should be consumed to

29

produce each mole of copper. Pholman and Olson (1976) stated that proton
consumption varies linearly with regard to copper production. Averill (1976)
obtained a 1:2.2 ratio of copper to proton. Sullivan (1933) reported that the average
Cu to H ratio was 1 to 1.54 by weight by investigating five chrysocolla samples.

The leaching of chrysocolla is a fluid-solid reaction in which a product layer


is

formed around an unreacted shrinking core.

The leaching process is

heterogeneous, involving the following steps as shown in Figure 2.3.1 (Wadsworth,


1972; Sohn 1979):

External Mass Transfer. The reactant H is transferred from the bulk of

the fluid to the external surface of the solid particle and the product Cu2 goes
opposite.

For example, the mass transfer of product Cu

is described by the

following equation,
cu =km(CcuCcu)

where

is the flux of Cu, km is the mass transfer coefficient, C

(2.3.2)

and CCUb are the

surface and bulk concentration of Cu, respectively.


Diffusion Through the Product Layer.

The remaining silica lattice provides

a diffusion barrier with a moving reaction boundary, which is described by Fick's law

in the following form,

n=-DVC

(2.3.3)

30

Flux outward

Distance

Figure 2.3.1 Shrinking core model for dissolution of chrysocolla.

31

where

11Cu

is the flux of Cu, VC

is the concentration gradient in the product layer,

and D is the effective diffusivity of Cu. D is proportional to porosity (cp) and


inversely proportional to tortuosity (r).
3) Chemical Reaction.

The chemical reaction between fluid and solid

occurs on the surface of the shrinking unreacted core.


The slowest step will control the dissolution rate. The overall dissolution rate,
however, is frequently not single rate controlled because of more or less equal effects

of several steps. The rate controlling step will change once reaction conditions
change;

therefore, the obtained rate information must be specified under its

conditions.

Valensi (1935) mathematically derived a diffusion control model in a study of

"Kinetics of Oxidation of Metallic Spherule and Powders". Later Ginstling and


Brounshtein (1950) proposed a simplified diffusion control model for a sphere solid

by neglecting the volume change during the reaction, and by assuming the constant
lixiviant concentration. In 1957, Crank published the same equation of diffusion
control model as Ginstling and Brounshtein. Now, the so-called Crank, Ginstling and

Brounshtein's (CGB) diffusion equation (Habashi, 1969) is widely applied to kinetic

studies as reactions controlled by diffusion. It is expressed as:

1 --cc (1 a)213
3

(2.3.4)

32

where a refers to the fraction reacted, t is time, and k is a rate constant, which is
proportional to the effective diffusivity.
Wadsworth (1972) published a reaction zone model (Fig. 2.3.2) based on the

steady-state conditions in which the rate of diffusion through the porous production

layer equals the rate of reaction at the interface between the unreacted core and the

reactant. He also assumed that the mineral is isotropic, the lixiviant concentration

is constant, and the leach solution is well agitated. The reaction zone model is
mathematically expressed as:
-(1 _)2/3I3[1 -(1 _)1/3] =k't

(2.3.5)

where a is the fraction reacted, $ is a constant, k' is a rate constant, and t is time.

The first part represents the diffusion and the second part refers to the surface
reaction.
Braun et al. (1974) confirmed that the model was well correlated with leaching

of copper sulfide ores. Pohiman and Olson (1974), using different sizes of particles

and a continuous flow system to analogous in-situ leaching, demonstrated that the
chrysocolla dissolution kinetics were correlated by the diffusion controlling model

(Eqn. 2.3.4), but did not go through the origin. The reaction zone model (Eqn.
2.3.5), however, correlated the experimental data as well as passed through the zero

time. Their results showed that initially the chemical reaction at the interface of

33

Diffusion of soluble products outward

Figure 2.3.2 Reaction zone model (After Wadsworth, 1972).

34

fluid and solid is significant but the diffusion rapidly become controlling, especially

for large particles. Their results also showed that the dissolution rate is inversely
proportional to the particle size.

Hsu and Murr

(1975)

proposed an oppositing opinion. They indicated that

the acid leaching of chiysocolla follows the shrinking core model, but the dissolution

rate is controlled by chemical reaction rather than the diffusion.

Roman et al.

(1974)

developed a computer model for the leaching of oxide

copper heaps as the rate of dissolution of copper is controlled by the diffusion of the

reactant through the product layer. They simplified the leaching process into a onedimensional system by dividing the heap into several equal-size square columns. The

model was verified by two column tests with malachite ore. They presented a sharp
boundary between the unreacted core and the leached rock, therefore concluding that

the diffusion limited the rate of reaction. Their results showed that the effects of
acid strength, size distribution, and dump height on the recovery of copper. The
particle size did not influence copper recovery during the early stages of leaching.

Finer size particles, however, obtained a higher recovery over a longer period of
time. Increasing acid concentration increases the rate of copper dissolution during

the whole leaching period. Higher heap has lower initial recovery rate, but higher
total production in a given time period.
Shafer et al.

(1979)

applied the shrinking core model to copper oxide leaching

of different grain sizes by using the column experiments. They indicated that copper

35

recovery depends on the particle size distribution and uniformity, and the
permeability of the rock.

Pohlman and Olson (1976) recognized that after leaching with acid, the
structure changes from microcrystalline to an amorphous gel with a lower surface
area. By investigating chrysocolla in Cuajone ores, Horlick (1981) confirmed a
marked surface structural change as hydrogen ions replace the copper during sulfuric

acid leaching. Raghvan and Gajam (1986) used nitrogen gas adsorption-desorption
technique to measure surface area and pore volume of chrysocolla which was leached

with ammonia. From this work, they concluded that pores are enlarged and the
porosity is increased to a maximum of 60% recovery during leaching process. Their

experimental results agree with Petersen's enlarging pore model (Petersen, 1957),
which considers the changes in pore structure during reaction under the assumption

of constant particle size, up to 45% conversion.

Wadworth (1972) also indicated that the leaching process is temperature


dependent in terms of changes in activation energy. Pohlman and Olson (1974)
calculated the activation energy of the experimental data and confirmed that the
leaching of chrysocolla was followed by the reaction zone shrinking core model.

Mena and Olson (1985) demonstrated that higher copper recovery is obtained at
higher temperature during leaching of chrysocolla by ammonia-ammonium carbonate
solution.

Khalezov (1984) confirmed the diffusion regime during the leaching of

36

chrysocolla but proposed that the silica acid that films on the surface of chrysocolla

has a negative effect by inhibiting the diffusion of the solvent to the surface.

Auck and Wadsworth (1972) demonstrated that the copper recovery is a


function of flow velocity, particle size, and pH in leaching, using column leaching

studies for porphyry copper ores; and concluded that higher recovery obtained at
faster flow velocity, smaller particle size, and lower pH.

Shelly (1975) indicated that H2SO4 was shown to be superior to the


ammonium solution as a lixiviant by comparing leaching of oxide copper using H2SO4
and NH3-NH4CO3.

So far, the effectiveness of ionic strength on the dissolution rate of chrysocolla

and the effects of precipitation of secondary minerals during the leaching of


chrysocolla has not been reported.

37

CHAPTER 3

METhODS

Standard laboratory procedures and standard methods of analysis were used


where applicable in all cases. Pyrex volumetric glassware was used in the preparation

of standard solutions, and new Nalgene plastic bottles were used for storage of
reagents, standards and samples. All glassware and plasticware was cleaned as
follows: it was first washed with soap, followed by a tap water rinse, an acid rinse

with approximately 1 N HNO3, and a final rinse three times with deionized water (a
four-chambered ion exchange purification system by millipore, refers to as Milli-Q).

All standards and solutions were prepared with Milli-Q. All chemicals were ACS
reagent grade.

Solid compounds were weighed using METFLER balances, model Ri 3000


with an accuracy of 0.Olg or model AE 240 with an accuracy of 0.0000lg, depending

on the accuracy required. A Thermolyne Stir Plate (Nuova2) was used when
necessary, to ensure thorough mixing of solutions.

Sample pH was measured using a pH meter (Accumet 950) with a Ross


combination electrode. Buffers of pH 4 and 7 were used to calibrate the pH meter
before and after measurements.

38

3.1 Chrysocolla Sample Preparation

A relatively pure chrysocolla specimen was obtained from Wards Natural


Science Establishment, Inc. It was crushed and sieved into two size fractions, the
smaller size was 60x150 mesh and the larger one was 16x20 mesh, i.e. particle sizes
at diameters between 16.93 to 42.33 m and 127.00 to 158.75 Mm, respectively. The

smaller size fraction was retained for subsequent batch leaching experiments, the
higher surface area would react more rapidly with sulfuric acid under the different
experimental conditions. The larger size fraction was used for making cylindrically
shaped samples or "plugs", which were convenient for SEM examination.

The following procedure was used to fabricate the epoxy plugs. Chrysocolla

particles were spread into the bottom of the cylindrical plastic container (2.5 cm

diameter/2.5 cm depth); the inner wall of the containers were pretreated with
Buehler Release Agent to prevent sticking. Epoxy, which was made from mixing of

1 part of Buehler Epoxide Hardener to 5 parts of Buehler Epoxide Resin, was then
slowly poured into the plastic container to a depth of approximately 1 cm to envelop

the chrysocolla particles. Vacuum extraction was used to prevent air bubbles from
forming in the epoxy. The samples were allowed to harden for 24 hours. The upper

surface of each plug sample was polished until a cross section of the chrysocolla
particles was exposed on the surface, and the surface was smooth with no visible
scratches.

39

3.2 Batch Dissolution Experiment

Batch experiments were designed for investigating two different processes: 1)

determine the kinetics of dissolution at different pH and ionic strength, and 2)


determine the morphology of particle surface after acid dissolution. A schematic
diagram is shown of the experimental setup for the investigation of dissolution of

chrysocolla particles (Figure 3.2.1).

The reaction pH was controlled using a

computer program pH-STAT (Appendix A), which delivers precise volumes of acid

with a spring pump as directed by a microcomputer. The reaction vessel was a five-

necked round bottom glass flask filled with approximately 400 ml of the initial
solution. To lower the pH to the desired value, sulfuric acid was delivered into the
flask using a 665 Dosimat syringe pump system. A stirrer (model RZR1) was used
to stir the chrysocolla slurry. A pH meter (Accumet) with 4-digital accuracy, and a
pH glass electrode combination (Orion Research) was used to monitor the solution

pH. A large circular constant temperature glass water bath with heating, model
1266-00 Immersion Circulator and a built-in thermometer, was used to maintain a

constant temperature of 25 C. For each dissolution experiment, one gram of


chrysocolla sample was added into the flask. To monitor the extent of reaction,
aliquot of 5 mls each were removed periodically using a volumetric pipet with a glass

frit filter tip. A total of 11 dissolution experiments were conducted. The solution

compositions and concentrations of each experiment are listed in Table 3.2.1.

40

Stirrer

Pipet is withdrawn to extract samples

Heating

Call

pH

electrode

Five-neck

Acid syringe

flask

p urn p

Filter tipped
pipet

Constant temperature water batch

pH meter

Figure 3.2.1

Printer
Computer

Schematic diagram of batch experimental apparatus for chrysocolla


dissolution experiments.

41

Table 3.2.1 Solution composition for batch dissolution experiments


Exp. No.

Solution Composition

Conc. (M)

pH

Na2SO4

0.1

Na2SO4

1.0

Na2SO4

2.0

MgSO4

0.5

Na2SO4

1.0

MgSO4

0.5

Al2(SO4)3

0.5

Milli-Q

10

KC1

0.3

11

KC1

3.0

42

Depending on the experiments, the reaction period varied from 1 to 3 hours and the

total number of sample aliquot withdrawn varied from 6 to 15 for any given
experiment.

The second set of experiments were conducted on the epoxy embodied plugs

to examine the surface morphology of the reacted particles of chrysocolla. The


experimental apparatus for these reactions were similar to the dissolution study. The

principal difference in the setup is that a beaker is used as a reaction vessel instead
of a five-neck flask, because the sample plug was too large to pass through the neck

of the round bottom flask. The sample plug was placed in the beaker to react for
one hour under the same conditions as the dissolution studies at the desired constant

pH value and 25C to obtain mid-reaction conditions. The intent was to observe
surface degradation and the sites of extensive leaching. To obtain a dry surface after

reaction, samples were removed from solution, washed with methanol 3 times,
immersed in methanol for 2 hours to remove the residual water, and then finally air

dried. It was assumed that methanol was non-reactive with the chrysocolla surface.
Three plugs were reacted under different solution compositions and concentrations
(Table 3.2.2).

3.3 Field Sampling

Field samples were taken from four locations in the solution mining operation

43

Table 3.2.2 Solution composition for batch reaction of epoxy plugs. (All experiments

were conducted at 25C and pH 2).


Conc. (M)

Exp. No.

Plug No.

Composition

12

Milli-Q

13

MgSO4

0.5

Na2504

2.0

Al2(SO4)3

0.5

MgSO4

0.5

Na2SO4

1.0

14

44

at Magma Copper Mine, San Manuel, Arizona. As shown in Figure 1.2.1, sample 1
was taken from a pregnant leached solution (PLS) pond (Site 1), sample 2 was taken
from a collecting channel draining the leaching operation (site 2), sample 3 was taken

from a pipe on the surface cariying leach solution pumped from a subsurface leach
operation (Site 3), and sample 4 was also taken from a pipe carrying leach solution
from a second subsurface leach operation (Site 4).

Samples were collected in new Nalgene plastic bottles without additional


acidification, because all samples were approximately pH 2.0 under solution mining

conditions. The temperature was measured in the field. The bottles were prerinsed
three times with the solutions to be sampled. Samples were pressed filtered through
a

im pore-size cellulose acetate membrane, refrigerated and stored for

subsequent analysis.

3.4 Chemical Analysis

The anionic concentration of F, Cl, Bf, SO42 in the field samples was
measured by the author using ion chromatography (IC), with an SP8700 solvent

delivery system, an SP8700 organizer, a DIONBX conductivity detector, and a


SP4100 computing integrator. Instrument parameters were optimized to acquire
optimum sensitivity. The eluent was composed of a 1:1 mixture of Na2CO3 (3.6mM)

and NaHCO3 (3.4mM), and had been degassed with helium prior to being pumped

45

through the column to ensure that air was excluded from the column. The anion
micro-membrane suppressor column was regenerated with 0.025N H2SO4.

The

column pressure was maintained at approximately 900 psi, and the flow rate of the
eluent was adjusted to approximately 2.0 ml/min.

A multi-component standard stock solution (500 mg/L) was made by


dissolving NaF, NaC1, NaBr and Na2SO4 in Milli-Q A series of four standards at
concentrations of 2.5, 1.5, 0.5, and 0.25 M were made from the mixed stock solution

for the anions of inherent by the serial dilution method. These mixed solutions were
used to create calibration curves for fluoride, chloride, bromide and sulfate by using

least squares linear regression. Samples were diluted using serial dilution until they

were in the range of standards. The peak height method was used to measure both
standards and samples.

The concentrations of the elements Cu, Ni, Na, K, Ca, and Mg in the field

samples were measured by the author using an AAS (Atomic absorption flame
spectrometer, Perkin Elmer 3100). Data collection and instrument parameter were
controlled by a Digital DEC Station 316sX computer, and linked to an Epson LQ-

850 printer. The computed concentrations of samples were reported directly.


The concentrations of all standard stock solutions were 1000 ppm within 0.5%

error and in 2% HNO3. Working standards were made from stock solutions using
serial dilution technique. Three standards were used to create the calibration curves.

The concentrations of standards used for constructing the nonlinear calibration

46

curves were determined as follows: the first standard was at the maximum value of

the linear range, the second standard was three times the concentration of the first,

the third standard was two times of the second standard. Instrumental parameters
of atomic absorption spectrometry were optimized to obtain best absorbance; the

pressure of air was 60 psi, and the acetylene fuel was set up at 12 psi. The other

parameters of operation for each element are summarized in Table 3.4.1. The
original samples were diluted using serial dilution until the results were within the
detecting range.

Analysis for the following elements Co, Al, Fe, U, Y, and Si in the field
samples were done commercially by Skyline Labs, Inc., using inductively coupled
plasma atomic emission spectrometry (ICP-AES).

