Você está na página 1de 145

BISMUTH IN THE SUPERGENE ENVIRONMENT

Timothy D. Murphy BSc (Hons), UWS

This thesis is submitted for the degree of

Doctor of Philosophy

in the University of Western Sydney

Supervisors: Professor Peter A. Williams and Professor Peter Leverett

March 2015
ACKNOWLEDGEMENTS

I sincerely thank my supervisor Professor Peter Williams, for sharing his

extensive knowledge in chemistry and mineralogy which made the project possible. His

guidance, support and encouragement throughout has been greatly appreciated. I would

also like to thank my co-supervisor Professor Peter Leverett for his invaluable

assistance and perspectives on my thesis which have been greatly appreciated. Dr Ric

Wuhrer and the Advanced Materials Characterisation Facility provided assistance with

X-ray diffraction as well as support, guidance and equipment training. Jim Sharp is

thanked for his assistance with fieldwork and sharing his experience in mineralogy and

X-ray diffraction. John Rankin is acknowledged for sharing his knowledge on the

mineralogy of the New England Oregon.

I would like to thank my fellow PhD students Matthew Sciberras, Adam Roper,

Mitchel Nancarrow and Simon Hager for their invaluable help, support and friendship

throughout the project. Finally I would like to thank the people who have helped and

encouraged me outside of the academic environment throughout this endeavour,

especially my Mother, Father and Sister whose help, support, encouragement and love

have helped me become the person I am today, as well as my friends.


STATEMENT OF AUTHENTICATION

This thesis contains work that, to the best of my knowledge and belief, is original except where

due acknowledgment appears in the text. I declare that material in this thesis has not been

submitted in any form for a degree or diploma at any other university or institution of tertiary

education.

..........…………………

Timothy David Murphy

March 2015
TABLE OF CONTENTS

Chapter 1 Introduction 1

1.1 Abstract 2
1.2 Geochemical Exploration 3
1.3 Bismuth as a path finder element 6
1.4 Project Outline 13
1.5 References 13

Chapter 2 Bismoclite, Bismtutite, Cannonite and Riomarinaite 22

2.1 Abstract 23
2.2 Introduction 23
2.2.1 Bismite, bismoclite, bismutite, cannonite and riomarinaite 23

2.2.2 Rare and other bismuth oxide, carbonates, chloride, and sulfate minerals 26

2.3 Experimental 28
2.3.1 Syntheses of cannonite and riomarinaite 28
2.3.2 Solubility Studies 29
2.4 Results 30
2.4.1 Synthesis of cannonite and riomarinaite 30
2.4.2 Solubility Studies 30
2.5 References 40

Chapter 3 Rooseveltite, Preisingerite and Atelestite 45

3.1 Abstract 46
3.2 Introduction 46
3.3 Experimental 50
3.3.1 Syntheses of rooseveltite, tetrarooseveltite, atelestite and preisingerite 50
3.3.2 Solubility Studies 51
3.4 Results 51
3.4.1 Syntheses of rooseveltite, tetrarooseveltite, atelestite and preisingerite 51
3.4.2 Solubility Studies 55
3.5 References 60
Chapter 4 Russellite, Koechlinite and Sardignaite 66

4.1 Abstract 67
4.2 Introduction 67
4.3 Experimental 73
4.3.1 Syntheses of russellite, koechlinite and sardignaite 73
4.3.2 Solubility Studies 74
4.4 Results 75
4.4.1 Syntheses of russellite, koechlinite and sardignaite 75
4.4.2 Solubility Studies 79
4.5 References 84

Chapter 5 Smirnite and Chekhovichite 90

5.1 Abstract 91
5.2.1 Introduction 91
5.3 Experimental 94
5.3.1 Syntheses of smirnite and chekhovichite 94
5.3.2 Solubility Studies 95
5.4 Results 95
5.4.1 Syntheses of smirnite and chekhovichite 95
5.4.2 Solubility Studies 98
5.5 References 102

Chapter 6 Conclusion 105

6.1 Abstract 106


6.2 Introduction 106
6.3 Discussion of Data 108
6.3.1 Chapter 2 Discussion 108
6.3.2 Chapter 3 Discussion 110
6.3.3 Chapter 4 Discussion 112
6.3.4 Chapter 5 Discussion 114
6.4 Reverse Ground Water Modelling 116
6.5 Bismuth Model for the Cobar Region 117
6.6 Bismuth Model for the New England Region 125
6.7 Conclusion 130
6.8 References 131

Appendix A 137
CHAPTER 1

INTRODUCTION

1|Page
1.1 ABSTRACT

Bismuth minerals associated with Mo, W, and Sn, are often found amongst the

highly acidic deposits of the eastern ranges of Australia (Plimer, 1975; Weber et al.,

1978). It is important to gain an understanding of the mobility and dispersion of Bi in

the supergene zone and make an assessment of these areas, as they have been the focus

for geochemical exploration to develop prospects and mining operations. A review of

the literature on bismuth as a pathfinder element, with respect to its ground water and

regolith concentrations, uncovered significant documentation including scientific,

industrial and government reports, the use of various sampling methods, and the use of

assumptions in previous studies due to the information and techniques available at the

time (MacDuff, 1971; 1971a; 1971b; 1972; Siegal, 1974; Roes et al., 1979; Levinson,

1980; Plant et al., 1989; Plant et al., 1991; Fiella; 2010). Furthermore, information on

the Gibbs free energy of formation values was limited to 3 out of the 65 known bismuth

secondary minerals (Clissold, 2007).

A study on a range of secondary bismuth minerals in the supergene zone,

(Rankin et al., 2001, 2002; Sharpe and Williams, 2004) showed that even though

bismuth minerals are considered to be rare, localised areas of Bi concentration are in

fact quite common. Examples of this can be found in certain deposits in eastern

Australia such as the New England Orogen. Therefore, the geochemical modelling

carried out in this thesis has focused on eastern Australia and examines potential

impacts on geochemical exploration where Bi has been used as a pathfinder element.

Furthermore this work can be been applied to exploration sites around the world where

2|Page
Bi is employed as a pathfinder element. To do this, a rigorous investigation including

Bi mineral synthesis, solubility and stabilities was undertaken thus yielding an

assessment of the suitability of bismuth as a pathfinder for future geochemical surveys.

1.2 GEOCHEMICAL EXPLORATION

Due to the nature and growth of manufacturing and consumption worldwide and

the need for mineral resources to meet these demands, the discovery of large-scale

near-surface deposits are in decline and thus the push to find deeper deposits has

inevitably increased (Aspandiar et al., 2008). Therefore, it is necessary to constantly

review and scrutinise new, current and historic exploration techniques to discover

hidden ore bodies that are commercially significant. Exploration geochemistry is used

in virtually every exploration program whereby geochemical prospecting for minerals

include any method of mineral exploration based on systematic measurement of one or

more chemical properties of a natural occurring material e.g. Hawkes and Webb, 1962.

The earliest reports of geochemical research date back to the 1930’s which were

carried-out in the former Soviet Union, and were known as metallometric surveying,

which we know today as geochemical surveys (Hawkes and Webb, 1962). It was not

until the late 1930’s to 1940’s that there was a rise in geochemical research by Western

countries, largely due to the increased resource requirements of World War II. Since

then, geochemical exploration methods have become increasingly sophisticated due to

advances in analytical instrumentation, yet much is still to be discovered and a greater

understanding to be obtained

3|Page
Geochemical exploration strategies include a number of key phases such as

planning, sampling, chemical analysis, interpretation and follow-up. Each stage is

critical and exploration geologists have to be cautious at each phase to ensure the

integrity of the data that is to be applied to the succeeding stages. Although for a variety

of commodities, or for a specifically targeted commodity, most deposits in a given

geological setting have, on average, the same profile or type deposit settings with

respect to elemental dispersion characteristics over regional landscapes. However,

specific environments within this may contain deposits with a degree of uniqueness (i.e.

geochemical signature) due to the differences in the geological, geomorphological and

environmental setting. Furthermore, many localities have well documented information

from which geochemical exploration studies can be undertaken while others do not.

One of the greatest challenges in geochemical surveys is the confidence to distinguish

between anomalous and background concentration in soils and ground waters. A review

of mining reports and reported data created by Siegal (1974), Rose et al. (1979), Roes

et al. (1979), Levinson (1980), Plant et al. (1989) and Plant et al. (1991) provide a solid

in site. However, due to the targeted nature of these reports extraction of additional data

requires the application of newer analytical methods which were/may not have been

available (or un-needed for the specific task) at the time. For example, for the elements

Bi and W, in some cases (MacDuff, 1971; 1971a; 1971b; 1972 and Fiella, 2010) reports

have honestly stated that because of a lack of references in the literature to similar

sampling programs, the results of their survey and previous data have to be arbitrarily

interpreted.

4|Page
Exploration across depositional landforms, both simple and complex, and with

shallow to deep transported regolith cover by using various geochemical techniques has

reaffirmed the important need to conduct and understand surface geochemistry and to

make it effective for new green field locations. Australia exhibits a unique landscape

with many of the landforms having their origins 300 Ma ago at the beginning of the

break-up of Gondwana (Aspandiar, 2008). Since that time the Australian land mass has

been exposed to wide ranging climatic conditions, from tropical to glacial periods as

well as undergoing various tectonic events which have been well documented (Anand,

2005). The impact of these weathering events has impacted the chemical and physical

environment which has resulted in continuous alteration of exposed rocks (Anand,

R.R., 2005). Furthermore, during most of the Phanerozoic era (0 to 540 Ma), eastern

and south-eastern Australia were undergoing active plate tectonic processes leading to

the amalgamation of various accreted terranes from east to west from approximately

550 Ma up to approximately 200 MA. It has only been recently, during the Quaternary

period where the weathered crust has mostly formed over the last few million years.

This has preserved the weathered mantle as a wide spread cover, 20 to 100m thick, over

much of the landscape. Therefore, weathering in a geochemical exploration context

causes the destruction of primary ore deposits and the dispersion of ore and pathfinder

elements into the surrounding regolith. Conversely, it may also result in the supergene

enrichment of some deposits and promote the formation of secondary ore bodies. To

understand the history and potential mechanisms and pathways of migration of ore and

pathfinder elements in regolith, it is necessary to resolve the complex superposition of

events that may have occurred during regolith-landscape evolution.

5|Page
1.3 BISMUTH AS A PATHFINDER ELEMENT

Bismuth is a rare element which is concentrated in a wide variety of ore types,

principally associated with high-temperature acid intrusives, related skarn deposits and

epithermal systems (Angino and Long, 1979; Lang and Baker, 2001; Baker et al.,

2005). It is (together with Te, W, Sn and other accessory elements) often associated

with gold mineralisation (Angino and Long, 1979; Baker et al., 2005). In the Australian

Figure 1.1: Location of the Cobar, Kingsgate, Wolfram Camp, Tennant Creek and

Henty and Mt Julia mines.

context (Figure 1.1) accessory Bi mineralisation is a feature of the Cobar- style deposits

(Stegman and Reynolds, 2005), the Henty and Mount Julia gold deposits, Tasmania

6|Page
(Callaghan, 2001), certain deposits in the Western Australian shield (Roberson et al.,

2001; Hassan and Clarke, 2005) and the Cu-Au-Bi deposits of the Tennant Creek area,

Northern Territory (Skirrow, 2002). Bismuth minerals are also important constituents

of the New England Orogen of eastern Australia (Figure 1.2), where they have reported

associations with Mo, W and Sn, and with highly acidic and very rich (albeit of low

tonnage) gold deposits of the eastern ranges of Australia (Plimer, 1975; Weber et al.,

1978). It is important to remember that this group of sites is indicative only, and is far

from being comprehensive but allows for a good starting point for the study of

accessory mineralisation of their target materials.

Figure 1.2: Location map showing the distribution of eastern Australian

orogens (Champion, 2009).

7|Page
Bi has been extensively used as a pathfinder element in geochemical exploration

for a variety of deposits (Angino and Long, 1979; Hale, 1981; Robertson et al., 2001;

Collins et al., 2004). The behaviour of Bi in terms of its solution speciation as a

function of pH is well-established (Baes and Mesmer, 1976; Smith and Martell, 1976:

Norman, 1998; Thurston et al., 2005) and provides a general first order assessment of

its mobility in pore solution in the regolith. However, very little is known of the

chemistry of formation of secondary Bi species in the supergene zone or of their

stabilities. To build on existing knowledge a more comprehensive understanding of the

dispersion of Bi in the regolith requires detailed study of its low-temperature aqueous

chemistry, and the investigation of the secondary Bi minerals that serve to buffer the

element between the solid and solution states.

Bismuth is estimated to have an average crustal abundance of 8.5 ppb and is

highly fractionated (Emsley, 1991). Mafic to intermediate igneous rocks average 40

ppb, whereas rhyolites and granites average 900 and 270 ppb, respectively. However,

not all granites exhibit this degree of enrichment (Lueth, 1999). Sedimentary rocks

contain higher amounts, with shales having the greatest average (0.26 ppm) and

limestone the least (30 ppb) (Lueth, 1999). Coal (5 ppm on average) and other organic-

rich materials are considerably enriched in Bi (Brandenstein et al., 1960).

Concentrations greater than 20 ppm have been reported for deep-sea manganese

nodules (Ahrens and Erlank, 1978) and similar enrichment occurs in bauxite and

sedimentary iron oxides (Goldschmidt, 1970), Kabata-Pendias (2001) reported an

average Bi concentration in soils of 0.2 ppm, much lower than previous reports in the

literature where the limits of detection were 5 ppm (McDuff and Snow, 1971).

8|Page
To understand the use of bismuth as a pathfinder element in fresh and marine

waters, Filella (2010) collated and critiqued data and analytical techniques from 125

published papers and tested methods to establish a reasonable typical range of total

dissolved bismuth in natural waters. She concluded that this was impossible to achieve

due to the wide, unexplained and dated dispersion of the data, as well as inadequate

limits of detection of past and present analytical techniques based on literature

published to date. Furthermore, Filella (2010) asserted that it was not possible to make

a confident estimation of bismuth concentrations in seawater and that it was also not

possible to identify a range of probable concentration values for fresh waters which

were not heavily polluted. Thus, in most cases, the use of bismuth as a pathfinder

element is untested to draw any conclusions for the geochemist with respect to

background concentrations or anomalous values.

Like bismuth, tellurium is also used as a geological pathfinding element due to

its occurrence and association with ore-forming elements in hydrothermal deposits

(McPhail, 1995). The speciation of Te has been established (McPhail, 1995) and the

thermodynamic properties for Te-bearing mineral, aqueous and gaseous species have

been reported (Ahmad et al., 1987; Afifi et al., 1988; Jaireth, 1991; Zhang and Spry,

1994; McPhail, 1995). The literature to date forms a solid starting point to assess the

role Te plays in different mineral systems, yet the exact thermodynamic and solubility

data for specific minerals in various geochemical settings are still to be reported. With

this in mind, along with themes presented in previous chapters, there are a number of

reports from geological surveys of various environments (Lett et al. 1998) where Te

and Bi anomaly values have been reported. However, these numbers are arbitrarily

derived or are at the limit of the analytical technique used.