In a total of 11 experiments, the concentration of copper in samples from 6


experiments were analyzed by the author using AAS, Perkin Elmer 2380 for Exp. 1,

4, and 8; Perkin Elmer 3100 for Exp. 9, 10 and 11. Samples of Exp. 7 were
determined by another analyst, Damaris Chong Diaz, using AAS at the University
of Arizona. The samples of 4 experiments, Exp. 2, 3, 5, and 6, were sent to Skyline
Labs, Inc. One sample of dissolved chrysocolla in acid solution was also analyzed by

Skyline Labs, Inc. for the following elements: As, Fe, Al, Na, Mg, K, and Ca.

47

Table 3.4.1 Parameters used in the atomic absorption analysis


Name

Calib. Type

Stds (ppm)

Wavelength

Slit

(3 each)

(nm)

(nm)

Lamp Current
(mA)

Cu

nonlinear

5,15,30

324.8

0.7

30

Ni

nonlinear

10,30,60

341.5

0.2

30

Na

linear

10,20,30

330.2

0.7

12

linear

50,100,200

404.4

0.7

12

Ca

nonlinear

5, 15,30

422.7

0.7

30

Mg

nonlinear

10,30,60

285.2

0.7

30

48

3.5 Surface Analysis by Scanning Electron Microscope

A JEOL JSM-840A scanning electron microscope was used to take


micrographs of four plug samples to examine the morphology of the chrysocolla
grains before and after reaction. Prior to analysis by SEM, samples were first coated

with a 200 A layer of gold-palladium in a Himmel sputter-coater, then mounted in

the sample chamber. An accelerating voltage of 15 KV was used to take the


micrographs at different magnification settings. Elements in the samples were
identified by the energy dispersive X-ray attachment and reported as weight percent.

3.6 Calculation of Copper Extraction

The copper extraction (as) is defined as the mass of copper dissolved at a


given time, divided by the total mass (mg) of copper in the initial chrysocolla sample

(Cu0). The general equation is expressed as:

[Cu]11)x1O6

([Cu](V,,-

(3.6.1)

Cu0

where [Cu] = the concentration of dissolved copper at t (mg/L);


t =

the period from the time chrysocolla is added into the flask to the

49

time the

sample is taken (minutes);

VS = the volume of the

11th

sample taken (mis);

V = V0 + Vaddn (mis);
V0 = the volume of solution in the reaction flask before the chrysocolla
was added (mis);

Vaddn = the total volume of acid delivered into the reaction flask at t.

50

CHAPTER 4
RESULTS AND DISCUSSION

4.1 Batch Experimental Results and Discussion

4.1.1 Chemical Composition of Chrysocolla

A chemical analysis of the chrysocolla specimen used in this batch dissolution

study was obtained by dissolving 1 gram of chrysocolla in 100 ml 6N nitric acid and

analyzing the resultant solution. The analytical results are listed in Table 4.1.1.
Copper is 30.5% by weight with only insignificant impurity, which is comparable to

the literature results (Reference Ch. 2.2).

4.1.2 Stoichiometry of Proton Consumption

Theoretically, the leaching of chrysocolla by sulfuric acid is a stoichiometric

reaction in which two moles of proton are consumed with the release of each mole

of copper. This relationship is confirmed by batch dissolution experiments. Figure


4.1.1(a-k) illustrates the results of 11 experiments at different pH values (2, 3 & 4)
and ionic strengths ('.'0, 0.3, 3, 9 & 14 m). Least squares linear regression is used to

correlate the experimental data. The reaction conditions and the results of the

51

Table 4.1.1. Chemical composition of acid solution after dissolution of chrysocolla

Element

Formula

Conc. (ppm)

Copper

Cu

3050.0

Silica

Si

6.00

Aluminum

Al

1.0

Arsenic

As

165

Calcium

Ca

50

Iron

Fe

10.0

Potassium

<1

Magnesium

Mg

9.0

Sodium

Na

1.2

52
0.8
0.6
H

d) pH=4. 1M N2SO4

1.5

0.4
0.2

y2.52l*-O.O52 15,r=O9949

J0.5

0.2o

0.5o

05

05
Copper (mmol)

Copper (mmol)
b) pH3, .1M N2SO4

y=2.027x+-O. 1501

0_o_ -

pH=3. 1MN2SO4

4
3

0.9988

- y2.792x+-O. 1754.=O9c6

0410

0
10

0.5

0
10

15

0.5

Copper (mmol)

15

Copper (mmol)

I
2

1.5

2.5

2
2.5
Copper (mmol)

1.5

Copper (mmol)
) pH=2, 2M N2SO4-.-.5M MpSO4

15
0-

3h) pH=Z, 1M N2SO4-.- 3M MSO4.-.5M Al2(SO4)3

y3.O62x*-171. rO.9785

-o y=L727x+-O9490,r=O9805

y=2.5O9x-t-2.985, r=0.9976

10
H

y2O82+-1.494rO.99O7

00

Copper (mmol)

1.5

Copper (mmol)

Figure 4.1.1 (a-h) The relationship between the production of copper to the
consumption of proton from batch dissolution experiments
at different pH values and solution compositions.

53

pIi2, Milli-Q

a)

0.5

1.5

2.5

2.5

2.5

Copper (mmol)
b) pH=2 0.3M KC1

0.5

1.5

Copper (mmol)
c)

pH=2, 3M KC1

y=2.208x+- 1 .221,r=0.9967

2
0
0.5

1.5

Copper (mmol)

Figure 4.1.1 (i-k) The relationship between the production of copper to the
consumption of proton from the batch dissolution experiments
at different pH values and solution compositions (continual).

54

regression analysis for each case are summarized in Table 4.1.2. The correlation

coefficient varies from 0.9498 to 0.9993 with a mean of 0.989 1 and a standard
deviation of 0.0142. The ratio of proton to copper varies from 1.606 to 3.062 with

a mean of 2.288 and a standard deviation of 0.435 for the 11 experiments. This is
similar to the results presented by Averill (1976) with a proton to copper ratio of 2.2.

The proton to copper ratio decreases slightly with increasing pH in dilute solutions
(0.1M Na2SO4).

These results are reasonable because the dissolution rate of

chrysocolla is controlled principally by proton attack as described in Eqn. 2.3.1.


Figures 4.1.1(g,h) show a break in the slope of the proton to copper ratio; the initial

slope is steeper in Figure 4.1.1(g) as opposed to a lower initial slope in Figure


4.1.1(h). It implies that in the early stage of dissolution reaction the rate is changing.

The effect of ionic strength on the proton to copper ratio is not obvious. Higher
proton consumption in some cases may be due to the impurity of the chrysocolla
grains.

4.1.3 Copper Extraction and Rate of Dissolution

During the solution recycling process, pH and ionic strength will increase as

acid is consumed and minerals continue to dissolve.

The batch dissolution

experiments performed in this study are related to the field solution recycling process

and were designed to examine the two factors affecting the dissolution rate for

55

Table 4.1.2 Stoichiometry of proton consumption for producing 1 mole of copper in


the batch dissolution experiments.
No

pH

Solution

Conc.(M)

H/ Cu

I (m)

Na2SO4

0.1

2.180

0.9980

0.3

Na2SO4

0.1

2.027

0.9988

0.3

Na2SO

0.1

1.606

0.9992

0.3

Na2SO4

2.749

0.9830

Na2SO

2.792

0.9960

0.3

Na2SO4

2.521

0.9949

0.3

Na2SO4

0.9785

MgSO4

0.5

3.062
2.509

Na2SO4

0.9860

14

MgSO4

0.5

1.727
2.082

Al2(SO4)3

0.5
-

1.977

0.9498

Milli-Q

10

KC1

0.3

2.3 15

0.9993

0.3

11

KC1

2.208

0.9967

56

sulfuric acid leaching of chrysocolla at a constant temperature of 25 C and a well


agitated condition without the complication of other mineral phases being presented.

The dissolution rates are determined experimentally and then fit to a diffusion
controlled shrinking core model.
The first group of experiments was conducted with the same aqueous solution
concentration (0.1 M Na2SO4) but with different pH values (2, 3, and 4). The results

are shown in Figure 4.1.2(a). As can be seen, the copper extraction rate is strongly

pH dependent; the higher the pH, the lower the copper extraction percent.

As shown in Figure 4.1.2(b), the pH-dependant relationship was again


demonstrated by leaching at a tenfold increase of sodium sulfate concentration (1M

Na2SO4), keeping the rest of the conditions similar to those of the first group. The
copper extraction percentage is similar between solutions of different ionic strengths.

pH value represents the negative log activity, i.e. the thermodynamically


effective concentration, of protons. Higher pH has a lower activity of the proton, and

a stoichiometric reaction releases a lower activity of copper. If the ionic strength is


kept constant, the activity coefficient of copper will remain constant; therefore, the
concentration of copper will decrease.

As a result of this pH dependence, all subsequent experiments dealing with

the ionic strength effect on the dissolution rate were conducted at pH 2, which is
about the operating pH at the San Manuel Copper Mine, as well as in most of the
other mining operations.

The results are divided into two sets: group 3 (0.1M

57

Na2SO4=.1M

o 08

opH2

1 oI

+pH4

x pT13

0
0
0

00

20

40

80
100 120
Time (minutes)

60

0.6

0.8
0.6

0.4

+pfl4

0.6

180

200

140

160

180

200

140

160

180

200

0
0

20

40

60

80
100
120
Time (minutes)

140

pH=2
* 1M Na2SO4 + .5M MgSO4 + 5M Al2(SO4)3
c)

+ 2MN2SO4+.5MMgSO4
x 1M Na2SO4

oO.IMNa2SO4

+
*

20

40

60

80

100

120

Time (minutes)
d) pH=2
o MiTh-Q

x 03MKC1
+3MKC1
0

00

160

0.2
00

200

x pH3

00

180

opH2

0.4

0.2

160

140

Na2SO4=1M

0.8

20

40

60

80
100
120
Time (minutes)

Figure 4.1.2(a-d) Copper extraction versus time for batch dissolutionexperiinents.

58

Na2SO4, 1M Na2SO4, O.5M MgSO4+2M Na2SO4, and O.5M Al2(SO4)3+O.5M

MgSO4+ 1M Na2SO4) and group 4 (Milli-Q, O.3M KC1, and 3M KCl).

The

corresponding ionic strength for each experiment is calculated from the computer
models PHREEQE and PHRQPITZ, depending on the ionic strength range. As can

be observed in Figure 4.1.2(c), the highest ionic strength (I=14m) results in the
lowest dissolution rate.

The middle value of ionic strength (I=9m) has an

intermediate copper recovery. The highest fraction of copper extracted can be


reached either at ionic strength 0.3 or 3 m, but no differences between them can be
observed. Another set of experiments shows similar results, as can be seen in Figure

4.1.2(d). The amount of copper extraction is almost the same between 0.3 M and 3

M KC1, but it reveals a higher value using distilled water. The absolute values
between the two groups can not be compared because the experimental conditions
were not exactly the same.

The results also show a tendency for decreasing fraction of extracted copper
with time. This phenomena is compatible with the diffusion controlled shrinking core
model.

The surface area of an unreacted chrysocolla core reduces with time,

correspondingly, the diffusing proton requires more time to penetrate the increasingly

thick porous product layer to reach the reaction interface Similarly the copper ion
requires more time to diffuse back to the bulk solution. The experimental data were

plotted according to CGB equation in Figures 4.1.3(a-d).

Excellent fits were

obtained when the data points were correlated by the least squares linear regression.

59

O.1M Na2Sp4

a)

0.1

-o pH2.y=4.612e-4x+1.8004-3.r=0.9967
0.08
C

-x pH3,y=6. 154o-5x-2.25-4.r=0.9990

0.06 - .. pH4.y6.92&-6+-5.756o-5.0.9974

0.04
0.02
ii.

40

20

60

80

120

1CID

140

160

180

200

140

160

180

206

120

140

160

180

2(X)

120

140

160

180

206

Time (mrnutes)

1M Na2SO4

0.1

-o p112. y&425e-4x+2.064-3,r=0.9926
x pH3.y=4.544o-5x-2.08S.4,r=0.9993

0.08

0.06 -

-+- pH4,yr5.234C-6x-1.266-5,,=O.9992

0.04

0.02

0
0

20

40

60

..

80

---+-106

120

------.-

Time (minutes)

C)

0.1

p}l=2

-* 1M Na2SO4+.5M MgSO4.5M Al2(SO4)3,y=1.708o-4x-1.024.3.rC0.9983


0.08
0.06

--I.- 2M Na2SO4+.5M MgSO43.425c-4x-6.284-4.r0.9975


-x- 1M Na2SO4,y=4.425-4x+2.064.3.r0.9926
-o 0. 1M Na2SO4,y=4.612c-4x+1.809-3,r0.9967

0.04
0.02

20

40

60

80

100

Time (minut)

d)

0.08

pH=2

-o Milli-Q,y=3.053o-4x1 .382o-3,0.9922
0.06

-x 0.3M KC1.y=2.277-4x+1.142-3.r0.9960
-+- 3M KC1,y=2.42-4x+1.O79-3,rrO.998O

0.04
0.02

20

40

60

80

100

Time (minut)

Figure 4.1 .3(a-d) Diffusion mode correlating data from the batch dissolution
experiments.

60

The rate constant for each case is tabulated in Table 4.1.3. The pH dependent
relationship is obvious; the dissolution rate has a tenfold difference when the pH
value changes one unit, and the lowest dissolution rate occurs at the highest pH value

for the supporting electrolyte, either 0.1 M Na2SO4 or 1 M Na2SO4. The rate
differences among different ionic strengths are within the same order of magnitude.

At constant pH of 2, the activity of the proton in solution will be constant.


The same activity of copper, therefore, should result according to the stoichiometric
reaction. The reason for the decreasing dissolution rate of copper with the increasing

ionic strength is not understood at present but may be related to the following
mechanisms:

The activity coefficient for the copper decreases with ionic strength increase

until higher ionic strengths are reached (larger than 3 m in this study), at which it
may increase again. It, therefore, leads to the decrease of copper concentration.
The indistinguishable dissolution rate between ionic strengths of 0.3 to 3 m indicates

that the activity coefficient of copper is insensitive to the changing of ionic strength
at this range. At lower ionic strength, represented by Milli-Q plus some sulfuric acid
at pH 2 in this study, the activity coefficient of copper is close to unity, therefore, the

concentration of copper approaches its activity. The maximum concentration of


copper then can be obtained.
The tortuosity of the travel path through the residual layer of chrysocolla
may increase while the porosity of chrysocolla decreases, with an increase in ionic

61

Table 4.1.3 Rate constants determined from CGB equation.


Exp. No.

Solution
Composition

Conc.
(M)

pH

Rate (k)

(m)

(E-4)

(min')
1

Na2SO4

0.1

0.3

4.6 12

.6154

.0693

Na2SO4

1.0

4.425

.4544

.0523

14

3.425

1.708

3.503

KCI

0.3

0.3

2.277

KC1

3.0

2.420

Na2SO4

2.0

MgSO4

0.5

Na2SO4

1.0

Mg504

0.5

Al2(SO4)3

0.5

Milli-Q

10
11

62

strength, because some secondary minerals may precipitate inside the pores of the

product layer. The effective diffusivity of copper would, therefore, decrease. This
would lead to the decrease of the rate constant (k) because it is proportional to the
effective diffusivity. No secondary minerals were observed on the chrysocolla surface,

but the possibility of precipitation within the residual layer cannot be eliminated.

4.1.4 Morphology

The surfaces of the chrysocolla specimens, one unreacted plug and three

reacted plugs under conditions similar to those of the particle reactions, were
examined by the scanning electron microscope. The plugs were photographed at

different magnifications then subjected to an X-ray scan of the surface area to


qualitatively identify the composition.

Plug 1 - Before reaction

Figure 4.1.4(a) shows the chrysocolla specimen before reaction at a


magnification of 3000X.

Figures 4.1.4(b,c) are the X-ray scans of the surface

indicating a qualitative composition of the exposed surface. The bright phase is


composed principally of copper with a minor amount of silica and the dark phase is
mainly composed of silica with a minor amount of copper.