9|Page
To understand the dispersion of bismuth in the supergene environment a review

of the secondary bismuth minerals and their occurrences has been undertaken which is

summarised in Table 1.1. In-depth reviews of specific minerals are indicated at the

beginning of chapters 2 to 5.

Table 1.1 The secondary minerals of bismuth. The number after the formula is the

number of known occurrences as listed by Mindat.org.

Aiolosite Na2(Na2Bi)(SO4)3Cl 1
“Arsenowaylandite” BiAl3(AsO4)2(OH)6 1
Asselbornite1,2 Pb(BiO)3(UO2)4(AsO4)2(OH)7·4H2O 2
Atelestite Bi2(AsO4)O(OH) 35
Beyerite Ca(BiO)2(CO3)2 27
Bismite Bi2O3 174
Bismoclite BiOCl 50
Bismutite (BiO)2CO3 543
Bismutoferrite Fe2Bi(SiO4)2(OH) 47
“Bismutostibiconite” BiSb2O7 6
Bleasdaleite1 (Ca,Fe3+)2Cu5(Bi,Cu)(PO4)4(H2O,OH,Cl)13 1
Bouazzerite Bi6(Mg,Co)11Fe3+14(AsO4)18(OH)4·86H2O 1
Brendelite (Bi,Pb)2(Fe3+,Fe2+)(PO4)O2(OH) 2
Cannonite Bi2(SO4)O(OH)2 16
Chekhovichite Bi2Te4O11 4
Chiluite1 Bi3TeMoO10.5 1
Chrombismite Bi16CrO27 1
Clinobisvanite BiVO4 32
Cobaltneustädtelite Bi2Fe3+(Co,Fe3+)2(AsO4)2(OH,O)4 6
Daubréeite BiO(OH,Cl) 10
Dreyerite BiVO4 3
Dukeite Bi24Cr8O57(OH)6·3H2O 2
Francisite Cu3Bi(SeO3)2O2Cl 2
Gelosaite BiMo6+(2-5x)Mo5+6xO7(OH)·H2O 2
Hechtsbergite Bi2(VO4)O(OH) 4
Juanitaite1 Bi(Cu,Ca,Fe)10(AsO4)4(OH)11·2H2O 4
Kettnerite CaBi(CO3)OF 29
Koechlinite Bi2MoO6 26
Medenbachite Bi2Fe3+(Cu,Fe3+)(AsO4)2(OH,O)4 3
Mixite BiCu6(AsO4)3(OH)6·3H2O 139
Montanite1 Bi2TeO6·2H2O; Bi2(OH)4TeO4 19
Mrázekite Bi2Cu3(PO4)2O2(OH)2·H2O 10
Namibite Cu(BiO)2(VO4)(OH) 23
Neustädtelite Bi2Fe3+(Fe3+,Co)2(AsO4)2(OH,O)4 16
Nickelschneebergite (Bi,Ca)(Ni,Co,Fe)2(AsO4)2·2(OH,H2O) 1
Orthowalpurgite (BiO)4(UO2)(AsO4)2·2H2O 1
Paganoite NiBi(AsO4)O 1

10 | P a g e
Paulkellerite (BiO)2Fe(PO4)(OH)2 1
Perite PbBiO2Cl 16
Petitjeanite Bi3(PO4)2O(OH) 13
Phosphowalpurgite3 (BiO)4(UO2)(PO4)2 3
Pingguite Bi6Te2O13 3
Pottsite1 HPbBi(VO4)2·2H2O 3
Preisingerite Bi3(AsO4)2O(OH) 39
Pucherite BiVO4 46
Riomarinaite BiSO4(OH)·H2O 2
Rooseveltite BiAsO4 26
Russellite Bi2WO6 23
Sardignaite BiMo2O7(OH)·2H2O 1
Schlegelite Bi7(AsO4)3(MoO4)2O4 1
Schneebergite (Bi,Ca)(Co,Ni,Fe)2(AsO4)2·2(OH,H2O) 2
Schumacherite Bi3(VO4)2O(OH) 11
Sillénite Bi12SiO20 14
Smirnite Bi2TeO5 4
Smrkovecite4 Bi2(PO4)O(OH) 2
Sphaerobismoite Bi2O3 2
Šreinite5 Pb(BiO)3(UO2)4(PO4)2(OH)7·4H2O 1
Tetrarooseveltite BiAsO4 1
Uranosphaerite Bi(UO2)O2(OH) 17
Walpurgite (BiO)4(UO2)(AsO4)2·2H2O 31
Waylandite BiAl3(PO4)2(OH)6 25
Ximengite BiPO4 3
Yecoraite1 Fe3Bi5(TeO4)2O9·9H2O 5
Zaïrite Bi(Fe,Al)3(PO4)2(OH)6 2
Zavaritskite BiOF 12
1
Structure unknown. 2Arsenate analogue of šreinite. 3Walpurgite structure type. 4Atelestite structure type.
5
Structure unknown; phosphate analogue of asselbornite ; “ “ unapproved mineral species.
Chemical formulae and mineral approval status derived from the IMA mineral database list (Rakovan,
2007).

An inspection of Table 1.1 indicates that the most common phases reported are

bismutite, bismite, mixite and bismoclite. Stoichiometries indicated in Table 1.1 are

based for the most part on single-crystal X-ray determinations (Anthony et al., 1990;

1995; 1997; 2000; 2003) and reference is made to these where appropriate or to other

sources for more recent structure determinations. Topographical listings in mindat.org

are recognised as being imperfect and other localities for a number of the secondary

minerals may be gleaned from the literature (vide infra). Nevertheless, the tabulation

11 | P a g e
does indicate which phases are common and thus likely to be generally present in

significant amounts in the oxidized zones of Bi-rich ore deposits.

To gain a better understanding and to develop a comprehensive base-line of the

distribution of secondary Bi mineralogy with respect to prospecting in Australia, a

preliminary study on the frequency, occurrence and associates is summarised in

Appendix A Some of the results are based upon more than 150 powder X-ray

diffraction determinations (Rankin et al., 2002; Sharpe and Williams, 2004). A further

study of some 40 specimens from the collections of the Australian Museum has

provided supplementary data for other deposits; rare minerals were encountered but a

distribution pattern of common phases does emerge.

This in turn has a bearing on the selection of species that should be incorporated

in any supergene dispersion model for Bi as several minerals, cannonite,

Bi2(SO4)O(OH)2, for example, are seen to be important even though they are found to

be rare. Of course, each secondary mineral has its own geochemical history and can

only be formed as a result of appropriate prevailing redox, temperature and

concentration conditions; some may be metastable with respect to others. However,

with a view towards a functional model for their influence on bismuth dispersion in the

regolith, very rare species may be largely passed over because they almost certainly

form under geochemical conditions that are rarely encountered.

12 | P a g e
1.4 PROJECT OUTLINE

In summary, the deficiency in the interpretation of Bi anomalies in the regolith

formed as a result of the oxidation of primary Bi-bearing mineralisation and forms the

basis of the argument for this thesis. To draw conclusions on a proper understanding of

the dispersion of Bi in the environment will depend on appreciation of its low-

temperature aqueous chemistry and knowledge of the secondary Bi minerals that serve

to buffer the element between the solid and solution states. Of the known bismuth

minerals reliable thermodynamic data are available only for bismite, bismoclite and

bismutite. This highlights the gap in the literature when attempting to construct a model

for the stability and dispersion of the bismuth secondaries. To determine their

solubility, calculations have been made to deduce reasonable values of bismuth in

ground waters (discussed in Chapter 6). The common and key phases that were taken

into account in this work included:

The minerals bismite, bismoclite, bismutite, cannonite, and riomarinaite,

discussed in Chapter 2.

The minerals rooseveltite, tetrarooseveltite, preisingerite and atelestite discussed

in Chapter 3.

The minerals koechlinite, russellite, and sardignaite discussed in Chapter 4.

The minerals cherckovivhite and smirnite, discussed in Chapter 5.

13 | P a g e
1.5 REFERENCES:

Afifi, A., Kelly, M., and Essene, E.J. (1988) Phase relationships among tellurides,

sulphides and oxides: I. Thermochemical data and calculated equilibria. Economic

Geology, 83, 337-394.

Anand, R.R. (2005) Weathering history, landscape evolution and implications for

exploration In: Anand, R.R. and de Broekert, P. (Eds.) "Regolith Landscape

Evolution Across Australia: a compilation of regolith landscape case studies with

regolith landscape evolution models" CRC LEME 1v, p2-40.

Ahmad, M. Solomon, M., and Walshe, J.L. (1987) Mineralogical and geochemical

studies of the Emperor gold telluride deposit, Fiji. Economic Geology, 82, 234-270.

Ahrens, L.H. and Erlank, A.J. (1978) Bismuth. In: Wedepohl, K.H. (Ed.) Handbook of

Geochemistry. Springer-Verlag, Berlin, Volume II, Sections 83-D-1 to 83-O-1.

Angino, E.E. and Long, D.T. (Eds) (1979) Geochemistry of Bismuth. Dowden,

Hutchinson and Ross, Inc., Stroudsburg, PA, USA

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (1990) Handbook of

Mineralogy Volume 1: Elements, Sulfides, Sulfosalts. Mineral Data Publishing,

Tucson, Arizona.

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (1995) Handbook of

Mineralogy Volume 2: Silica, Silicates. Mineral Data publishing, Tucson, Arizona.

14 | P a g e
Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (1997) Handbook of

Mineralogy Volume 3: Halides, Hydroxides, Oxides. Mineral Data publishing,

Tucson, Arizona.

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (2000) Handbook of

Mineralogy Volume 4: Arsenates, Phosphates, Vanadates. Mineral Data

Publishing, Tucson, Arizona.

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (2003) Handbook of

Mineralogy Volume 5: Borates, Carbonates, Sulfates. Mineral Data Publishing,

Tucson, Arizona.

Aspandiar, M.F., Anand, R.R. and Gray, D.J., (2008) Geochemical dispersion

mechanisms through transported cover: Implications for mineral exploration in

Australia. CRC-LEME, CSIRO Open File Report Series, 230R.

Baes, C.F. Jr and Mesmer, R.E. (1976) The Hydrolysis of Cations. Plenum Press, New

York.

Baker, T., Pollard, P.J., Mustard, R., Mark, G. and Graham, J.L. (2005) A comparison

of granite-related tin, tungsten and gold–bismuth deposits: implications for

exploration. Society of Exploration Geologists Newsletter, 61, 5–17.

15 | P a g e
Brooks, R. R. (1972) Geobotany and biogeochemistry in mineral exploration: New

York, Harper and Row, 290 p

Brandenstein, M., Janda., I. and Schroll, E. (1960) Seltene elemente in österreichischen

kohlen- und bitumengesteinen. Tschermaks Mineralogische und Petrographische

Mitteilungen, 7, 260-269.

Callaghan, T. (2001) Geology and host-rock alteration of the Henty and Mount Julia

gold deposits, western Tasmania. Economic Geology, 96, 1073-1088.

Champion, D.C., Kositcin, N., Huston, D.L., Mathews, E. and Brown, C. (2009)

Geodynamic Synthesis of the Phanerozoic of eastern Australia and implications for

metallogeny. Geoscience Australia Record 2009/18, 255 p.

Clissold, M.E. (2007) Aspects of the supergene geochemistry of copper nickel and

bismuth. PhD thesis unpbulished, University of Western Sydney.

Collins, P., Hooper, B. and Cornelius, M. (2004) Whim Creek Cu-Zn-Pb deposit,

Pilbara, WA. In: Butt, C.R.M., Cornelius, M., Scott, K.M. and Roberson, I.D.M.

(Eds), Regolith Expression of Australian Ore Systems. CRC LEME, Perth, pp. 1-3.

Emsley, J. (1991) The Elements. Second edition, Oxford University Press, Oxford.

Filella, M. (2010) How reliable are environmental data on 'orphan' elements? The case

of bismuth concentrations in surface waters. Journal of Environmental Monitoring,

12, 90-109.

16 | P a g e
Goldschmidt, V.M. (1970) Geochemistry. Oxford University Press, Oxford.

Hale, M., (1981) Pathfinder applications of arsenic, antimony and bismuth in

geochemical exploration. Journal of Geochemical Exploration, 15, 307-323.

Hassan, L.Y. and Clarke, R.M., (2005) Some unusual gold and bismuth mineralization

at Mardoonganna Hill, Murchison mineral field. Annual Review of the Geological

Survey of Western Australia for 2004-2005, Technical Papers, 83-87, and

references therein.

Hawkes, H.E. and Webb, J.S. (1962) Geochemistry in Mineral Exploration. Harper and

Row, New York.

Jaireth, S. (1991) Hydrothermal geochemistry of Te, Ag2Te, and AuTe2 in epithermal

precious metal deposits. Economic Geology, 37. James Cook University of North

Queensland, Australia.

Kabata-Pendias, A. (2001) Trace Elements in Soils and Plants. 3rd Edition, CRC Press,

Boca Raton, Florida.

Lang, J.R. and Baker, T. (2001) Intrusion-related gold systems: the present level of

understanding. Mineralium Deposita, 36, 477-489.

Lett, R.E., Jackaman, W. and Yeow, A. (1998) Detailed Geochemical Exploration

techniques for base and precious metals in the Kootenay Terrane. Geological

Fieldwork 1998. 297-306.

17 | P a g e
Levinson, A.A., (1980) Introduction to Exploration Geochemistry: The Supplement,

Applied Publishing Ltd, Wilmette, USA, 1980, 615-924.

Lueth, V.W. (1999), Bismuth: Element and Geochemistry. In: Fairbridge, R.W. (Ed.)

The Encyclopaedia of Geochemistry. Kluwer Academic Publishers, London, 43-44.

MacDuff, R. and Snow, A. (1971) Quarterly Report Period to 8th January 1971, EL208,

AOG Minerals PL. New South Wales Department of Primary Industries (Mineral

Resources) Open File Reports, No. R00028059.

MacDuff, R. and Zerwick, J. (1971a) EL208, Quarterly Report Period to 8th July 1971,

AOG Minerals PL. New South Wales Department of Primary Industries (Mineral

Resources) Open File Reports, No. R00007895.

MacDuff, R. and Zerwick, J. (1971b) EL208, Quarterly Report Period to 8th October

1971, AOG Minerals PL. New South Wales Department of Primary Industries

(Mineral Resources) Open File Reports, No. R00007896.

MacDuff, R. and Zerwick, J. (1972) EL208, Quarterly Report Period to 8th January

1972, AOG Minerals PL. New South Wales Department of Primary Industries

(Mineral Resources) Open File Reports, No. R00007897.