63

Figure 4.1.4(a)

SEM micrograph for plug 1 at a magnification of 3,000X.

64

BRIGHT PHASE

C
U

U
D

Figure 4.1.4(b,c)

C
U

VF

Non-dispersive X-ray scan of b) the bright, and c) the dark


surface area of plug 1 indicating qualitatively the composition.

65

Plug 2 - After reaction with H2SO4 + Milli-Q at pH 2 for one hour


Two different grains are observed in plug 2 after reaction, as shown in Figures

4.1.5(a,b) at a magnification of 65X. The bright phase of grain 1 appears brighter.

Figures 4.1.5(c-f) show the bright phase and the dark phase of grains 1 & 2 at a
magnification of 7000X. It can be seen that the bright phase of grain 1 has been
highly dissolved. No secondary minerals are visible. The X-ray scan for the surface
of the bright phase of grain 1 indicates that it is composed of Cu, Ca, and Al (Figure

4.1.5(g)). Ca and Al are derived from the impurities in chrysocolla. The X-ray scan
for the bright phase of grain 2 indicates major amounts of Si and minor amounts of
Cu (Figure 4.1.5(h)). The SEM photo clearly shows a tight silica product layer with

few fractures. No secondary precipitation can be observed on either phase. The


small amount of copper remaining in both bright phases implies incomplete reaction.

The X-ray scan for dark phases of both grains shows only pure Si remaining (Figure
4.1.5(i,j)).

This verifies that the small amount of copper in dark phases has

completely dissolved.

The tiny holes were formed as a result of the copper

dissolution.

Plug 3 - After reaction with H2SO4 + 3M MgSO4 + 2M Na2SO4 at pH 2 for one


hour.

Figure 4.1.6(a) shows a typical grain with bright phase, dark phase, and some

white spots after reaction and rinsing with methanol.

Figure 4.1.6(b) is the

67

Figure 4.1.5(c,d)

SEM micrographs for bright phases of c) grain 1, and d) grain


2 of plug 2 at a magnification of 7,000X.

69

LI

C
H

LI

LU4

5t1PLE 2 IHITE FIAEA

LI
C

U \.j

.L).44

C
U

51_E 2 BRIGHT PHASE GRAIN *2

Figure 4.1.5(g,h)

Non-dispersive X-ray scan of the bright surface area of g) grain


1,

and h) grain 2 of plug 2 indicating qualitatively the

composition.

70

SIRMPLE 2 DAR}< P!4E

SIRMPLE 2 OFiK PI-4E RIN *2

Figure 4.1.5(i,j)

Non-dispersive X-ray scan of the dark surface area of i) grain

1, and j) grain 2 of plug 2 indicating qualitatively the

composition.

72

Figure 4.1.6(b)

SEM micrograph for the bright phase of plug 3 at a


magnification 7,000X.

AMFLE 3 BRIGHT PHASE

Figrue 4.1.6(c)

Non-dispersive X-ray scan of the bright surface area of plug 3


indicating qualitatively the composition.

73

enlargement of the bright phase at a magnification of 7,000X. As can be seen, the


bright phase is an etched surface; pores are large and clean. The X-ray scan for the
surface (Figure 4.1.6(c)) indicates that it is mainly composed of the porous Si residual

product layer with a small amount of Cu, and minor Mg impurity. It implies that the

copper dissolution is almost completed in the bright phase

The dark phase is

enlarged to a magnification of 7,000X as shown in Figure 4.1.6(d). The surface


appears less reacted. The X-ray scan of the dark phase indicates that it is composed
principally of Si with minor amounts of Cu (Figure 4.1.6(e)). Figure 4.1.6(f) shows

a large white precipitant on the surface of the dark phase. The X-ray scan verifies
that it is a secondary mineral, composed of Si, Cu, Na, Mg, and S (Figure 4.1.6(g)).

The X-ray scan for one of the small white spot shows about half the amount of Na

and S with minor amounts of Mg (Figure 4.1.6(h)). The small white spots were
formed by the supersaturation of the supporting electrolyte solution due to water
release after rinsing with methanol.

Plug 4 - After reaction with H2SO4 + O.5M MgSO4 + O.5M Al2(SO4)3 + 1M Na2SO4

at pH 2 for one hour.

Figure 4.1.7(a) shows a representative grain after the reaction. The bright
phase is enlarged to a magnification of 7,000X (Figure 4.1.7(b)) and 20,000X (Figure

4.1.7(c)). It can be seen that the mineral surface is clean and the pores are large and
connected to each other. The X-ray scan for both photos (Figures 4.1.7(d,e)) reveals

5,

ir

75

Figure 4.1.6(f)

SEM micrograph for the white spot of plug 3 at a magnification


7,000X.

SAMPLE 3 PRECIPITATE

Figure 4.1.6(g)

Non-dispersive X-ray scan of the white spot surface area of plug


3 indicating qualitatively the composition.

76

POJOR SRIIPLE

Figure 4.1.6(h)

Non-dispersive X-ray scan of the white powder of plug 3


indicating qualitatively the composition.

77

Figrue 4.1.7(a)

SEM micrograph for a grain of plug 4 at a magnification of


65X.

78

Figure 4.1.7(b,c)

SEM micrographs for the bright phase


magnifications of b) 7,000X, and c) 20,000X.

of plug 4

at

79

C
F

RIHT P1E +4OLE FEA

Figure 4.1.7(d,e)

Non-dispersive X-ray scan of the bright surface area of d)


7,000X, and e) 20,000X of plug 4 indicating qualitatively the
composition.

80

that they are composed of Cu, Ca, and Al, indicating that the chrysocolla is not pure.

The X-ray scan for the dark phase (Figure 4.1.7(1)) identifies that it is mainly
composed of Si with a miner amount of S. From Figure 4.1.7(g), a SEM photo for
the dark phase at magnification of 7,000X, it can be seen that the surface is tight and

uniform, with only few tiny holes on the surface but not linked to each other. No
precipitation is observed.

4.2 Aqueous Chemistry Results for Field Samples

Chemical analyses of water samples collected from San Manuel Copper Mine

are summarized in Table 4.2.1. pH values vary from 2.0 to 3.0 and temperature
varies from 14.5 to 20.5 C. Themajor cations are Mg, Al, Fe, and Cu, and SO4 is

the dominate anion. Aluminum might come from the dissolution of aluminumbearing minerals such as gibbsite or feldspar. Iron may be due to the dissolution of

iron bearing minerals such as limonite or geothite. High concentration of copper


indicates that the leaching is effective in extracting the copper from the oxide ores.

Magnesium in the leaching solution may imply that there are some Mg-bearing
minerals associated with chrysocolla, which have not been identified.

The concentrations of dissolved species at sites 1 and 2 are similar. This is


reasonable since the site 1 includes a combination of pregnant leach solution from
site 2, along with the leach solutions from other leaching areas. The temperature of

CD

p
0

CD

CD
CD

0CD

CD

<p

p
p

-.

rn

01

:'

82

Table 4.2.1 Analytical results for samples collected at the San Manuel Copper Mine.
(The concentration of dissolved species is given as mg/L).
Site 1

Site 2

Site 3

Site 4

pH

2.02

2.28

2.22

2.97

T (C)

16.5

20.5

24.0

14.5

Ca

20

20

26

41

Mg

3600

3800

2800

1700

Na

250

255

226

184

138

136

74

34

Fe

920

770

390

155

Cu

1111

1170

2144

1351

Al

3200

3500

2350

1050

Cl

63

66

55

42

SO4

49820

43404

36438

18685

Br

11

11

10

Ni

29

31

24

15

Co

37

39

30

21

Si

80

65

88

50

152

162

102

43

83

site 2 represents the temperature of the atmosphere at the time of sampling.


Samples from sites 3 and 4 were derived from the in-situ leaching of a site

with different mineral composition from that of the dump leaching site (site 2).
Samples from sites 3 and site 4 should be similar, because both are in-situ leaching

operations; however, significant differences are found in the concentrations of the


dissolved species, as well as pH and temperature. Solubilities of most minerals are

related to temperature; in general, lower temperature leads to a decrease in the


solubility of minerals. pH is of course a master variable in the leaching process. A

pH of approximate 3 at site 4 is higher than the optimal leaching pH of 2. Redorange secondary iron coatings were observed at site 4 during field sampling, as well

as on the filter after treatment of the sample. The precipitation of ferric hydroxide

would be expected at this higher pH; as a result the measured dissolved iron
concentration is the lowest in this sample. Generally, Eh is proportional to the pH.
Ferric iron is more stable as iron hydroxide at higher Eh. Furthermore, dissolution
of most minerals is acid consuming; therefore, higher pH would dissolve minerals less
efficiently.

K, Na, Ca, Ni, Co, Br, F, Cl and trace amounts of U and Y detected in
samples of all sites indicate that ores are not pure. Extraction of metals such as Ni

and Co as by-product from mine tailings may be sources of Ni and Co. Trace
amounts of U and Y are not concentrated enough to be valuable, but the potential
impact on the environmental contamination cannot be neglected.

84

CHAPTER 5

GEOCHEMICAL COMPUTER MODELING

The purpose of this chapter is to use geochemical computer models to


interpret the geochemistry involved in the dissolution of chrysocolla in the laboratory

as well as in the field. The goal is a better understanding of the mechanism for
dissolution of chrysocolla and subsequently one should be able to predict the optimal

solution composition for copper solution mining of oxide ores.


The chapter is divided into four sections. The first section provides a general

explanation of the geochemical models used in this study. The second section

describes the modification required to update the computer code PHRQPITZ.

Aluminum and copper were added to the data base and then the model was
calibrated with data from the literature for solubilities of minerals in the Cu-Al-SO4
system.

The third and forth sections discuss the application of the models in

interpreting the experimental and field phenomena.

5.1 Background

The four geochemical models, PHRQPITZ, PHREEQE, SOLMINEQ, and

WATEQ4F are used in this study to compute the aqueous ion speciation and the
mineral saturation indices of experimental and field aqueous samples.

85

Speciation is the equilibrium distribution of aqueous species between free ions

and ion pairs or complexs when using the ion association theory (Nordstrom, 1986).
Speciation is not physically defined in the ion interaction theory used in PHRQPITZ.

Saturation index (SI) represents the degree to which a water sample is saturated with

respect to a specific mineral or phase. It is defined as,


lAP
Sl=logK

(5.1.1)

in which the JAY is the ion activity product and K is the equilibrium constant for the

mineral of interest, which is adjusted for temperature.

When the SI is greater than zero the water composition computes to be


supersaturated with respect to the mineral, and the mineral has the potential to
precipitate.

It indicates that the water composition may be thermodynamically

supersaturated but may remain so as a result of kinetic hindrances. When the SI is

below zero the solution is undersaturated with respect to the mineral, and the
mineral would have the potential to dissolve. When the SI is zero, the solution is
defined to be at equilibrium with the mineral of interest.
Four models are used in this study because each has a specific application for
the appropriate ionic strength range, implicit theory, or specific thermodynamic data
base.

No single model has the breadth of capability required for solving all

experimental situations.

86

The most important difference among these models is the method of


calculation for the activity coefficient. SOLMINEQ, PHREEQE, and WATEQ4F
use ion association theory, while PHRQPITZ uses ion interaction theory.

5.1.1 Ion Association Theory

Ion association model (TAM) theory (Garrels and Christ, 1965; and Helgeson,

1967) requires the distribution of aqueous species for a given composition of water

at specified pH, temperature, and Eh (if redox reaction is considered) to be


calculated by solving all the appropriate mass-action and mass-balance equations.
Consider the reaction,

aA+bB=cCdD

(5.1.2)

the corresponding mass action equation is,

K-

[qcfD]d

(5.1.3)

[A]z [B]1'

where { ] denotes ion activity. The thermodynamic activity or effective concentration

equals the concentration of the elemental ion (i) in molality (mi) times the activity

coefficient (y.).

There are numerous expressions for computing the activity

87

coefficient; most are extensions of the Debye-HUckel model (1923), and generally

work well only for dilute solutions; some may be extended for use up to an ionic
strength of near one molal. These activity coefficient equations are summarized in
Table 5.1.1. In the table, a. is the hydrated radius of the particular ion. B is a

fitting constant, which depends on the temperature, the species and the ionic
strength; b1 is a WATEQ constant, and A and B are the molal Debye-Huckel

coefficients at temperature T. This defines the effect of the solvent, given by,
l.8248xlO6p'
(7)3/2

B(T)-

5O.29x1O8p1

(5.1.4)

(5.1.5)

(e 7) 1(2

where c is the dielectric constant of water at temperature T, and p refers to the


density of water at temperature T.
The ionic strength (I) is defined as,

m1z,

(5.1.6)

where m is the concentration of species i in molality and z1 represents the charge on

that species. The ionic strength is often considered as an index representing the

88

Table 5.1.1 Equations for Calculation of Activity Coefficients


Name

Equation

DebycHUckel

logy1-

WATEQ
DebyeHckel

logy1

Used in
Codes

References

Range

PHREEQE

DebyeHckel 1923

I<O.lm

WATEQ4F

Truesdell
and Jones,

-Az,2[I
1 +Ba Of!

_42/i

PHREEQE

1974

1 +Ba OfI

B method

SOLMINEQ

logy1-

-Az/7

+B OJ

Lewis and
Randall,

1961;

Helgeson,

1 +Ba fI

1969

Davies
2

logy1= -A; (

fi
i+ri

0.31)

WATEQ4F
PHREEQE

Davies, 1962

89

electrostatic effectiveness of polyvalent ions in solution.

Different geochemical

models calculate the ionic strength in different ways. lAM codes calculate the ionic
strength after speciation adjustment. The formation of complexes, which sometimes

have zero charge, will decrease the value of ionic strength. The result of activity
calculated based on this ionic strength, however, still represents the most true value,
because the ion association model has taken into account of all of free ions and their

complexes. PHRQPITZ computes the ionic strength only on the basis of the input
species concentrations and charges (more to IIM section). WATEQ4F reports the
ionic strength in both ways, i.e. the effective ionic strength and the total analytical
ionic strength. The value of effective ionic strength is used to calculate the activity

coefficient in WAIEQ4F.
The general mass balance equation is expressed as,
I

m,,=> ni/nj

(5.1.7)

where mit is the analytical (total) concentration in molality of the component i, n


refers to the stoichiometric coefficient of component i in species j, and mj denotes
the computed concentration in molality of species j. For a system of i components

and j species, i mass-balance and j-i mass action equations need to be solved
simultaneously. The final results are obtained by iteration to the extent that the
residual is within the desired range.

90

5.1.2 Ion Interaction Theory

Ion interaction model (IIM) theory (Pitzer, 1973; Pitzer and Mayorga, 1973,
1974; Pitzer and Kim, 1974: and Pitzer, 1975) does not explicitly compute aqueous

complexes rather it parameterized the ionic evaluation with a separate equation for

each element. It takes into account not only the long range effects for dilute
solutions from Deaby-Huckel theory, but also the short range effects of concentrated

solutions at high ionic strength. Harvie and Weare (1980) have tested the accuracy
of the IIM in a Na-K-Mg-Ca-Cl-SO4-H20 system using mineral solubility data, and

verified that this model performs well under ionic strengths from 0.06 to 20 molal.
IIM equations, based on a virial expansion of the excess energy, calculate the

osmotic coefficient

the activity coefficient of the cation

M'

and the activity

coefficient of the anion x as follows:


2

(4>-i)(

m.)

A1312
1 +bJt2

mmI'I ,+ mrcc 'a+


c<

m1Jr)]
C

a1

mC(2cZMC+> m0qr)

mmC

m4mCI1pFM+Iz
a1

(5.1.8)
mama

a<

lnyM=zMF+2 ma(2BMa+ZCMa)+
a<

mcma(B+ZC)

(5.1.9)

91

1ny=zF+

m(2B1ZC) +

ma(2'I1xa+y: md)

mmC
c<

(5.1.10)

where c, M, and c' represent cations, and a, X, and a' refer to anions. c < c' and a < a'

denotes all pairs of dissimilar cation and anions, respectively. The F term refers to
Debye-Huckel theory, representing the long-range ion interactions.

The coefficients B and C, referring to short-range ion interactions, are


parameterized from single-salt data. The data includes fitting parameters of/3
/3(2)

for the second virial coefficients and

j3(1)

refer to opposite signed ions.

/3(2),

d for the third virial coefficients.

j3(1),

j3(0) and

as a representative of ion pairing, is necessary

for higher valence electrolytes, but drops to zero for monovalent ions.
The short-range ion interactions for mixed electrolytes are represented by
the second virial coefficients, and i

the third virial coefficients. The former refers

to the interactions between anion and anion, or cation and cation. The latter
represents cation-cation-anion and anion-anion-cation interactions.

ijk

is

independent of ionic strength.

The thermodynamic data base is also an important factor when considering


which model to use. Different codes have different sources of thermodynamic data.

WATEQ4F includes more minerals than the others and most of the data have been

updated and reviewed recently. It may, therefore, be more reliable. The data base

92

in PHREEQE was primely derived from an earlier version of WATEQ4F.


PHRQPITZ does not required thermodynamic data for aqueous species, but has a

limited data base for minerals. SOLMINEQ has a large data base, many of the
constants are given for higher temperature and rely on estimation and approximation
(Bassett, 1992).

The application of each code is also different. SOLMINEQ has a specific

function to calculate the density on the basis of total dissolved solids, which is
especially useful for concentrated electrolytes with high density. WATEQ4F can
compute the solution Eh on the basis of preselected redox pairs, as well as compute

the distribution of redox species on the basis of input Eh values. SOLMINEQ can

only do the latter; PHRQPITZ has no redox function incorporated at present.


PHRQP1TZ is, however, the most accurate code for simulating high ionic strength
solutions. The study described below was intended to further expand the capabilities

of PHRQPITZ.

5.2 Calibration of the PHRQPITZ code in a Cu-Al-SO4 system

The PHRQPITZ code includes two data source files: PITZER.DATA and
PHRQPITZ.DATA.

parameters

The P1TZER DATA contains all Pitzer ion interaction

and the PRRQPITZ.DATA includes

element definition and

thermodynamic data for species and minerals. These two data files can be called

93

directly from the main code. The current version of PHRQPITZ code contains data

only for 14 elements: Ca, Mg, Na, K, Fe, Mn, Ba, Sr, Cl, C, S, B, U, and Br. The

ion interaction parameters for CuSO4 from Pitzerts study are only listed in the

appendix of this code but without validation. The purpose of this section is to
enlarge the working range for Al and Cu by calibrating the Cu-Al-SO4 data against

experimental solubility data so that experimental and field data from this
investigation may be modeled.
Reardon (1988) reevaluated the Pitzer ion interaction parameters for the Al2(SO4)3-H20 system by using all published isopiestic data at 25C. The revised values

are 0.854, 18.53, -500, and -0.0911 for/3, p(l), p(2), and C, respectively. The ion

interaction parameters p (0),

f3

(1),

and C

were obtained from regression analysis of

the osmotic coefficient data, while p (2) was estimated because it cannot be
determined definitively. In the mixed salt Cu-Al-SO4 systems, it was found that the
ternary interaction parameters 0Cu

and

Cu-A1-SO4

are zero and 0.035, respectively.

Two minerals, chalcanthite and alunogen, are used to test the model by
comparing the thermodynamic equilibrium constants (K6) of these minerals and the

model-calculated solubility products (lAP). Solubility data were derived from the
experimental results obtained by Occieshaw (1925). The ion interaction parameters

for the Cu-Al-SO4 system are added in the PITZER DATA (Appendix B) and the

mineral thermodynamic data of chalcanthite and alunogen are added in the


PHRQPITZ.DATA (Appendix C). The above data are summarized in Table 5.2.1.

94

Table 5.2.1 Ion interaction parameters for Cu-Al-SO4 system and solubility products

(k,) for chalcanthite and alunogen.


Name

Fomula

-log K,,

$()

p(1)

p(2)

c()

eCM

Chalcanthite

CuSO45H2O

2.621'

0.2342

2.5272

.45332

0.00442

0.00'

Alunogen

Al2(SO4)17H2O

6.204'

0.8541

18.53'

-500.0'

.0.09111

Note: 1 Reardon (1988),


2 Plummer (1988).

0.035'

95

The input file for PI{RQPITZ requires the concentration in molality of each
element. Occieshaw's solubility data are, however, reported in weight percentage of

aluminum sulfate and copper sulfate. The conversions from the weight percent to
molality are calculated using the following equation:

m- l000xA

(5.2.1)

B(100 -A)

where:

m = molality (moles of solute / kg of solvent);

A = weight percent of solute (kg solute / 100 kg solution);


B = molecular weight of solute (gram / mole).

The input concentrations for each element are listed in Table

5.2.2.

Theoretically, saturation indices of these two minerals should be zero, by using

the solubility data as the input concentrations; the results demonstrate an excellent

match. As shown in Figure


sulphate solution

(No.14),

5.2.1,

chalcanthite is in equilibrium in the pure copper

and close to equilibrium when the aluminum sulphate as

supporting electrolyte (from No.8 to No.13); Alunogen is close to equilibrium from

No.1 (pure aluminum sulphate solution) to

No.4

(copper sulphate is the supporting

electrolyte); both minerals are close to equilibrium from No.5 to No.7.

96

Table 5.2.2 Solubility data in Al2(SO4)3-CuSO4 -1120 System from Occleshaw's


(1925).
UNIT

WEIGHT PERCENT

MOLAUTY

(g solute/100 g solution)

(moles solute/kg solvent)

Al2(SO4)3

CuSO4

H20

Al

Cu

SO4

#1

27.80*

72.20

2.251

3.376

#2

26.60*

2.50

70.90

2.118

0.161

3.338

#3

25.66*

3.75

70.59

2.017

0.244

3.271

#4

25.64*

3.83

70.53

2.015

0.250

3.273

#5

24.86*

5.18*

69.96

1.934

0.342

3.243

#6

24.65*

5.32*

70.03

1.912

0.352

3.220

#7

24.32*

5.36*

70.32

1.878

0.355

3.173

#8

23.18

6.11*

70.71

1.764

0.408

3.053

#9

18.83

8.18*

72.99

1.356

0.558

2.592

#10

18.10

8.75*

73.15

1.292

0.601

2.539

#11

12.05

11.33*

75.62

0.877

0.801

2.117

#12

7.50

14.69*

77.81

0.474

1.079

1.790

#13

3.79

16.59*

79.62

0.230

1.246

1.592

#14

18.49*

81.51

1.421

1.421

* Equilibrated with the minerals.

97

Figure 5.2.1 Saturation index diagram of alunogen and chalcanthite.

98
5.3

Saturation Analysis for Batch Dissolution Experiments by PHREEQE and

PHRQPITZ

Geochemical computer models, PHREEQE and PHRQPITZ are used to


calculate the ionic strength and to examine the mineral saturation states of the batch

dissolution experiments to see whether the secondary minerals are precipitated to


influence the dissolution rate.

5.3.1

Model Input

In a total of eleven experiments, four (Experiments

1, 2, 3

and 10) with lower

solution ionic strength (approximately 0.3 m) were computed by PHREEQE, and the

rests were modeled by PHRQPITZ. Temperature, pH, concentration of each


element, and density of sequential samples in each experiment were input into the
computer models. The densities were precalculated from another code, SOLMINEQ.

The equilibrium constants of three copper sulphate hydroxyl minerals, antlerite,


brochantite, and langite (Table 5.3.1), taken from the data base of WATEQ4F, were

added into the PHIRQPITZ.DATA (appendix 3).

99

Table 5.3.1 Equilibrium constants and reactions for antlerite, brochantite, and langite

(from the data base of WATEQ4F).


Name

log K

Reaction

Antlerite

8.92

Cu3(OH)4SO4 + 4H

= 3Cu2

+ 4H20 + SO42

Brochantite

15.34

Cu4(OH)6SO4 + 6J1

= 4Cu2

+ 61120 + SO42

Langite

16.79

Cu4(OH)6SO4H20 + 6H

= 4Cu2

+ 71120 + SO42

100

5.3.2 Model Results

Figures 5.3.1(a, b) show the values of ionic strength from PHRQPITZ and
PHREEQE, respectively. It can be seen that the change in ionic strength during the

dissolution process is small. Exp 8 has the highest ionic strength.


In a total of eleven experiments, Exp 7 (2M Na2SO4 + 0.5M MgSO4) was found

to be little supersaturated with respect to mirabilite (Figure 5.3.2), while Exp 8 (1M
Na2SO4 + .5M MgSO4 + .5M Al2(SO4)3 was supersaturated with respect to jurbanite

(Figure 5.3.3). The modeling results from Exp 7 agreed with those from SEM

morphological examination. However, SEM had not found any precipitation in Exp

8. Jurbanite was also found to be a potential surface coating in the field leaching
process (Ch.5.4.3). The modeling results further verified that the secondary mineral
precipitation would not be serious enough to reduce the dissolution rate in the batch
dissolution condition.

Figure 5.3.4(a,b) intuitively shows the equilibrium surface boundary for


antlerite in three dimensions at different horizontal rotation and vertical elevation.

At lower pH, it will be undersaturated at most situation (below the surface plane),
except at higer pH value and extremely higher concentrations in copper and sulphate
(above the surface plane). Experimental results showed all to be undersaturated with
respect to antlerite (Figure 5.3.4(a,b)), as well as brochantite (Figure 5.3.5(a,b)), and

langite (Figure 5.3.6(a,b)).

101

a) Ionic strength of batch experiments from PHRQPITZ

15
*

* 1M Na2SO4+.5M MgSO4+.5M Al2(SO4)3pH*2


x 2M Na2SO4+.5M MgSO4,pH2

10

3M KCI,pH=2

o 1M N*2SO4,pH=2

--1MNa2SO4.pH=3

1MNa2SO4,pH=4

00

50

0.2
C,,

200

150

250

300

Time (mm)
b) Ionic strengh of batch experiments from PHREEQE

0.4
0.3

100

0 0 00

+ 1M Na2SO4,pH=3
* 1M N*2SO4,pH2

xxxxx

1M Na2SO4.pH=4

-r

o .3M KC1,pH=2
x MiIli-Q,pH2

0
0
00

50

100

150

200

250

Time (mm)

Figure 5.3.1(a,b) Ionic strength versus time from a) PHRQPITZ and


b) PHREEQE for batch dissolution experiments.

300

102

Mirabilite
0.5
* 1M Na2SO4+.5M MgSO4+.5M Al2(SO4)3,pH=2

V-

xx

x 2M Na2SO4+5M MgSO4,pHr2

- 0. 1M Na2SO4,pH=2

++

0.5 -

-- 0.1M Na2SO4,pH=3

- -- -.-.-...0.1MNa2SO4,pH=4
+

o 1M Na2SO4,pH2

1M Na2SO4,pH=3

+ 1M Na2SO4,pH=4

-1.5

50

100

150

200

Time (mm)

Figure 5.3.2

Saturation index of mirabilite versus time.

250

300

103

Jurbanite
3

2.5

* 1M Na2SO4+.5M MgSO4+.5M Al2(SO4)3,pH=2

0.5

00

20

40

60

80

Time (mm)
Figure 5.3.3

Saturation index of jurbanite versus time.

100

120

140

104

o Batch dissolution experimental data

86-

a)

Horizontal Rotation=30 degree


Vertical Elevation=20 degree

2-6

-4

-2

2 -6

-2

-4

log[SO4--}

log[Cu+]
8

b)
Horizontal Rotation=-10 degree
Vertical Elevation=10 degree

6
4
2

oq
-5

log[SO4--]

Figure 5.4.4(a,b)

-4

-3

-2

-1

log[Cu++]

Three dimensional equilibrium diagram for antlerite.

105

o Batch dissolution experimental data


Horizontal Rotation=40 degree

a)

Vertical Elevation20 degree

6-

2-4

-2

log[SO4--]
log[Cu++]
8

b)

Horizontal Rotation=-10 degree


Vertical Elevation=25 degree
4
2
x))

2
0

-2
-4

log[SO4--]

Figure 5.4.5(a,b)

-5

-4

-3

-2

log[Cu-i-+]

Three dimensional equilibrium diagram for brochantite.

106

o Batch dissolution experimental data


Horizontal Rotation=20 degree
Vertical Elevation=30 degree

-6

-5

-4

-3

log[SO4--]
log[Cu++]
8

b)

Horizontal Rotation=-5 degree


Vertical Elevation=30 degree

4
2

o I'iI))((sub)
o

-6
log[SO4--]

-6

-5

-4

-3

-2

-1

log[Cu-H-]

Figure 5.4.6(a,b)

Three dimensional equilibrium diagram for langite.

107

5.4 Saturation Analysis for Field Samples by Geochemical Computer Models

Geochemical computer models, SOLMINEQ, WATEQ4F, and PHRQP1TZ,

are used to compute the speciation, saturation indexes of minerals of field samples

to predict the potential for formation of surface coatings caused by the secondary
mineral precipitation during solution mining operation in the Sun Manuel Copper
Mine.

This section is divided into three parts. The first part describes the estimation

of the important parameter, Eh, for model input when redox species and relative
minerals are considered. The second part discusses the different model applications

and parameter inputs. The third part provides model results and interprets the
geochemistry involved in field samples.

5.4.1 Estimation of Eh

A redox potential (Eh) is required in order to effectively investigate the


oxidized mineral precipitation. Eh values have not been measured for solution
mining process, two Eh estimation methods are used here and compared with each
other.

The first method calculates Eh values from Sato's empirical equation (Sato,
1960),

108

Eh=O.682 -O.0591pH

(5.4.1)

on the basis of pH values measured in the field, which will be called Eh-s below.
The Eh-pH relationship from the Sato equation is shown in Figure 5.4.1. The Sato
equation is widely applied in weathering environments; and is essentially the standard

potential of the H202-02 couple. Sato used his experimental and field investigations

to conclude that 'the oxidation potential of an aqueous system should be at or above

the potential of H202-02 couple so long as a detectable amount of free oxygen is


present in the system" (Sato, 1960). He also gives an approximate upper marginal
value, 0.859 (v), as Eh-axis intercept.
The second method creates a relationship between Eh and pH by WATEQ4F

simulation using a set of analytical data from acid mine drainage in West Squaw
Creek, which has similar conditions as those of samples from San Manuel Copper
Mine. The analytical results from acid mine drainage in West Squaw Creek and its
tributaries (Filipek et al, 1987) are listed in Appendix D. A total of 32 sample sites
are reported with results for pH, conductivity, concentrations of total iron, Fe2, Mn,

Zn, Cu, Al, Na, K, Ca, Mg, SO42, HCO3, Cl-, F, Si02, and As. It can be seen that

the concentration of total iron at most sites with pH value around 2 to 3 are close
to the results from San Manuel Copper Mine. WATEQ4F calculates Ehon the basis

of the ratio of Fe37Fe2t We assume that the Eh from this computation represents

somewhat the redox potential of the solution mining process at the San Manuel

109

Eh-pH Diagram
1

0
Sample Data

WATEQ4F, 16 points, Eh=.864-.0395pH, r=.5

0.9 1NATEQ4F, 15 points, Eh=.91 0-.053pH, r=.7

16-point regression
0.8-

2q8

21

Sato Eqn.

0.7-

0.6-

1J
D

1 5-poInt regression

Sato Eqn. Eh=.682-.0591 pH


-

0.5-

0.4
2.2

2.4

2.6

2.8

3.2

3.4

3.6

3.8

PH

Figure 5.4.1 Eh-pH relationship for Eh estimated from the Sato equation (Eh-s)
and from WATEQ4F simulation of a set of chemical data from acid
mine drainage in West Squaw Creek with similar conditions as those
of Magma samples.