McPhail, D.C. (1995) Thermodynamic properties of aqueous tellurium species between

25 and 350°C. Geochimica et Cosmochimica Acta, 59, 5, 851-866.

18 | P a g e
Norman, N.C., (1998) Chemistry of Arsenic, Antimony and Bismuth. Blackie, London.

Plant, J.A., Breward, N., Forrest, M.D. and Smith, R.T. (1989) The gold pathfinder

elements As, Sb and Bi - their distribution and significance in the southwest

Highlands of Scotland. Transactions of the Institution of Mining and Metallurgy,

98, B91-101.

Plant, J.A., Cooper, D.C., Green, P.M., Reedman, A.J. and Simpson, P.R. (1991)

Regional distribution of As, Sb and Bi in the Grampian Highlands of Scotland and

English Lake District: implications for gold metallogeny. Transactions of the

Institution of Mining and Metallurgy, 100, B135-147.

Plimer, I., (1975) Wolfram Camp wolframite-molybdenite-bismuth-quartz pipes, north

Queensland. In: N Knight C.L. (Ed.), Economic Geology of Australia and Papua

New Guinea - 1. Metals. Australasian Institute of Mining and Metallurgy,

Melbourne, Monograph Series, 5, 760-762.

Rakovan, J. (2007) Words to the Wise- More than 4000 To Be Exact. Rocks and

Minerals, 82, 423-424.

Rankin, J., Sharpe, J.L. and Williams, P.A. (2001) Betpakdalite from the tin deposits of

Elsmore, New England district of New South Wales. Australian Journal of

Mineralogy, 7, 15-17.

19 | P a g e
Rankin, J., Lawrence, L.J., Sharpe, J.L. and Williams, P.A., (2002) Rare secondary

bismuth, tungsten and molybdenum minerals from Elsmore, New England district

of New South Wales. Australian Journal of Mineralogy, 8, 55-60.

Roberson, I.D.M., King, J.D. and Anand, R.R., (2001) Regional geology and

geochemical exploration around the Stellar and Quasar gold deposits, Mt Magnet,

Western Australia. Geochemistry: Exploration, Environment, Analysis, 1, 353-364.

Rose, A.W., Hawkes, H.E. and Webb, J.S., (1979) Geochemistry in Mineral

Exploration Academic Press, New York, N.Y., pp. 490--517.

Siegel, F. R., (1974) Applied Geochemistry: New York, John Wiley and Sons, 353p.

Sharpe, J.L. and Williams, P.A., (2004) Secondary bismuth and molybdenum minerals

from Kingsgate, New England district of New South Wales. Australian Journal of

Mineralogy, 10, 7-12.

Skirrow, R.G., (2002) Copper-gold-bismuth deposits of the Tennant Creek district,

Australia: a reappraisal of diverse high-grade systems. In: Porter, T.M. (Ed.),

Hydrothermal Iron-Oxide Copper-Gold and Related Deposits. PGC Publishing,

Adelaide, pp. 149-160.

Smith, R.M. and Martell, A.E., (1976) Critical Stability Constants. 4. Inorganic

Complexes. Plenum Press, New York.

20 | P a g e
Stegman, C. and Reynolds, I., (2005) Primary mineralisation in Cobar deposits, with

emphasis on gold deposits of the Cobar Goldfield. Australian Journal of

Mineralogy, 11, 63-72, and references therein.

Thurston, J.H., Swenson, D.C. and Messerle, L., (2005) Solvolytic routes to new

nonabismuth hydroxy- and alkoxy-oxo complexes: synthesis, characterization and

solid-state structures of novel nonabismuth polyoxo cations Bi9(μ3-O)8 (μ 3-OR)65+

(R: H, Et). Chemical Communications, 4228-4230.

Weber, C.R., Paterson, I.B.L. and Townsend, D.J., (1978) Molybdenum in New South

Wales. Geological Survey of New South Wales Mineral Resources Series, 43.

Zhang, X. and Spry, P.G. (1994) Calculated stability of aqueous tellurium species,

calaverite, and hessite at elevated temperatures, Economic Geology, 87, 1152-

1166.

21 | P a g e
CHAPTER 2

BISMOCLITE, BISMUTITE, CANNONITE AND


RIOMARINAITE

22 | P a g e
2.1 ABSTRACT

From studying the relative frequency, associations and occurrence of secondary

Bi minerals found from a host of deposits, bismite, Bi2O3, bismoclite, BiOCl,

bismutite, (BiO)2CO3, and cannonite Bi2(SO4)O(OH)2 appear to play a significant role

in the immobilisation of Bi in oxidised settings. Thus, these phases are important in

developing a preliminary model for the dispersion of Bi in the supergene environment.

Thermochemical data was determined for cannonite, ΔGfө(Bi2(SO4)O(OH)2, 298.15K)

= –1315.98±0.5 kJ mol–1, allowing for the development of a geochemical model.

Understanding the immobilising capacity of these important secondary minerals, under

various conditions, further develops our understanding of the dispersion and mobility of

bismuth in the supergene zone.

2.2 INTRODUCTION

The minerals of interest that were studied in this chapter were bismite,

bismoclite, bismutite, cannonite and riomarinaite. The discovery and associations are

discussed below in section 2.2.1. These were chosen due to their known associations

and their common assemblage associations as shown in the compiled data of Appendix

A. Other rare or uncommon phases which are of note, however, were not studied in

depth are reported in section 2.2.2.

2.2.1 Bismite, bismoclite, bismutite, cannonite and riomarinaite

Bismite Bi2O3 as a mineral was first described in 1753, a definitive analysis

was first reported in 1848 (Frondel, 1943). The single-crystal structure of bismite was

23 | P a g e
reported by Sillén (1940) and refined by Malmros (1970). The Colavi mine in the

Machacamarca district, Cornelio Saavedra Province, Potosí Department, Bolivia is

taken to be the type locality and bismite is reported in the vast majority of bismuth

deposits around the world.

Two basic carbonates of bismuth are known, bismutite, Bi2O2CO3 (Figure 2.1),

and a synthetic phase of composition Bi4O4CO3(OH)2 (Taylor et al., 1984). The

stabilities of these phases are known (Clissold, 2007). Bismutite was first named by

Breithaupt (1841) using material from Ullersreuth, Vogtland, Germany. However, a

single type locality cannot be definitely ascribed. Frondel (1943) reviewed the available

evidence and concluded that the mineral had the composition Bi2O2CO3. The single

crystal structure was originally reported by Lagerkrantz and Sillén (1948) and refined

by Grice (2002). The predominance of bismutite amongst the vast majority of Bi-rich

deposits consequently renders it the most common secondary Bi mineral. Bismite,

Bi2O3, is a rare phase by comparison and it should be noted again that Bi2O3 reacts with

CO2(g) under ambient conditions to give Bi2O2CO3 (Barreca et al., 2001).

The bismuth halide bismoclite, BiOCl, was first described from Jackal’s Water,

Steinkopf, Namaqualand, Northern Cape Province, South Africa, by Mountain (1937).

The structure was reported by Bannister (1935) and refined by Keramidas et al. (1993).

Bismoclite is commonly associated with bismutite, bismuthinite, jarosite, alunite, and

cerussite.