110

Copper Mine. The Eh calculated from this method will be called Eh-w. Some
sample sites were omitted because the measurement of Fe2 was inaccurate, making
the calculation of Eh meaningless. Site 23 was also neglected because of an obvious

analytical error: the concentration of total iron is less than that of Fe2 alone. Only
16 sample sites (marked with an asterisk in Appendix D) were input into WATEQ4F

to calculate the Eh. The Eh values computed from WATEQ4F versus measured pH

values are plotted in Figure 5.4.1. A least squares linear regression was applied to
fit these data. The correlation coefficient is 0.5, which is considered too low to be
accepted. Graphical analysis indicates that site 15 shows poor statistical fitting and
may be an outlier. After eliminating this point, a better correlation coefficient, 0.7,

is obtained. The regression equation is,


Eh =0.9 10 -0.0534pH

(5.4.2)

From Figure 5.4.1, it can be seen that the regression equation line for Eh-w

from acid mine drainage is almost parallel to the Sato equation line (the slope
difference is 9.6%), and the values of Eh-w are approximately 0.23 (v) higher than

those of Eh-s. It is in good agreement with Sato's results, but extends Sato's upper
boundary. The Eh values of Magma samples are calculated on the basis of the Sato
equation (Eqn. 5.4.1) and the regression equation (Eqn. 5.4.2). The estimated results

are summarized in Table 5.4.1.

111

Table 5.4.1 Summary of Eh values for samples from the San Manuel Coppper Mine
Site 1

Site 2

Site 3

Site 4

pH

2.016

2.283

2.217

2.966

Eh-w

.802

.788

.791

.751

Eh-s

.563

.547

.551

.507

Note: Eh-w means that the Eh values come from WATEQ4F calculation using a set
of analytical data from acid mine drainage in West Squaw Creek with similar

conditions as those of Magma samples; Eh-s means that the Eh values are
calculated from the Sato equation.

112

5.4.2 Model Input

Three of the geochemical codes, SOLMINEQ, WATEQ4F, and PHRQPITZ

are used to compute speciation and saturation indexes of field samples from San
Manuel Copper Mine. The model requires temperature, pH, and chemical analysis

data of given samples. The number and choice of elements which can be used is
different in each code. Of the total of 16 elements in the field samples (Table 4.2.1),

the concentrations of twelve elements, Ca, Mg, Na, K, Cl, SO4, Si02, Al, Cu, Fe, F

and U were input into SOLMINEQ because of the limitation. These same 12, plus
three others, Ni, U, and Br, were input into WATEQ4F. Only ten elements, Ca, Mg,

Na, K, Fe, Cu, Al, Cl, SO4, and Br can be calculated by PHRQPITZ. SOLMINEQ
used Eh-w values as system oxidation potentials (Table 5.4.1). WATEQ4F, however,

calculated the same samples twice by using different system Eh values, Eh-w and
Eh-s (Table 5.4.1), to see the Eh effects on the speciation and the mineral saturation

for the redox pairs. SOLMINEQ was used to calculate the density for all samples

first, and the results (Table 5.4.2) used as input values for WATEQ4F and
PHRQPITZ. Three minerals, A1(OH)3(a), gibbsite, and jurbanite, were added into

the PHRQPITZ.DATA (Appendix C) with the same thermodynamic data base as


those of WATEQ4F.

The reaction equations for the minerals of interest and their identification
numbers and equilibrium constants at 25C are summarized in Table 5.4.3.

113

Table 5.4.2 Densities from SOLMINEQ calculation for samples from the San Manuel

Copper Mine

Site Numbers

Density

1.0409

1.0368

1.0308

1.0164

114
Table 5.4.3 Identification numbers, reaction equations, and equilibrium constants at 25C for minerals
of interest in WATEQ4F and SOLMINEQ
Equatioo.

ID No.

WThQ

Io8ICT

SOLMINE

W&TEQ

SOLMINEO

10.8

-34.42

140

48

Ai(OH)3a

AI(OH)3+3H+AIS.f+3H20

50

AImit

KAI3(SO4)2(OH)6 K +3A13 +O42-+60W

-14

-2.79

97

26

Cba1ccdo

S102+21120=H4SiO4

.355

.3.73

99

37

Criatobalitc,a

Si02+2H20=H4S1O4

.3587

-354

38

Cristobalit

SiO2+2H2O=H,SiO4

-2.94

CupricFrrIt

cu1p4+sn+ =Cu2 +2Fe

249
229

181

+41120

CuF02+4H+=Cu++F+2H2O

.8.92

F(OH)27CI3

F(OH)22Ci3+2.7H

-3.04

Cup

aFrrit

+2.7H20+0.3Cf

112

Frrilydrit

F(OH)3+3H+F+3Hp

4551

51

49

Giitc

A1(OH)s+3H+=AJS++3H20

8.11

-34.92

110

152

Gothi11

F00H+3H'=Fe+2H2O

-1.0

0.48

108

148

1frmatit

F2O3+6H.2Fe+3Hp

-4.008

0.04

205

Jamait K

KF3(SO4)2(OH)6+6H+ = K+ +3F+ +2SO42 +61120

-14.8

337

JamaiI H

(H30)F3(SO4)2+5H' _3.?+ +260j +71120

.5.39

204

Ja,oaI(e Na

NaFc3(SO4),(OH)6+6H

-11.2

133

.9.93

+3Fe3' +20042 +61120

(K,Na 03(1130) 2]Fc3(SO42+5.&1 + -.77K +.O3Na+ +3F3

+25042 +6.2H20+6Hj)
-

471

3ubamt

A10HSO4+H =433+ +SO42+H

109

149

Maghm

F2O3+6H+ 2Fe+ +3H20

6.386

6.4

107

Magtite

F04+8H2F33++F2++4H20

3.737

86

No,ilmtht.Na

Na3FA1 Si3 62010(011)2+73250 + +2.68H20=33Na + 2F

41.48

.iisi

-12.5

33
-

87

NoHronit,K

+ 33A133 +3.67H4S104

K3sFA1Si3o20jo(OH)2+73211 + +2.681120 .33K +2F33


+ + .33A13

88

No,tmnit,H

+3.6711,45104

H33FeA1%60j0(OH)2+699H + +2.681120 3311+ +2F33


+ + .33A33

+367H.4SiO4

89

No11mmt,Ca

CaFAI37Si3670j0(OH)2+732H + +2.681120 .165Ca2+ +


2I
+,33A1'+3.67H4SiO4

-1154

90

Noatronile,Mg

Mg.j68FAI33S0j0(OH)2+732H +2.681120 .165Mg2

-1156

+2Fc' + + 33A13 +3.6711.45104


+41145404+211 +

53

Pymptruit

Al2S1040(OH)2+ 121120 .-2Ai(OH

101

103

Quartz

S102+2H20.H,4S104

-3318

-3.93

100

112

SIlic,(kI

Si02+2H=H,4SiO4

-3.018

.2.71

395

111

SiIk,Am

Si02+2H20.-H45iO4

-2.71

-2.7

.48.31

115

5.4.3 Model Results

Table 5.4.4 summarized the results of ionic strength obtained by above


geochemical computer models.

As we expected, the values of ionic strength from

SOLMINEQ are similar to the results of the effective ionic strength from
WATEQ4F, while the results from PHRQPITZ are close to the total ionic strength
values from WAIEQ4F. The miner differences between ionic strength values from

WATEQ4F using Eh-w and Eh-s are due to oxidation potential effects on the
distribution of redox pairs. The different values between different codes are mainly

the products of the different input elements, even though they are using the same
calculation methods. The value of total ionic strength calculated from WATEQ4F
probably represents the most accurate value of analytical ionic strength, since most

of the elements in samples have been input and the results are not adjusted by
complexes. The value of ionic strength varies from 0.77 molality (m) for sample
collected at site 4 to 2.0 m for sample collected at PLS pond (site 1). They are very

close between sites 1 and 2, because site 2 is part of input for site 1. The values of

ionic strength at sites 3 and 4 are less than those of sites 1 and 2. The differences
may mean that the dump leaching produces higher ionic strength than that of in-situ
leaching. The value of ionic strength at site 4 is less than that at site 3. It shows that
less minerals have been dissolved at second subsurface leach operation.

The mineral saturation states of field samples from modeled results are

116

Table 5.4.4. Values of ionic strength for field samples calculated from geochemical
computer models.
Ionic Strength

Site No.

PHRQPITZ

WATEQ4F

SOLMINEQ

Total

Effective

Eh-w

Eh-s

.638

1.991

.643

.660

1.879

.576

1.914

.569

.581

1.843

.498

1.498

.493

.500

1.430

.338

0.767

.313

.316

0.760

117

discussed by dividing the minerals into seven groups: Si-bearing minerals, Al-bearing

minerals, Fe-bearing minerals, jarosite series, nontronite series, Cu-Fe minerals, and
Cu-SO4 minerals.

Si minerals

Using WATEQ4F, five Si minerals, chalcedony, cristobalite (ci), quartz, silica

(amorphous), and silica (gel) are computed to be supersaturated at all sites as shown

in Figure 5.4.2(a). The SI decreases slightly with increasing pH. The results from
SOLMINEQ (Figure 5.4.2(b)) for Si-minerals are in agreement with the results from
WATEQ4E

SOLMINEQ also includes another silica polymorph in the

thermodynamic data base, cristobalite (a), which is also computed to be


supersaturated. The over saturation of the Si minerals are caused by the elevated

concentration of silica as a result of the continued dissolution of chrysocolla as


lixiviant is recycled. The above Si minerals, therefore, may precipitate with time,

when kinetics are feasible.

Surface analysis for reacted chrysocolla samples in

laboratory dissolution experiments has demonstrated that the silica remains as a


product layer. The field results are confirmed by the laboratory phenomena. Both
are following the leaching reaction of chrysocolla (Eqn.2.3.1).

Al-bearing minerals

The results from SOLMINEQ show that the solution is undersaturated with

118

Si-WATEQ

Chalcedony
Crlstoballte,alpha
-4wQuartz

-9Sillca,Gel

-9Sllica.Am

2.1

2.2 2.3

2.4 2.5 2.6 2.7 2.8 2.9

pH

Si-SOLMINEQ
-AChalcedony

-9Cristobalite,alpha

Cristobalitebeta

-9-

Quartz

-NSIlica,Ani

-8Sillca.Gel
2
4
2

2.1

2.2 2:3

2.4 2.5 2.6

2.7 2.8 2.9

pH

Figure 5.4.2(a,b)

Saturation index (SI) versus pH for silica bearing minerals from


a) WATEQ4F and b) SOLMINEQ.

119

respected to gibbsite and Al(OH) at all sites. It is unsaturated at site 1 with


respected to alunite, whereas close to equilibrium at site 3 and saturated at sites 2
and 4 (Figure 5.4.3(a)). These results suggest that if gibbsite and A1(OH)3a are

present at field sites, they could both dissolve and contribute Al and OH to the
solution, but the precipitation of these two minerals would be impossible. The
alunite may dissolve at site 1, maintain equilibrium at site 3, and precipitate at sites
2 and 4.

The results from WATEQ4F for Al-bearing minerals are plotted in Figure
5.4.3(b).

Gibbsite and Al(OH)3a do not saturate at all sites. It agrees with the

SOLMINEQ results but demonstrates a greater degree of undersaturation. In


contrast, the leaching solutions from all sampling sites are supersaturated with
respect to jurbanite.

It implies that jurbanite may precipitate with time as a

secondary mineral to occlude the pores or coat mineral surface during the solution

mining operation. Alunite, in contrast to the results from SOLMINEQ, does not
computed to be supersaturated at sites 1, 2 and 3, but is slightly supersaturated at site

4. The results suggest that alunite may not precipitate at sites 1, 2, and 3, but may
be lost from solution at site 4. The different results for alunite between WATEQ4F

and SOLMINEQ are controlled primarily by the differences in the log K values in
the models which are from different sources. Determining which modeled results are

more reliable will require an evaluation of the equilibrium constants. The values of

the L&P are similar (Figure 5.4.4), since they have similar activity coefficients

120

b)
AI-WATEQ4F
2

J(OH)3a

AlunS.
GIbbsIte

kjrbiit
Pyroph4Ute

2.1

2.2

2.4

Z5 ao 22 Z8

2.9

pH

Figure 5.4.3(a-c)

Saturation index (SI) versus pH for aluminum bearing minerals


from a) SOLMINEQ, b) WATEQ4F, and c) PHRQPITZ.

121

Alunite
IogK-Wateq4f
IogK-So!mineq
IogIAP-Wateq4f

-9-

IogIAP-Solmlneq

14 15 16 17 18 19 20

21

22

23

24

Temperature (oC)

Figure 5.4.4 Log K and log TAP versus temperature for alunite from WATEQ4F
and SOLMINEQ.

122

calculated by the ion-association method.

The results derived from PHRQPITZ are similar to those from WATEQ4F

and SOLMINEQ. Gibbsite and A1(OH)3a do not saturate, but jurbanite saturates

with larger SI values than those from WATEQ4F (Figure 5.4.3(c)). It further
demonstrates the potential of jurbanite precipitation as a secondary mineral.

Fe-bearing minerals

SOLMINEQ uses Eh-w to calculate the speciation of redox pairs. A review


of one type of the redox minerals, Fe-bearing minerals, indicates that hematite and

geothite saturate at all sites, while maghemite does not saturate at sites 1, 2, and 3,

but saturates at site 4 (Figure 5.4.5(a)). It implies that ferric-containing minerals

have thermodynamic potential to precipitate during present solution mining


operation, and are most likely to precipitate in the second subsurface operation (site
4).

WATEQ4F uses two different approaches for redox potential, Eh-s and Eh-w,

for determining the effects of Eh values on the distribution of redox species and
saturation states of redox minerals. It can be seen from Figure 5.4.5(b,c) that, in Femineral series, hematite, geothite, and Fe(OH)27C13 saturate, but ferrihydrite dose

not saturate, either at higher Eh (Eh-w) or lower Eh (Eh-s).

Magnetite and

maghemite does not saturate at lower Eh (Eh-s) at all sites, but saturates at site 4
at higher Eh (Eh-w). It implies that the potential precipitation of redox minerals can

123

Figure 5.4.5(a-c)

SI versus pH for iron bearing minerals from a) SOLMINEQ, b)


WATEQ4F using Eh-w, and c) WATEQ4F using Eh-s.

124

be avoided by adjusting the Eh at desirable range to decrease the saturation indices.

The SI results for hematite, geothite, and maghemite at higher Eh (Eh-w) are in
good agreement with those from SOLMINEQ. Each mineral has larger saturation
index value at higher Eh.

Jarosite Series

The Eh effect on the oxidized mineral saturation is also observed in the


jarosite mineral series. As can be seen in Figure 5.4.6(a), when higher Eh values
(Eh-w) are used by WATEQ4F, the solution is supersaturated with respect to jarosite

K and jarosite ss at all sampling sites, while is only saturated with respect to jarosite

Na at site 4, and is close to equilibrium with respected to jarosite H at site 4.


However, once the Eh values decrease when using Eh-s, the solution at all sites is
undersaturated with respect to all jarosite minerals at the same conditions (Figure
5.4.6(b)). It further demonstrates that the secondary mineral precipitation caused by
oversaturated oxidized species can be reduced if suitable system redox potential are

applied. The desired lower Eh values can be obtained by adding some reductant.

Nontronite Series

Only SOLMINEQ offers the thermodynamic data base of nontronite mineral

series. Nontronite is a M-Al-Fe-Si type of mineral. Here M is represented by Na,


K, H, Ca, and Mg. From Figure 5.4.7, all types of nontronite minerals supersaturate

125

a)

Jarosite-WATEQ4F(Eh-w)
Jarosite K

-.-

jarosite H

Jarosite Na

--

Jarosite 55

-8

2.1

2.2

2.3

2.4

2.5

2.6

2.7

2.8

2.9

pH

-.Jarosite K
Jarosite H

Jarosite Na
Jarosite 55

2.1

Figure 5.4.6(a,b)

2.2

2.3

2.4

2.5
pH

2.6

2.7

2.8

2.9

Saturation index (SI) versus pH for jarosite mineral series from


WATEQ4F calculation, a) using Eh-w, and b) using Eh-s.

126

Nont,Na

Nont, K

-*-

Nont,H

-8-

Nont,Ca

w-

Nont,Mg

2.1

2.2

2.3

2.4

2.5

2.6

2.7

2.8

2.9

pH

Figure 5.4.7 Saturation index (SI) versus pH for nontronite mineral series from
SOLMINEQ.

127

at all sites. It implies that nontronite minerals will also precipitate as secondary

minerals if the kinetics for these minerals are feasible.