24 | P a g e
Figure 2.1: Bismutite from Schmiedestollen dump, Wittichen, Schenkenzell, Black
Forest, Baden-Württemberg, Germany (Bonacina, 2011)

The minerals cannonite, Bi2(SO4)O(OH)2, (Figure 2.4) and riomarinaite

BiSO4(OH)·H2O form under typical secondary conditions. The structure of cannonite

was reported by Golic et al. (1982), Sadanaga and Bunno (1974) first reported

cannonite from the Ashio mine, Ashio, Tochigi Prefecture, Japan, where it was

associated with native bismuth, bismuthinite, Bi sulfosalts, koechlinite, and zavaritskite.

However, a specimen from the Tunnel Extension No. 2 mine, Bullion Canyon, Piute

County, Utah, USA is taken to be the type locality (Stanley et al., 1992). The structure

of riomarinaite, BiSO4(OH)·H2O, was reported by Graunar and Lazarini (1982).

According to Rögner (2005) riomarinaite was found at the Falcacci stope, Rio Marina

mine, Elba Island, Tuscany, Italy and it was associated with primary phases including

25 | P a g e
bismuth, bismuthinite, Bi sulfosalts and molybdenite and secondary minerals

bismoclite, bismutite, cannonite and daubréeite.

Figure 2.2: Cannonite from the Duadello mine (Baitello mine), Palotto Valley (Palot
Valley), Fraine, Pisogne, Camonica Valley, Brescia Province, Lombardy, Italy
(Bonacina, 2009).

2.2.2 Rare and other bismuth oxide, carbonates, chloride, and sulfate minerals
The structure for sphareobismoite was reported by Blower and Greaves (1988)

from the Schmiedestollen mine, Wittichen, Baden-Württemberg, Germany which was

the reported type locality (Walenta, 1995), where it was found to be the oxidation

product of wittichenite and emplectite. Sphareobismoite is predominantly associated

with bismutite and mixite. Other associated secondary Bi minerals include atelestite,

beyerite, bismite, bismutoferrite, orthowalpurgite, preisingerite, rooseveltite and

26 | P a g e
walpurgite (Walenta, 1992, 1995). Sphaerobismoite is a rare phase compared to

bismutite, which was is found repeatedly in Nature.

The type locality of beyerite, Ca(BiO)2(CO3)2 is Schneeberg, Saxony, Germany,

which was first reported by Frondel (1943) where it was associated with bismutite.

Reported as a carbonate of Bi and Ca, a definite formula was not confirmed until later

investigations by Heinrich (1947). The single crystal structure is known, originally

reported by Lagerkrantz and Sillén (1948) and later refined by Grice (2002).

Another Bi-halide, perite, PbBiO2Cl, is commonly associated with bismutite.

The orthorhombic, pseudotetragonal structure for perite was reported by Gillberg

(1960) space group Bmmb, a = 5.591, b = 5.431 and c = 12.200 Å. However, Bridge

(1976) reported that perite from Glen Florrie Station, Western Australia, was tetragonal

(a = 5.579(3) and c = 12.45(1) Å). Both orthorhombic and tetragonal forms are known

synthetically (Deschanvres et al., 1970; Ketterer and Krämer, 1985; Lopatkin, 1987).

The minerals aiolosite, Na2(Na2Bi)(SO4)3Cl, campostriniite, Bi12.67O14(SO4)5,

and leguernite, (Bi,Na)3(Na,K)4(SO4)6·H2O, have been reported as primary volcanic

sublimates or as the reaction product of sublimates (Demartin et al., 2010; Demartin et

al., 2013; Garavelli et al., 2013). As such, they have not been negated from the future

consideration regarding secondary Bi mineralogy.

27 | P a g e
2.3 EXPERIMENTAL

2.3.1 Syntheses of cannonite and riomarinaite

Cannonite was synthesised by adding, H2SO4, (49.78 mL, 0.1067 M, AR ≥98%)

to a flask containing Bi2O3 (0.51315 g, ReagentPlus® 99.9%) and magnetically stirred

(500 rpm) at room temperature. After 1 hour the initially yellow powder changed colour

to white, however the mixture was left to stir for a further 5 days. The solid was then

collected at the pump on GF/F grade filter paper, rinsed with DI water and acetone, and

sucked dry.

Riomarinaite was synthesised by mixing H2SO4 (49.78 mL, 1.067 M, AR

≥98%) in a flask containing Bi2O3 (0.5093 g, ReagentPlus® 99.9%) and magnetically

stirred (500 rpm). The mixture was again observed to change colour after 1 hour from

yellow to white but was left to stir for 5 days. The precipitate was isolated as above.

Powder X-ray diffraction studies of the products were carried-out using a

Bruker D8 Advance diffractometer (Ni-filtered CuKα1 radiation, λ = 1.5406 Å, 40 kV

and 30 mA). Traces were produced between 5-70o 2θ, with a step size of 0.02o and a

rate of 1.2 o min–1. Diffraction Technology Data processing software (Eva) and JCPDS-

ICDD data base files were used to identify the phases produced.

28 | P a g e
2.3.2 Solubility Studies

Solubility studies were undertaken using sealed 250 cm3 conical QuickfitR flasks

maintained at 25.0  0.2oC in a thermostatted water bath inside a controlled temperature

room at the same temperature. Measurements of pH were made using a Radiometer

ION450 apparatus fitted with a combination electrode. The solubility studies were

conducted by precipitation of the desired product following the synthetic procedures for

cannonite and riomarinaite. Bismuth oxide (ca 0.5 g) was added to a series of seven

conical flasks containing 49.97 cm3 of standardised H2SO4; 0.1067 M for cannonite and

1.067 M for riomarinaite. The flasks were left for 10 weeks during which time the pH

of a paired flask was monitored periodically until the pH of the test mixture remained

stable. The solutions were filtered using GF/F grade filter paper and the filtrates

collected in clean PET specimen jars. Dissolved Bi concentrations were determined

using ICP-MS by a NATA-compliant commercial laboratory (mgt|LabMark

Environmental Testing Australia Pty Ltd).

29 | P a g e
2.4 RESULTS

2.4.1 Synthesis of cannonite and riomarinaite

High purity, single phase samples of cannonite and riomarinaite were obtained

in quantitative yield from the synthesis and dissolution of cannonite as shown by

powder X-ray diffraction. Unit cell parameters were refined in LAPOD (Langford,

1973) and values obtained were a = 7.690(2), b = 13.861(6) and c = 5.682(2) Å for

cannonite and a = 6.023(2), b = 13.362(1) and c = 6.490(5) Å for riomarinaite which

are in good agreement with those reported by Golic et al. (2007) and Graunar and

Lazarini (1982) respectively.

2.4.2 Solubility Studies

Equilibrium was achieved after 6 weeks for cannonite but not for riomarinaite,

X-ray diffraction examination of the original riomarianite samples after 2 weeks and 10

weeks in solution indicated the formation of cannonite. Further solubility studies of

riomarinaite were not pursued due to the apparent metastability. The apparent

metastability of riomarinaite in contact with its mother liquor is shown in the X-ray

diffraction traces presented in Figures 2.3-2.5 below.

30 | P a g e
1200

1100

1000

900

800

700
Lin (Counts)

600

500

400

300

200

100

5 10 20 30 40 50 60 70

2-Theta - Scale

Figure 2.3: Powder XRD trace of riomarinaitre with peak positions of the ‘type’ pattern from JCPDS-ICDD file 01-076-1103 (red).

34 | P a g e
430
420
410
400
390
380
370
360
350
340
330
320
310
300
290
280
270
260
250
Lin (Counts)

240
230
220
210
200
190
180
170
160
150
140
130
120
110
100
90
80
70
60
50
40
30
20
10
0

6 10 20 30 40 50 60 70 80

2-Theta - Scale

Figure 2.4: Powder XRD trace of riomarinaite after 2 weeks showing a mixture of riomarinaite and cannonite with peak positions of the ‘type’

patterns from JCPDS-ICDD files for riomarinaite 01-076-1103 (red) and cannonite 00-045-1439 (blue).