They have increasing

saturation tendency with pH increase.

Cu-Fe minerals

WATEQ4F includes thermodynamic data base for cupric ferric and cuprous
ferric. At lower Eh (Eh-s), most Cu species exist in reduced states as Cut As

shown in Figure 5.4.8, cuprous ferric is the stable phase while cupric ferric does not
saturate at all sites. When higher Eh values (Eh-w) are used, however, both minerals

undersaturate at sites 1 and 2, but equilibrate at site 3, and supersaturate at site 4.

The Eh effect on the redox mineral precipitation is further demonstrated by the


above results. Corrected Eh values need to be measured or estimated to minimize
the secondary mineral precipitation caused by redox species. Therefore, the most
efficient leaching process can be predicted.

Cu-SO4 minerals

Figure 5.4.9 shows the SI values of Cu-SO4 minerals, chalcanthite, antlerite,

brochantite, and langite, from PHRQPITZ calculation. No Cu-SO4 mineral stable


phases can be found in the solution mining operation at the San Manuel Mine, and
is unlikely that Cu-SO4 minerals would precipitate during solution recycling.

128

Cupricferric(Eh-S)
Cuprousferric(Eh-S)

-*-

Cuprlcferric (Eh-w)

-B-Cuprousferrlc(Eh-w)

2.1

2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9

pH

Figure 5.4.8 Saturation index (SI) versus pH for copper-iron mineral series from
WATEQ4F by using Eh-w and Eh-s.

129

Cu-PHRQPITZ
Antlerile

-+Brochantite

---

chalcanthite

-B-Langite

-20

2.1

2.2

2.3

2.4

2.5
pH

2.6

2.7

2.8

2.9

Figure 5.4.9 Saturation index (SI) versus pH for copper bearing minerals.

130

CHAPTER 6
CONCLUSION

The chemical analysis for the chrysocolla specimen used in the


experimental study indicates that the chrysocolla is relatively pure, containing 30.5%
copper.
Batch dissolution experiments demonstrate that the leaching of chrysocolla

by sulfuric acid is a stoichiometric reaction, consuming approximately two moles of

proton to produce one mole of copper. pH has a slight effect on the ratio of proton
to copper in a dilute solution, having lower ratio at higher pH value; while the ionic
strength has no obvious effects.

Batch dissolution experiments show that the copper dissolution rate is


strongly pH dependent, so higher dissolution rates occur at lower pH value.

Batch dissolution experiments illustrate that higher ionic strength has a


negative effect on the copper dissolution rate, but the effect will vanish when the
ionic strength of the supporting electrolyte falls below 3 m.
Batch dissolution experiments demonstrate that the chrysocolla dissolution

process follows the diffusion controlled shrinking core model. The dissolution rate
constants can be determined from the Crank, Ginstling and Brounshtein's diffusion

equation. The differences for rate constants are significant for different pH values,
but only slightly different for ionic strength.

131

The morphology of chrysocolla specimens before and after acid dissolution

at different ionic strength was examined by the scanning electronic microscope.


Element composition was qualitatively identified by the X-ray scanning

Results

showed that the secondaly mineral precipitation only occurred in one case at 2M
Na2SO4 + O.5M MgSO4 as supporting electrolyte at the laboratory conditions but

would not be serious enough to reduce the dissolution rate.


Field aqueous samples collected from the San Manuel Copper Mine have

been analyzed for chemical composition. The major ions are Al, Fe, Cu, Si02, and

so4.

The two data source

files of PHRQPITZ, P1TZER.DATA and

PHRQPITZ.DATk were expanded to include Al and Cu, and calibrated with


Occlehaw's experimental solubility data for minerals alungen and chalcanthite. The

required ion interaction parameters and mineral thermodynamic data were taken
from Reardon's and Pitzer's studies.
Geochemical computer models SOLMINEQ, WATEQ4F, PHREEQE, and

PHRQPITZ were used to calculate the ionic strength, spciation, and saturation
indices for experimental and field aqueous samples. Results can be used to directly

predict the effective concentration and mineral saturation states for given samples

with conditions of pH, Eh, temperature, and solution composition. Geochemical


codes are useful tools for interpreting water-rock interaction.
Modeling results for 11 experiments are confirmed with morphological

132

analysis using the scanning electron microscope. No stable phase of copper-bearing


minerals can be found. Secondary mineral precipitation is not an influence factor at
dissolution study.

Modeling results for field samples demonstrate the potential for


formation of surface coatings caused by the secondary mineral precipitation for
present solution mining operation at the San Maunel Copper Mine. The potential
secondary minerals include: Si-bearing minerals, cristobalite (a), cristobalite (fi),
quartz, silica (amorphous), silica (gel), and chalcedony; Al-bearing minerals, jurbanite

and alunite; Fe-bearing minerals, geothite, hematite, maghemite, Fe(OH)23C13, and

megnetite; jarosite mineral series (K, ss, Na, and H); nontronite mineral series (Na,

K, H, Ca, and Mg); and cupric ferric and cuprous ferric.


The effects of pH, temperature, and Eh on mineral saturation state during

leaching process are observed from the geochemical computer model calculation for
field aqueous samples. All except Si-bearing minerals show a tendency for increasing

saturation with increasing pH. Generally, the thermodynamic equilibrium constants


of minerals are a function of temperature. Most minerals have smaller log K values

at higher temperature; therefore, minerals tend to reach equilibrium or saturate at

higher temperature rather than at lower temperature. Redox potential (Eh) is a


control factor in predicting the redox mineral stable phase. Correct Eh measurement

or calculation is necessary to determine the speciation of redox pairs. Under certain

conditions, properly adjusting pH, Eh, and temperature during solution mining

133

operation will help obtain optimum ionic strength, and solution compositions and
concentrations to minimize secondary mineral precipitation.

13). Proper geochemical computer model application is important, because


each code has a specific application for the appropriate ionic strength range, implicit

theory, or specific thermodynamic data base. Two different models, ion association

model and ion interaction model generate different activity coefficients at


concentrated solution, while activity coefficients

calculated from the different

equations within the ion association model are of no major difference. Results show

different saturation tendencies when different thermodynamic data are used.


Sensitivity check for thermodynamic data base should be carried out if reliable results

are desired.

134

APPENDICES

135

APPENDIX A

pH-Stat Computer Program, Written in Basic

136
5

10
15
20
25
30
35

40
50
55
60
70
80
90
100
110
120
130

250
260
270
280
290
300
305

310
320
330
331
332
333
337
338
340
345
349
350
351
352
353
354
355
356
357
360
370
380
381
382
383
390

400
410
420
430
470
475
480

KEY OFF: CLS


PRINT "THIS IS A PH STAT PROGRAM."

PRINT"
PRINT PLEASE TURI4 ON THE 665 DOSIMAT AND THE PH METER AND PRESS
PRINT 'ANY KEY TO CONTINUE."

PRINT""
PRINT""
W$=INKEY$: IF W$=" "THBN4O
OPEN "COM1:9600,E,7,1,LF" AS #3
OUT (&H3FC),O
EOT$ = CIIRS(13) + CHR$(10) + CHR$(4)
DIM S$(20),P(14)
S$ = "REM ON" + EOT$
GOSUB 470
REM CHOOSE DISPENSE REPETITIVE MODE

P=2

IFP=2THENS$="DIR"EOT$
GOSUB 470
INPUT "ENTER EXPERIMENT ID =
INPUT "WHAT PH DO YOU WANT TO KEEP YOUR SAMPLE? ",M
INPUT "PLUS OR MINUS HOW MANY PH UNITS? ",G
INPUT "TITRATING WITH ACID(1) OR BASE(2)? ",.J
INPUT "ENTER THE DISPENSING VOLUME FROM .001 TO 5.00 ML

S$="VDS" i-Z$+EOT$
N=VAL(Z$)
GOSUB 470
INPUT "WHAT INTERVAL DO YOU WANT TO CHECK PH IN HH:MM:SS 7 ",I$
HR = VAL (MID$(I$,3,2))
MIN = VAL (MIDS(I$,4,2))
SECS = VAL (MID$(I$,7,2))
X = MIN + HR6O + SECS/60
INPUT "FILENAME TO STORE DATA (IN QUOTES, UP TO 8 CFIARACIERS)? ",FILE$
INPUT "ENTER TITRANT AND CONCENTRATION => ",C$
INPUT "ENTER A NUMBER 1 TO START THE PUSTAT OR ANY OTHER TO EXIT => ",K

IFK<>1THEN5SO
OPEN FILE$ FOR OUTI'UT AS #2
PRINT #2, "EXPERIMENT ID =
PRINT #2, "TITRANT AND CONCENTRATION =
PRINT #2, "PH SET POINT =
PRINT #2, "PH TOLERANCE =
PRINT #2, "ELAPSED TIME
PH
VOLUME ADDED"
PRINT #2, "-----------------.--------------------CLOSE #2

V=0
T=0
TIMES ="00:00:00"

GOSUB 700

DEFSEG=0
K = PEEK(1048) AND 128

IF K 128 THEN GOSUB 900


IF J=1 AND M<(A-G) THEN S$="G" + CHR$(4) :GOSUB 470
IF J=2 AND M>(A+G) THEN S$="G" + CHRS(4) :GOSUB 470
V$=TIME$
IFV$ >= I$THENTT+X: GOTO 370
GOTO 410

V=V+N
C%=1
OUT (&H3FC),2

137
490
500
510
520
530
540
550
560
580
590

D% = INP(&H3FE)

600
610
700
710
720
750
760

PRINT 'YOU HAVE EXITED THE PROGRAM"


END
ON ERROR GOTO 800
OPEN "COM2:600,s,7,2,CS,DS,CD" FOR INPUT AS #1
INPUT #1,PH$
A=VAL(PH$)
PRINT T,A,V
OPEN FILE$ FOR APPEND AS #2
PRINT #2," ';T,"
";A," ";V
CLOSE #2
CLOSE #1

761

762
763
770
780
790
800
810
820
830
840
900
910
920
930
940
950
960
970
990
1000
1010
1020
1030
1040
1050
1060
1070
1080
1090
1150
1160
1170
1180
1190
1200
1210
1220

S%=D%AND16
D% =ASC(MID$(S$,C%,1))
OUT (&H3FC),0
OUT (&H3F8),D%

C%=C%1
IF MID $(S$,C%,1) = CHRS(4) THEN RETURN
GOTO 480
S$="REM OFF" + EOT$
GOSUB 470

ONERRORGOTO700
RETURN
REM ERROR OCCURRED
CLOSE #1
OPEN "COM2:300,N,8,1,CS,DS,CD" AS #1
RESUME 720
RETURN
POKE 1047, PEEK (1047) AND 127
PRINT 'PARAMETER CHANGE MENU":PRINT:PRINT
PRINT "1) SET POJNT pH"
PRINT "2) SET PRINT INTERVAL"
PRINT "3) SET TOLERANCE"
PRINT "4) RESET ALIQUOT SIZE"
PRINT "5) RETURN TO PROGRAM"
PRINT "6) EXIT"

PRINT: INPUT "ENTER THE NUMBER OF YOUR CHOICE = > ";K9


CLS
IF 1(9 = 5 THEN RETURN
ON 1(9 GOSUB 1050,1070,1210,1150,1040,1200

GOTO 900
RETURN
INPUT "ENTER pH SET POINT = >
RETURN
INPUT "ENTER PRINT INTERVAL IN MINUTES > ';X
INPUT "ENTER PRINT INTERVAL IN HH:MM:SS = > ';I$
RETURN
INPUT "ENTER ALIQUOT SIZE FOR TITRANT =
S$="VDS" + Z$ + EOT$
N=VAL(Z$)
GOSUB 475
REFURN
GOTO 580
INPUT "ENTER TOLERANCE IN pH UNITS = > ';G
RETURN

138

APPENDIX B

The Update Data File For Geochemical Computer Model PHRQPITZ:


PITZER.DATA

139
SPECIES 29
1
4
5 6

45 76 85 86

789

10 11 12 13 14 15 16 18 21 22 31 34 35 40 41 42 43 44

BO

NM
MG+2
CA+2
MGOH+

CLCLCL-

ClCL-

ClLI+
SR+2

FE+2
MN+2

BA+2
CU+2
CABC)4+
MGBO4+

NM
K+
H+

MG+2
CA+2

LI+
SR+2
BA+2

MG+2
CA+2
H+

LI+
SR+2
FE+2
MN+2
CU+2

AL+3
NA+

M0+2
CA+2
H+

FE+2
NA+

CA+2

LI-'BA+2
NA+
MG+2

CA+2
SR+2
NM-

CLCLCLCLCLCLCLCLBRBRBRBRBRBRBRBR-

SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
HSO4HSO4HSO4HSO4HSO4HSO4OHOHOHOHOHHCO3HCO3HCO3HCO3HCO3-

CO3-2
CO3-2

0.0765
0.04835
0.35235
0.3159

-0i

0.17Th
0.1494
0.28575
0.335925
0.327225

0.12
0.2817
0.12
0.16
00973
0.0569
0.1960

04327
0.3816
0.1748
0.331125
0.31455
0.01958
0.04995
0.221

-777.03

-4.4706

0.008946
5.794E-4
-1.943E-4
-1.725E-4

-3.3158E-6

-3.081E-4
-1.685E-4
0717E-3

06405E-3

7692E-4
739E-4
-2049E-4
-5625E-5
-52275E-4
-1819E-4
-032775E-3
-033825E-3
2367E-3
L44E-3
-069E-3

0.2
0.0298
0.136275
0.220
0.2568
0.2065
0.2304

0.5055E-3

-29E-3

0854
00454
-0.0003

04746
0.2145
0.2065
0.4273
0.0864

7.00E-4

01298
-01747
0015
017175
0.0277

00296
0.329
0.4
0.12
0M399
0.1488
-0.0427

NA+

B(OH)4-

NA+
NA+

B30304- -0.056

1.00E-3
0.996E-3

1.79E-3
1.788E-3

B4050-- -011

0.035
B(OH)4B303Ui4- -0.13
K-'-

B4050-- -0.022

B1

NA+
K-'-

MG+2

CLCLCL-

0.2664
0.2122
1.6815

61608E-5
10.71E-4
3.6525E-3

1.0715E-6

140
CA+2

MGOH+
H+

LI+
SR+2
FE+2
14N+2

BA+2
CU+2
NA+
H+

MG+2
CA+2
LI+
SR+2
BA+2
NA+
110+2

CA+2
LI+
SR+2
FE+2
MN+2
CU+2
AL+3
NA+
110+2

CA+2
FE+2
NA+

CLCLCLCLCLCL-

CLCLCLBRBRBRBRBRBRBRBR-

SO4-2
SO4-2
SO4-2
s04-2
SO4-2
s04-2
SO4-2
SO4-2
SO4-2
SO4-2
HSO4HSO4HSO4HSO4HSO4HSO4OHOH-

CA+2
LI+
BA+2
NA+
MG+2
CA+2
NA+
NA+
NA+
NA+
K+

OHOHOH-

1.614
1.658
0.2945
0.3074
1.66725
1.53225
1.55025
1.49625
1.4677
0.2791
0.2212
0.3564
1.753
1.613
0.2547
1.7115
1.56975

3.9E-3
1.419E-4

5366E-4
28425E-3
32325E-3
1079E-4
17.40E-4
4.467E-4
3.8625E-3
6.0375E-3

6636E-4
6.5325E-3
6.78E-3
5.6325E-3

L113

66975E-3
L53E-2

0.7793
3.343
3.1973
1.2705
2.58
3.063
2.9511
2.527
18.53
0.398
0.1735

5.46E-2

L41E-3
27.OE-3

L729
2.53
0.5556
3.48
0.253
0.32
-0.2303
0.14

L34E-4

L2

HCO3HCO3HCO3HCO3CO3-2
CO3-2
B(OH)4-

0.0411
-0.013
0.6072

CL-

0.0023
-37.23
-54.24
-41.5
-42.0
-40.0
-48.33
-500.0

1.1OE-3

L1O4E-3

2977
205E-3
2051E-3

1.389
1.43
0.089
B30304- -0.910
84050-- -0.40
B(OH)40.14

B2

Cu+2
MG+2

CM?
SR+2
FE+2

MN+2
CU+2
AL+3
CA+2
CO
NA+
MG+2
CA+2
LI+

SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
OHCLCLCLCLCLCL-

-0.253
-0.516

-042

-572
0.00127
-0.00084
0.00519
-0.00034
0.0008
0.00359

33.317

0.09421

-4.655E-5
-5.095E-5

-L64933E-4

6213E-5
-4.520E-5

141
SR+2
FE+2
MN+2
BA+2
CU+2
NA+
11+

MG+2
CA+2
LI+
SR+2
BA+2
NMMG+2
LI+
SR+2
FE+2

MN+2
CU+2

AL+3
NA+

NA+
NA+

CLCLCLCLCLBRBRBRBRBRBRBRBRSO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
OHOHHCO3CO3-2
CO3-2
B(OH)4-

-0.00130461
-0.00860725
-0.0204972
-0.0193782
-0.01395
0.00116
-0.00180
0.00827
0.00312
-0.00257
0.0053
0.00122506
-0.0159576
0.00497
0.025
0.0438
-0.00399338
0.019
0.0209
0.01636
0.0044
-0.0911
0.0044
0.0041
-0.008
0.0044
-0.0015
0.0114

NA+
NA+
NA+
NA+

-0.012
0.07
0.07
0.036
0.032
0.005
0.007

THETA
MG+2
CA+2
CA+2
K+

CA+2

140+2

H+

140+2

CA+2

SO4-2
HSO4OHHCO3-

CO3-2
B(OH)4-

B30304B4050-OHOHHCO3CO3-2
B(OH)4-

B303Q4B4050-CO3-2
CO3-2

B30304B405Cj-LAMDA
NA+
K+
140+2

CA+2
CL-

SO4-2
HSO4-

CLCLCLCLCLCLCLCLBRSO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
OHHCO3HCO3HCO3-

H2CO3
H2CO3
H2CO3
H2CO3
H2CO3
H2CO3
H2CO3

0.1

0.092
0.02
-0.006
-0.05
0.03
-0.02
-0.065
0.12
0.074
-0.065
-0.013

001
0.02
-0.012
0.10
0.12
0_i

-0.04
-0.10
-0.087
0.1

0.051

0183
0.183
-0.005
0.097
-0.003

-1.53796E-4
-9.30E-5
-7.004E-5

-5685E-5

-2813E-5

-4.87904E-4
0.523E-3
-2.33345E-4
3.OE-3

-1894E-5

142
NA+
CL-

SO4-2

B30304-

B(OH)3
B(OH)3
B(OH)3
B(OH)3
B(OH)3

-0.097
-0.14
0.091

0.018
-0.20

ZETA
H+

NA+

CLSO4-2

B(OH)3
B(OH)3

-0.0102
0.046

CL-

-0.0018
-0.0022
-0.010
-0.003
0.003
-0.007
-0.055
-0.012
-0.015
-0.004
-0.012
-0.0129

PS I

NA+
NA+
NA+
NA+
NA+
NA+
WA+
NA+
NA+
NA+
NA+
NA+

K+

BR-

H+

SO4-2
HCO3CO3-2
CLSO4-2
CLSO4-2
CL-

11+

BR-

H+

HSO4-

CA+2
MG+2
MG+2

CLCL-

CA+2
CA+2
MG+2
MG+2

H+
H+

SO4-2

-0.022
-0.048

CL-

-0.011

BR-

-0.021

SO4-2

0.197
-0.0265
-0.012
0.024
-0.015
0.028

HSO4-

CA+2
CA+2
CA+2
MG+2
MG+2
MG+2
AL+3
CLCLCLCLCLCLCL-

CLCLCLCLCL-

CLCL-

CL-

CLSO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
SO4-2
OHOHOHOHHCO3-

MG+2
MG+2

MGOH+
H+
H+

CLSO4-2
CLCLCLHSO4-

CU+2

SO4-2

BR-

K+
NA+
CA+2

SO4-2
SO4-2
SO4-2
HSO4HSO4OHOHOHHCO3HCO3-

CO3-2
CO3-2
B(OH)4-

B30304B4050-HSO4HSO4HSO4OHOHHCO3HCO3-

CO3-2
CO3-2
CO3-2
CO3-2
BRBR-

CO3-2

-0025

140+2
IIA+
11+

NA+
K+
CA+2
NA+
MG+2
NA+
K+
NA+
NA+
NA+
NA+
K+
MG+2
NA+
NA+
140+2

NMK+
NMK+
NA+
K+
NA+

-0.011

-0.0178
0.035
0.0000
0.0014
-0.018
-0.004
-0.006
0.013
-0.006
-0.006
-0.025
-0.015
-0.096
0.0085
0.004
-0.0073
-0.024
0.026

-00094
-0.0677
-0.0425
-0.009
-0.050
-0.005
-0.161
-0.005
-0.009
-0.017
-0.01
-0.018
-0.014
0.002

143
HCO3-

CO3-2

K+

0.012

144

APPENDIX C

The Update Data File For Geochemical Computer Model PHRQPITZ:


PHRQPITZ.DATA

145
ELEMENTS
CA
MG
NA
K

40.08
24.305
22.9898
39.0983
55.847
54.9380
63.54

4
5

6
7

8
9

FE

MN
Cu
BA
SR
AL
CL

10

12
13
14
15
16

S
B
LI

18
21

BR

22

FE+2
MN+2
CU+2
BA+2
SR+2
AL+3

137.33
87.62
26.982
35.453
44.0098
96.06
10.81
6.941
79.904

11

CA+2
MG+2
NA+

CLCO2

SO4-2
B

LI+
BR-

SPECIES

0.0
1

100 1.0
0.0

0.0

9.0

0.0

100 -1.0
0.0

0.0

0.0

0.0

100 0.0
0.0

0.0

0.0

0.0

101 2.0
0.0

0.0

6.0

5.0

0.165

0.0

101 2.0
0.0

0.0

8.0

5.5

0.20

0.0

101 1.0
0.0

0.0

4.0

4.0

0.075

0.0

101 1.0
0.0

0.0

3.0

3.5

0.015

0.0

100 20

2.0

6.0

0.0

100 2.0
0.0

2.0

6.0

0.0

100 2.0
0.0

0.0

5.0

0.0

100 2.0
0.0

0.0

5.0

0.0

101 2.0

0.0

5.0

1.0

E-

0.0
2 1.0
3

H20
0.0

3 1.0
4

CA+2
0.0
4 1.0
5

MG+2
0.0
5 1.0
6
NA+
0.0
6 1.0

7
K+
0.0

7 1.0
8
FE+2
0.0
8 1.0

0.0

MN+2
0.0
9 1.0
10

CU+2
0.0
10 1.0
11

BA+2
0.0
11 1.0
12

SR+2

5.26

0.121 0.0

146
0.0
12 1.0
13

AL+3
0.0
13 1.0
14

0.0

101 3.0
0.0

101 -1.0

CL-

0.0
14 1.0

0.0

0.0

5.0

0.0

3.0

3.5

0.015

0.0

0.0

15

CO3-2
0.0
15 1.0
16

s04-2
0.0
16 1.0
18

B(OH)3
0.0
18 1.0

101 -2.0
0.0

4.0

4.5

5.4

0.0

2.0

101 -2.0
0.0

6.0

4.0

5.0

-0.04

0.0

1000.
0.0

0.

0.0

0.0
0.0

0.0
0.0

0.0
0.0

0.0
0.0

21

L1+
0.0
21 1.0
22
BR0.0
22 1.0

100 1.0
0.0

0.0

6.0

0.0

100 -1.0
0.0

0.0

3.0

0.0

200 -1.0
13.345
1
-1.0

0.0

3.5

1.0

211 -1.0

4.0
107.8975

31

OH-

-13.998
3 1.0
34
HCO310.3393
15 1.0
35
H2CO3
16.6767
15 1.0
40
HSO41.979
16 1.0

-3.561
1
1.0

310 0.0
-5.738
1 2.0
210 -1.0
4.91
1
1.0

4.0
464.1925
3 -1.0
6.0
-5.3585

4.5

5.4

0.03252849

-5151.79

0.0
0.09344813

-26986.16

0.0
1.0
-38.92561
563713.9

0.0
-165.75951
2248628.9

0.0

0.0183412

0.0

5572461

41

B(OH)4-9.239

300-1.
0.

31.

181.

0.

0.0
1-1.

0.

0.0

0.0
0.0

0.0
0.0

1.

0.0
0.0

0.0
0.0

1.

0.0
0.0

0.0
0.0

2.

0.0
0.0

0.0
0.0

1.

0.0
0.0

0.0
0.0

1.

0.0

42

B30304-7.528

300-1.
0.

3-2.

183.

0.

0.0
1-1.

0.

0.0

0.0

43

B4050--16.134

300-2.
0.

3-3.

184.
44

CABQ4+
-7.589

4001.
0.

181.
45

41.

MGBQ4+
-7.840

0.

0.0
1-2.
0.

0.0
31.

4001.
0.

0.

0.0

0.0

0.0

1-1.
0.

0.0

0.

0.0

0.

0.0

0.0

147
181.
76

CACO3
3.151
4 1.0

51.

31.

1-1.

210 0.0
4.0
3.547
-1228.806
15 1.0

0.0
-0.299440

2.0

35512.75

485.818

85

MGOH+
-11.809
5 1.0

300 1.0
15.419

3 1.0

0.0
1

0.0

1.0

-1.0

86

MGCO3
2.928
5 1.0
LOOK HIM
At(OU)3a

210 0.0
4.0
2.535
-32.225
15 1.0

2.0

1093.486

0.00

10.8
3.000
1
-3.000
ALUNOGEN
6.00
-6.204
13
3
16
3.000
17.000
ANHYDRIT
2
6.00
-4.362
4
1.000
16
1.000
422.950
0.0
-18431.
-147.708
1.
ANTLERIT
0.00
9.92
10
3.000 16
1.000
3
3.000
ARAGONIT
2
4.00
-8.220
15 1.0
4 1.0
-171.8607
-.077993
2903.293
71.595
ARCANITE
2
6.00
-1.776
7
2.000 16
1.000
2.823
0.0
-1371.2
BASALUI4I
4
0.00
22.7
13
4.000 16
1.000
3
10.000
BISCF1OFI
3
0.00
4.455
5
1.000 14
3
2.000
6.000
3.524
0.0
277.6
BLOEDITE
4
12.00
-2.347
6
2.000
5
1.000 16
2.000
BOEHMITE
3
0.00
8.584
13
1.000
3
2.000
1
-3.000
BRUCITE
2
0.00
-10.884.85
5
1.000 31
2.000
BROCHANT
4
0.00
15.34
10
4.000 16
1.000
3
6.000
BURKEITE
3
16.00
-.772
6
6.000 15
1.000 16
2.000
CALCITE
2
0.00
-8.406
15 1.0
4 1.0
-171.8329
-.077993
2839.319
71.595
CARNALLI
4
0.00
4.330
7
1.000
5
1.000 14
3.000
CHALCANT
3
-2.621
0.00
10 1.0
16 1.0
3 5.0
DOLOMITE
3
8.00 -17.083 -9.436
4 10
5 1.0
15 2.0
EPSOMITE
3
6.00
-1.881
5
1.000 16
1.000
3
7.000
1.718
0.0
-1073.
GAYLUSSI
4
8.00
-9.421
4
1.000
6
2.000 15
2.000
GIBBSITE
3
0.00
8.11
13
1.000
3
3.000
1
-3.000
GLASERIT
3
24.00
-3.803
6
1.000
7
3.000 16
2.000
GLAUBERI
3
12.00
-5.245
4
6
2.000
1.000 16
2.000
13

3
1.000
3
2.000

0.0
0.0

12.72433

0
1

-4.000

1
1

-10.000

1
1

4.000

0
0
0

-6.000

0
1

6.000
0
0

5.000
0
0
0

148
GYPSUM
3
6.00
-4.581
4 1.0
16 1.0
3 2.0
90.318
0.0
-4213.
-32.641
HALITE
2
0.00
1.570
6
1.000
14
1.000
-713.4616
- .1201241
37302.21
262.4583
HEXAHYDR
3
6.00
-1.635
5
1.000
16
1.000
3
6.000
-62.666
0.0
1828.
22.187
JURBANIT
4
0.00
-3.23
13
1.000 16
1.000
3
1.000
KAINITE
5
6.00
-0.193
7
1.000
5
1.000 14
1.000
KALICINI
3
4.00 -10.058
7
1.000
1
1.000 15
1.000
KIESERIT
3
6.00
-0.123
5
1.000 16
1.000
3
1.000
LANGITE
4
0.00
16.79
10
4.000 16
1.000
3
7.000
LABILE S
4
18.00
-5.672
6
4.000
4
1.000 16
3.000
LEONHARD
3
6.00
-0.887
5
1.000 16
1.000
3
4.000
LEONITE
4
12.00
-3.979
7
2.000
5
1.000 16
2.000
MAGNESIT
2
4.00
-7.834
-6.169
5
1.000 15
1.000
MIRABIL
3
6.
-1.214
62.

PIRSSONI

6
POLYHALI
7

PORTLAND
4
SCFIOENIT

2.000
1.000

1.000
0.0

14

4
4

3.000

2.000
4

16
0.

184.

B-ACID,S

31

2.000

7
TRONA

BORAX

2.000

16

2.000

7
SYLVITE
7
3.984
SYNGENIT

THENARDI

1577.756

-2106915.
1

-1.000

1.000

16

3.000

2.000

0
0

-6.000

2.000
0
0

4.000

3
0

0
0

1.000
0

10.000
0

3.000
-4.776

669365.0

-40.45154

1.000
3
8.00
-9.234
1.000 15
24.00 -13.744
1.000
4
0.00
-5.190
2.000
12.00
-4.328
1.000 16
0.00
.900
1.000
-919.55
12.00
-7.448
1.000 16
8.00 -11.384
1.000
15
4.00
-0.18
1.000
12.464
0.

62.

0.

35.

5.000
0

2.000

2.000

2.000

16

4.000
0
0

2.000

6.000

3
1

2.000

1.000
0

2.000
-0.57

2.000
0

0
1-2.

0.

0.

-0.030

0.

0.

0.

4.671

0.

0.

181.

KB5084W

7.000

0.01985076 -6919.53
3
6.00
-1.285
1.000

310.

161.

-3862.234
-1.19856
93713.54
MISENITE
3
42.00 -10.806
7
8.000
1
6.000 16
NAHCOLIT
3
4.00 -10.742
6
1
1.000
1.000
15
MATRON
3
4.00
-0.825
6
2.000 15
1.000
3
NESQUEHO
3
4.00
-5.167
1.000
5
15
1.000
3
PCO2
1
4.0
-1.468
35 1.0
108.3865
PENTAHYD

149
185.
4

0.

0.

184.
181.

HCL
U2SO4
HBR

END

10.840
62.

2
2
2
2
2
2
2
2

4
4
4

1.0
1.0
1.0

6
6
6
6
6

1.0
1.0
1.0
1.0
1.0
1.0

2
2
2
2
2

2
2
2
2
2
2
2

5
5

5
5

7
7
7
7
7

2.0

14
16
15
31
14
16
15
31
14
16

1.0

34

2.0

15
31

1.0
1.0

0.

3-2.

0.

181.

NACL
NA2SO4
NAHCO3
NA2CO3
NAOH
KCL
K2SO4
KHCO3
K2CO3
KOH

5.895
61.

TEEPLEIT

0.

33.

0.

185.

0.

3-1.

9.568
61.

NAB5085W

MG(0I1)2

13.906
72.

NABO24W

MEAN GAM
CACL2
CASO4
CACO3
CA(OH)2
MGCL2
MGSO4
MGCO3

3-3.

71.

K2B4074W

2.0

14
16

1.0

34

2.0

15

1.0
1.0

31
14

2.0

16

1.0

22

2.0
1.0
1.0

2.0
2.0
1.0
1.0

2.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0

0.
141.

1-1.
0
1-2.
0
1-1.
0
1-1.
0
31.

0.

0.
0.

1-1.