35 | P a g e
600

500

400
Lin (Counts)

300

200

100

6 10 20 30 40 50 60 70

2-Theta - Scale

Figure 2.5: Powder XRD trace of riomarinaite after 10 weeks converted to cannonite with peak positions of the ‘type’ pattern for

cannonite from JCPDS-ICDD file 00-045-1439.

36 | P a g e
Cannonite dissolved incongruently in aqueous H2SO4 as expressed in equation
(1).

Bi2(SO4)O(OH)2(s) + 3H+(aq) ⇋ BiOH2+(aq) + H2O(l) + HSO4-(aq) (1)

Individual ion activity coefficients were calculated using the Debye-Hückel

equation for 298.15 K, lg  = –0.5085z2(I/(1+I). For cannonite I = 0.128 mol dm–3,

3 = 0.093 2 = 0.348 and  = 0.768. In all instances, o is taken to be unity. The

activity of H+(aq), a(H+), in H2SO4(aq) (0.1084 M) was then calculated and

corresponding activity a(Bi3+), was calculated from the solubility data (Table 2.1).

Table 2.1 Dissolved metal concentrations for cannonite

Solution [Bi] ppm [Bi] mol dm-3


1 22 1.053 x 10-4
2 21 1.000 x 10-4
3 23 1.101 x 10-4
4 21 1.000 x 10-4
5 21 1.000 x 10-4
6 21 1.000 x 10-4
Mean 21.50 1.028 x 10-4
Error ±1.4 ±2.05 x 10-5

The pH at equilibrium (1.047) was used in the determination of individual ion

speciation using COMICS (Perrin and Sayce, 1967). Reliable lg K values for equation

(2), (4), (6), (8), (10), (12) and (14) of lg K(298.15 K) = –1.11, –3.30, –8.21, 2.01, 3.41,

4.60 and 4.88 at I = 0 mol dm–3 were reported by Fedorov et al. (1971) and van Der Lee

37 | P a g e
and Lomenech (2004). Correction to I = 0.128, mol dm–3 by the method of Baes and

Mesmer (1976), yields lg K(298.15 K) = –1.73, –4.30 and –9.33. Using the

relationships shown by equations (3), (5) and (7), (9), (11), (13) and (15), the

concentration of Bi(OH)2+(aq), Bi(OH)2+(aq) and Bi(OH)3o(aq) are 2.20 x 10-5, 7.20 x

10-7 and 1.30 x 10-10 mol dm–3 respectively.

Bi3+(aq) + H2O(l) ⇋ Bi(OH)2+(aq) + H+(aq) (2)

[Bi(OH)2+].[H+]/[Bi3+] = 10–1.11 (3)

Bi3+(aq) + 2H2O(l) ⇋ Bi(OH)2+(aq) + 2H+(aq) (4)

[Bi(OH)2+].2[H+]/[Bi3+] = 10–3.30 (5)

Bi3+(aq) + 3H2O(l) ⇋ Bi(OH)30(aq) + 3H+(aq) (6)

[Bi(OH)30].3[H+]/[Bi3+] = 10–8.21 (7)

Bi3+(aq) + SO42-(aq) ⇋ BiSO4+ (aq) (8)

[BiSO4+]/[Bi3+].[ SO42] = 10+2.01 (9)

Bi3+(aq) + 2SO42-(aq) ⇋ Bi(SO4)2- (aq) (10)

[BiSO4+]/[Bi3+].2[ SO42] = 10+3.51 (11)

Bi3+(aq) + 3SO42-(aq) ⇋ Bi(SO4)33- (aq) (12)

[BiSO4+]/[Bi3+].3[ SO42] = 10+4.60 (13)

Bi3+(aq) + 4SO42-(aq) ⇋ Bi(SO4)4 5- (aq) (14)

[BiSO4+]/[Bi3+].4[ SO42] = 10+4.88 (15)

38 | P a g e
This yields a value of lg K for equation (1) –8.00 ± 0.50 and yields a

corresponding value for ΔGrө(298.15 K) = +45.68 kj.mol-1. Thus,

ΔGfө(Bi2(SO4)O(OH)2, s, 298.15 K) = –1315.98 ± 0.5 kJ mol–1 The estimated error

takes into account the analytical error of the solubility experiments and errors quoted

for the thermochemical data used in Table 2.2.

Table 2.2 Thermodynamic quantities used in the calculations (T = 298.15 K).

ΔGfө/kJ mol-1 Reference


bismoclite –322.1 Kaye and Laby (1995)
bismutitea –916.2 ± 7.5 see text
cannonite –716.91 ± 0.5 this study
BiSO4+ (aq)b –663.65 see text
Bi(SO4)2-(aq)b –1415.64 see text
Bi(SO4)33-(aq)b –2177.91 see text
Bi(SO4)45-(aq)b –2931.5 see text
Bi(OH)3o(aq)c –572.61 see text
Bi(OH)2+(aq)c –363.54 see text
Bi(OH)2+(aq)c –138.94 see text
Bi3+(aq)c +91.82 see text
CO2(g) –394 Robie and Hemingway (1995)
Cl-(aq) –131.2± 0.1 Robie and Hemingway (1995)
H2O(l) –237.1± 0.1 Cox et al. (1989)
a
Values have been recalculated using values for H2O(l), CO2(g) and Bi2O3(s) at 298.2 K
taken from Robie and Hemingway (1995).bCalculated using values from Fedorov et al.
(1971) and dissociation constants of sulfuric acid from Martell and Smith (1982).
c
Calculated using standard electrode potentials from Lovreček et al. (1985) to determine
ΔGfo (Bi3+) and dissociation constants from van Der Lee and Lomenech (2004).

39 | P a g e
2.5 REFERENCES

Baes, C.F. Jr and Mesmer, R.E. (1976) The Hydrolysis of Cations. Plenum Press, New

York.

Barreca, D., Morazzoni, F., Rizzi, G.A, Scotti, R. and Tondello, E. (2001) Molecular

oxygen interaction with Bi2O3: a spectroscopic and spectromagnetic investigation.

Physical Chemistry and Chemical Physics, 3, 1743-1749.

Blower, S.K. and Greaves, C. (1988) The structure of β-Bi2O3 from powder neutron

diffraction data. Acta Crystallographica, C44, 587-589.

Bonacina, E., (2009) Cannonite: Bi2(SO4)O(OH)2 Duadello Mine (Baitello mine),

Palotto Valley (Palot Valley), Fraine, Pisogne, Camonica Valley, Brescia Province,

Lombardy, Italy, photograph, viewed 4 January 2013,

<http://www.mindat.org/photo-245067.html>.

Bonacina, E., (2011) Bismutite : (BiO)2CO3 Schmiedestollen dump, Wittichen,

Schenkenzell, Black Forest, Baden-Württemberg, Germany, photograph, viewed 4

January 2013, <http://www.mindat.org/photo-419556.html l>.

Breithaupt, A. (1841) Über das natürliche kohlensaure Wismutoxyd. Poggendorffs

Annalen der Physik und Chemie, 53, 627-630.

Clissold, M.E. (2007) Aspects of the supergene geochemistry of copper nickel and

bismuth. Unpublished PhD thesis, University of Western Sydney.