150

APPENDIX D
The Analytical Results from Acid Mine Drainage

in West Squaw Creek and Its Tributaries (Filipek et al., 1987)

6.1

3
3.25

31*
32*

6.15
6.85
5.25
2.45
2.55
3.05
6.15
6.3
3.55
6.2
2.75
4.2
3.25
3.85
2.4
3.6
5.65
5.65
5.65
3.65
2.35
2.8
5.55
2.7
5.55
2.7
5.25
2.75
2.95

pH

30

28*
29*

27

26*

25

24*

23

20*
21*
22*

17
18
19

13*
14*
15*
16*

12

10
11*

6
7
8
9

4*
5*

samp.
site

150
675
660

850
60
950
750

1500
120

45

1800

2000
400
1800
300
4500
500
90
70
33
450
5100

110

25
55

105

35
3000
850
250

100

1.7

5.2

12

0.9
<0.05

1.6

9
<0.05
2.2
<0.05

13.8
6.8
0.26

<0.05
<0.05
<0.05

6.6
0.13
500
0.47

11

0.009
32
12
0.019

41

72
0.006

120
0.24

9.2
0.009
<0.05
0.022
2.2
930

1600

27

94

<0.05
0.15
0.005
3.9
<0.05

<0.03
<0.03
<0.03
<0.03
<0.05
0.18
<0.05

100
10.3

730
110
0.17
0.007

0.01

<0.03
<0.03
<0.03

0.008
0.06

mg/I

100

Fe2+

tot.Fe

uS/cm mg/I

cond.

<0.01
0.92
<0.01
0.55
<0.01
0.55
0.48
<0.01
0.45
0.51

1.3

0.68
<0.01
<0.01
<0.01
0.63
4.4

1.9

0.62

1.2

2.3

0.01

0.25
0.24
0.04
<0.01
0.04

1.3

<0.01
<0.01
<0.01

mg/I

Mn

5.9
0.003
5.4

7.1
0.01
7.1

0.004

12.3

0.016

102
17.4

4.8
0.025
0.15
0.005
4

156

0.014
80
13.4
3.8
0.44
0.002
0.33
0.17
17.5
5.4
23
3.8

0.01

<0.012

mg/I

Zn

<0.01
4.8
4.3

5.1

<0.01
11.5
<0.01
6.4
<0.01
6.5

105
16.4

3.2
<0.01
0.05
<0.01
2.3

190

13.2

8.7
2.8

<0.01
<0.01
<0.1
137
22.4
5.2
0.46
<0.01
0.12
0.08

mg/I

Cu

0.015
9.2
8.8

10

0.009
8.9
0.029
9

15

<0.05

19

8.7
28
6.5
89
7.6
0.033
<0.05
0.022
6.8
90

51

0.22
0.05

,0.05

2.3
5.2
0.05

14

0.018
0.034
0.021

mg/I

AJ

4.9
4.9

7.1

4.6
4.6
4.7
1.9
4.6
4.8

2.1

13

13

27

21
5.1
16
10
14
5
13
12

8.5

16

0.24
0.5
0.29
0.08
0.28
0.08
0.19
0.15
0.2
0.33
0.04
0.23
0.29

3.6
7.6
4.2

7.4

2.9
7.4

8.2
7.8

1.6

8.1

1.2
0.88
7.5
6.6
16
1.3
12
2.3

32
7.6
7.9
8.3

11

33

1.9
1.2
1.9
2.1

2.2

1.7

0.7
0.82

1.9

1.2

5.2

SO4

187

237
178
54
204

11

290

13

634
8.2
380

140
3500

2.6

11

11

183

5100

140

602

150

3.8
20
26
506

112
21

371

4.7
2260

21

22

mg/I mg/I

Mg

7.2

16
8.5
16

16
23

11

7.2
6.3
15

14
18
6
12
18
17
18

mg/I

Ca

0.11

0.36
0.42
0.22
0.58
0.25
0.07
0.12

0.41

0.19
0.14

0.1

0.36
0.14

0.1

0.12
0.12
0.03
0.28

mg/I

3.1

3.6
3.8
3.6

13

3.3
2.2
3.9
3
3.4

2.8
3.3
3.3
5.5
3.7
3.7
3.4
2.5

mg/I

Na

<1
<1

<1
<1
<1
<1
47

<1
18

<1
<1

10

8.5

<1
<1
<1
<1
9
2

<1
23
<1

17

<1
<1
<1
34

30
45

0.4
2.2

1.7

0.4
0.6

0.3

1.2

2.3
0.5

0.4
0.6
0.6
0.7
0.5

1.1

2.7
0.3
4.4
0.4
0.7

1.1

1.2

0.4
0.4
0.4

1.7

0.4
0.5
3.7
2.7
0.7
0.4

HCO3 Cl
mg/I mg/I

Table 4.7.1-1 The Concentrations of Dissolved Elements in Waters from West Squaw Creek and Its Tributaries (Filipek et aI, 1987)
F

0.14
0.7

0.21
0.21

0.2
0.04
0.2
1.3
0.3
0.07
0.59
0.08
0.25
0.03
0.19

0.1

0.21
0.03
0.06
0.06
1.2
0.36
1.5
0.15
0,78
0.19

0.37

0.11

0.03
0.47

0.11

0.09

mg/I

Si02

28
24
29
29

28

12

27

18
18
18
24
64
30
12
42
19

21

54
28
42
25
95

17

7.5

13

18

26

21

37

18

17

16

mg/I

<1
<1
<1
<1
<1
<1

12

<1

<1
<1
<1
160

<1
<1
<1
<1
<1
450
3.9

1.7

<1
<1
<1
<1
<1

ug/I

As

152

LIST OF REFERENCE
Auck, Y.T., and M.E. Wadsworth, Physical and chemical factors in copper dump
leaching, International Symposium on Hydrometallurgy, Chicago, IL, D.J.I. Evans and
KS. Shoemaker edN, 1973.

Averill, W.A., The chemistry and modeling aspects of copper dump leaching, Ph.D.
Dissertation, Univ. of Utah, 1976.

Ball, J.W., and D.K. Mordstrom, User's manual for WATEQ4F, with revised
thermodynamic data base and test cases for calculating speciation of major, trace,
and redox elements in natural waters, U.S. Geological Survey, Open-File Report 91183, 189 pp., Menlo Park, California, 1991.
Bassett, R.L., Personal contact, 1992.
Bassett, R.L., and J.B. Hiskey, Proposal--Aqueous phase solution chemistry and the
chemical interactions at mineral surfaces during the leaching of copper ores -the
effect of solution recycle, 1989.
Braun, R.L., A.E. Lewis, and M.E. Wadsworth, In-place leaching of primary sulfide
ores: laboratory leaching data and kinetic model, Metallurgical Transactions, 5, 17171726, Aug. 1974.

Chukhnov, F. V. and F.J. Anosou, Nature of chiysocolla, Zapiski Vsesyuz mineral,


Obschestuz, 79, 127-136, 1950; Chem. Abstr., 44, 7724b, 1950.
Crank, J., diffusion with rapid irreversible immobile immobilization, Trans. Faraday
Soc., 53, 1083-1091, 1957.
Dana, J.D., Systems of mineralogy, 3rd Ed., John Wiley and sons, New Yorl 1878.
Davies, C.W., Ion Association. Butterworth, WashingtoniD.C., 1962.

Filipek, L.H., D.K. Nordstrom, and w.H. Ficklin, Interaction of acid mine drainage
with waters and sediments of west Squaw Creek in the west Shasta mining district,
California, Environ. Sci. Technol. 21, 388-396, 1987.
Garrels, R.M., and C.L. Christ, Solutions, minerals, and equilibria, Freeman-Cooper,
San Francisco, California, 45Op, 1965.

153

Ginstling, A M, and B.I. Brounshtein, the diffusion kinetics of reactions in spherical


particles, Zh. Priki. Khim., 23, 1249, 1950. English Trans. in J. App!. Chem. USSR; 23,
1327-38, 1950, CA 46, 9959c.

Habashi, F., Principles of Extrative Metallurgy, Vol.1, General Principles, Gordon


and Breach, New York, 143-164, 1969.

Harvie, C.E., and J.H. Weare, The prediction of mineral solubilities in natural
waters: the Na-K-Mg-Ca-Cl-SO4-H20 system from zero to high concentration at 25C,
Geochimica et Cosmochimica Acta, 44, 98 1-997, 1980.

Harvie, C.E., N. Moller, and J.H. Weare, The predication of mineral solubilities in
natural waters: the Na-K-Mg-Ca-H-Cl-OH-SO4-HCO3-0O3-0O2-H20 system to high
ionic strengths at 25C, Geochimica Cosinochimica Acta, 48, 723-751, 1980.
Helgeson, H.C., Thermodynamics of complex dissociation in aqueous solutions at
elevated temperatures, Journal of Physical Chemistry, 71, 3 121-3136, 1967.

Hiskey, J.B., Renaissance of Copper Solution Mining, Fieldnotes, 5-9, Fall 1986.

Horlick, J.M., W.C. Cooper, and A.H. Clark, Aspects of the mineralogy and
hydrometallurgy of chrysocolla, with special reference to the Cuajone, mt. J. Miner.
Process., 8, 49-59, 1981.
Hsu, P-C, and L.E. Murr, A simple kinetic model for sulfuric acid leaching of copper
from chrysocolla, Met. Trans., 6B, 435-440, 1975.
Khalezov, B. D., Krushkol, O.B., Kakovskii, l.A., Kiseleva, V.1., Dissolution kinetics
of chrysocolla in sulfate solutions, Izvestiya Vysshikh Uchebnykh Zavedenii, Tsvetnaya
Metallurgiya n6, 36-39, 1984.

Kharaka, Y.K., W.D. Gunter, P.K. Aggarwal, E.H. Perkins, and J.D.Debraal,
SOLMINEQ.88: A Computer Program for Geochemical Modeling of Water-Rock
Interactions, U.S. Geological survey, Water Resources Investigations Report 88-422 7,
1988.

Lewis, G.N., and M. Randall, Thennodynamics, McGraw-Hill, New York, 723 p.,
1961.

Magma Copper Company booklet, Magma, 1991.

154

Malouf E.E., Introduction to dump leaching practice, Second Tutorial Symposium


on Extractive Metallurgy, 1972.
Martinez, E., Chrysocolla studied by thermal analysis, thermal gravimetric analysis
and infrared spectrophotometry, Trans. A.I.M.E., 226, 424-427, 1963.

Mena, M. and F.A. Olson, Leaching of chrysocolla with ammonia-ammonium


carbonate solutions, MetaL Trans. B, 16B, 441-448, Sept., 1985.

Newbery, D.W., Geochemical implications of chrysocolla-bearing alluvial gravel,


Econ. Geol., 62, 932-956, 1967.

Nordstrom, D.K., Geochemical thermodynamics, Blackwell Scientific Publications,


264-282, 1986.

Occleshaw, V.J., The equilibrium in the systems aluminum sulphate-copper sulphate-

water and aluminum sulphate-ferrous sulphate-water at 25C, J. Chem. Soc. 127,


2598-2602, 1925.

Parkhurst, D.L., D.C. Therstenson, and L.N. Plummer, PHREEQE - a computer


program for geochemical calculations, U.S. Geological Survey, Water Resources
Investigations, Report 80-96, 1980.

Petersen, E.E., Reaction of porous solids, AIC/ZE J., 3,


443-448, 1957.

Pitzer, K.S., Thermodynamics of electrolytes, 1. Theoretical basis and general


equations, Journal of Physical Chemistry, 77(2), 268-277, 1973.

Pitzer, K.S., Thermodynamics of electrolytes, 5, Effects of higher-order electrostatic


term, Journal of Solution Chemistiy, 4 (3), 249-265, 1975.

Pitzer, K.S., and J.J. Kim, 1974, Thermodynamics of electrolytes, 4, Activity and
osmotic coefficients for mixed electrolytes, Journal of the American Chemical Society,
96, 5701-5707, 1974.
Pitzer, K.S. and Mayorga, G., Thermodynamics of electrolytes, 2, Activity and osmotic
coefficients for strong electrolytes with one or both ions univalent, Journal of Physical
chemistry, 77(19), 2300-2308, 1973.

Pitzer, K.S. and Mayorga, G., Thermodynamics of electrolytes, 3,

155

Activity and osmotic coefficients for 2-2 electrolytes, Journal of Solution Chemistiy,
3(7), 539-546, 1974.
Plummer, L.N., D.L. Parkhust, G.W. Fleming, and S.A. Dunkle, A computer program
incorporating Pitzer's equations for calculation of geochemical reactions in brines,
USGS-WRJ Report 88-4153, pp.310, 1988.

Pohiman, S.L. and F.A. Olson, a kinetic study of acid leaching of chrysocolla using
a weight loss technique, Solutions Mining .symposium, 446-460, 1974.

Pohlman, S.L. and F.A. Olson, Characteristics of chrysocolla pertinent to acid


leaching, in Extractive Metallurgy of Copper, J. C. Yannopoulos and J. C. Agarwal eds.,
A.I.ME, 943-959, 1976.

Prosser, A.P., and A.J. Weight, Physical and chemical properties of natural copper
silicates which resemble chrysocolla, Bull. Inst. Mining Met. 699, 233-58, 1965.
Raghavan, S. and D. W. Fuerstenau, Characterization and pore structure analysis of
a copper ore containing chrysocolla, International Jurnal of Mineral Processing, 4, 381394, 1977.

Raghavan S and S.Y. Gajam, Application of an enlarging pore model for the
ammoniacal leaching of chrysocolla, Hydrometallurgy, 16, 271-281, 1986.

Reardon, E.J., Ion interaction parameters for A1SO4 and application to the
predication of metal sulfate solubility in binary salt systems, The Journal of Physical
Chemistry, 92, 6426-643 1, 1988.

Romam, R.J., B.R. Benner, and G.W. Becker, Diffusion model for heap leaching and
its application to scale-up, Trans. SME/AIME, 256 (2), 103-105, 1974.

Sato, M., Oxidation of sulfide ore bodies, 1. geochemical environments in terms of


Eh and PH, Economic Geology, 55, 928-96 1, 1960.

Shafer, J.I., M.L. White, and C.L. Caenepeel, Application of the shrinking core model
for copper oxide leaching, Mining Eng (Feb.), 165-171, 1979.
Shelly, T.R., Sulfuric acid and ammoniacal leaching of Sar Cheshmeh Oxicidic copper
ore, Trans. 1MM, 84C, 174-176, 1975.
Sohn, H.Y., Kinetics of heterogeneous reaction systems, in Rate Processes of Extractive

156

Metallurgy, Edts: Sohn, H.Y. and M.E. Wadsworth, Plenum Press, 1979, 1-6.

Sullivan, J.D., Dissolution of various oxidized copper minerals, Trans. AIME, 106,
5 15-546, 1933.

Taylor, J.H., and Whelan, P.F., The leaching of cupreous pyrites and the precipitation
of copper at Rio Tinto, Spain, Institution of Mining and Metallurgy Bulletin, 457, 1-36,
1942.

Truesdell, A. H., and B.J. Jones, WATEQ, A computer program for calculating
chemical equilibria of natural waters, National Technical Information Seivice PB220464,

77p, 1973.

Valensi, G., Kinetics of oxidation of metallic spherule and powders, Compt. Rend.,
309-12, 1936, 202, 414-16, 1936, CA 30, 2084, Bull Soc. Chim. France, 5, 668,

202,

1935.

Van Oosterwyck-Gastruche, M.C. and C. Gregoire, Electron microscopy and


diffraction identification of some copper silicates, Inst. MineraL Ass., Pap. Proc. Gen.
Meet., 7th, Tokyo-Kyoto, No.1, 196-205, 1970.

Wadsworth, M. E., Rate processes in hydrometallurgy, Second Tutorial Symposium


on Extractive Metallurgy, Univ. of Utah, 1972.

Wadsworth, M.E., Hydrometallurgical Processes, in Rate Processes of Extractive


Metallurgy, Edts: Sohn, H.Y. and M.E. Wadsworth, Plenum Press, 1979, 188-189.

Windrel, A.N., Elements of mineralogy, Prentice Hall, New York 499-520,


1942.Wadsworth, M.E., Interfacing technologies in solution mining,
Mining Engineering,

29(12),

30-33, 1977.

Вам также может понравиться