40 | P a g e
Cox, J.D., Wagman, D.D. and Medvedev, V.A. (1989) CODATA Key Values for

Thermodynamics. Hemisphere Press, New York.

Demartin, F., Gramaccioli, C.M., Campostrini, I. and Pilati, T. (2010) Aiolosite,

Na2(Na2Bi)(SO4)3Cl, a new sulfate isotypic to apatite from La Fossa Crater,

Vulcano, Aeolian Islands, Italy. American Mineralogist, 95, 382-385.

Demartin, F., Gramaccioli, C.M., Campostrini, I. and Pilati, T. (2013) Campostriniite,

IMA 2013-086, CNMNC Newsletter No.18, December 2013, page 3255 -

Mineralogical Magazine, 77, 3249- 3258.

Fedorov, V.A., Kalosh, T.N., Chernikova, G.E. and Mironov, V.E. (1971) Sulphato-

complexes of bismuth(III). Russian Journal of Physical Chemistry, 45, 106.

Frondel, C. (1943) Mineralogy of the oxides and carbonates of bismuth. American

Mineralogist, 28, 521-535.

Golic, L., Graunar, M. and Lazarini, F. (1982) Catena-di-μ-hydroxo-μ3-oxo-

dibismuth(III) sulfate. Acta Crystallographica, B38, 2881-2883.

Grice, J.D. (2002) A solution to the crystal structures of bismutite and beyerite.

Canadian Mineralogist, 40, 693-698.

41 | P a g e
Graunar, M. and Lazarini, F. (1982) Di-μ-hydroxo-bis[aquasulfatobismuth(III)]. Acta

Crystallographica, B38, 2879-2881.

Kaye, G.W.C. and Laby, T.H. (1995) Tables of Physical and Chemical Constants. 16th

Edition, Longman, London, Section 3.10.5.

Lagerkrantz, A. and Sillén, G. (1948) On the crystal structure of Bi2O2CO3 (bismutite)

and CaBi2O2(CO3)2 (beyerite). Arkiv för Kemi, Mineralogi och Geologi, 25, 1-21.

Langford, J.I. (1973) Least-squares refinement of cell dimensions from powder data by

Cohen's method. Journal of Applied Crystallography, 6, 190-196.

Lovreček, B., Mekjavić, I. and Metikoš-Huković, M. (1985) Bismuth. In: Bard, A.J.,

Parsons, R. and Jordan, J. (Eds.) Standard Potentials in Aqueous Solution.

International Union of Pure and Applied Chemistry and Marcel Dekker, New York,

180-187.

Malmros, G. (1970) The crystal structure of α-Bi2O3. Acta Chemica Scandinavica, 24,

384-396.

Martell, A.E. and Smith, R.M. (1982) Critical Stability Constants. Volume 5:First

Supplement. Plenum Press, New York.

42 | P a g e
Mountain, E.D. (1937) Two new bismuth minerals from South Africa. Mineralogical

Magazine, 24, 59-64.

Perrin, D.D. and Sayce, I.G. (1967) Computer calculation of equilibrium concentrations

in mixtures of metal ions and complexing species. Talanta, 14, 833-842.

Rögner, P. (2005) Riomarinait, ein neues Wismutmineral vom Abbau Falcacci, Rio

Marina, Elba (Italien). Aufschluss, 56, 53-60.

Sadanaga, R. and Bunno, M. (1974) The Wakabayashi Mineral Collection. Bulletin of

the University Museum, University of Tokyo, 7

Sillén, L. (1940) Die Kristallstruktur des monoklinen α-Bi2O3. Naturwissenschaften, 28,

206-207.

Stanley, C.J., A.C. Roberts, D.C. Harris, A.J. Criddle, and J.T. Szymański (1992)

Cannonite, Bi2O(OH)2SO4, a new mineral from Marysvale, Utah, USA.

Mineralogical Magazine, 56, 605-609.

Taylor, P., Sunder, S. and Lopata, V.J. (1984) Structure, spectra, and stability of solid

bismuth carbonates. Canadian Journal of Chemistry, 62, 2863-2873.

van der Lee, J. and Lomenech, C. (2004) Towards a common thermodynamic database

for speciation models. Radiochimica Acta, 92, 811-818.

43 | P a g e
Walenta, K. (1992) Die Mineralien des Schwarzwaldes. Weise, Munich.

Walenta, K. (1995) Sphaerobismoite, a new mineral of the composition Bi2O3 from the

Black Forest. Der Aufschluss, 46, 245-248.

44 | P a g e
CHAPTER 3

ROOSEVELTITE, PREISINGERITE AND ATELESTITE

45 | P a g e
3.1 ABSTRACT

Secondary Bi-As minerals play a role in determining the extent of Bi dispersion

in the supergene environment. Syntheses and stability studies of atelestite,

Bi2(AsO4)O(OH), preisingerite, Bi3(AsO4)2O(OH), tetrarooseveltite, BiAsO4 and

rooseveltite, BiAsO4 have been undertaken in order to elucidate their contribution to Bi

immobilisation. Solubilities in aqueous HNO3 were determined at 298.2 K and the data

obtained used to calculate values of ΔGfө at the same temperature. The derived ΔGfө

(298.2 K) values for atelestite, (–1102.58 ± 5.6 kJ mol–1), preisingerite, (–1823.29 ± 2.5

kJ mol–1) and rooseveltite (–716.91 ± 1.5 kJ mol–1) have been used in subsequent

calculations to determine their relative stabilities and relationships with other secondary

Bi minerals.

3.2 INTRODUCTION

The minerals atelestite, Bi2(AsO4)O(OH), preisingerite, Bi3(AsO4)2O(OH),

tetrarooseveltite, BiAsO4 and rooseveltite, BiAsO4, have been reported from multiple

localities worldwide (Chapter 2, Table 1) and are the most abundant of the Bi-As

minerals found. Bismuth arsenates are rare minerals, but they can be a common feature

of some bismuth geochemical settings which are discussed herein. Therefore, it is

important to understand the role As plays in limiting the dispersion of Bi in the

oxidising environment. In an Australian context, the minerals rooseveltite and

preisingerite have been reported from the Elsmore mine, Elsmore, New South Wales,

Australia (Rankin et al., 2002, 2013), yet atelestite to date has never been reported.

Rooseveltite and preisingerite were found to be closely associated with bismoclite

(Rankin et al., 2013) at this locality. Bismuth and bismuthinite are primary phases.

46 | P a g e
Other secondary Bi minerals identified from Elsmore included bismite, bismutite,

“bismutostibiconite”, cannonite, russellite, sardignaite, waylandite and zavaritskite

(Rankin et al., 2002, 2013). This chapter has focused on the chemical modelling of

bismuth arsenates the data presented cab be used to determine the roles theses phases

play at different mine sites globally. As such the use of data from previous chapters can

be included to draw conclusions on the dispersion of bismuth in the environment.

Atelestite (Figure 3.1) was first reported from the Neuhilfe mine, Schneeberg,

Saxony, Germany (Breithaupt, 1832). It is often associated with primary bismuth

minerals and bismuth sulfosalts as well as secondary Bi minerals including, bismutite,

bismutoferrite and paulkellerite (Breithaupt, 1832; Dunn et al., 1988; Weiss, 2007)

mixite, preisingerite, rooseveltite and walpurgite (Palache et al., 1951; Walenta, 1972,

1992) and beyerite, bismite, orthowalpurgite and sphaerobismoite (Walenta, 1992,

1995a). The structure of atelestite was reported by Mereiter and Preisinger (1986).

Figure 3.1: 0.5 mm balls of yellow atelestite overgrowing spherules of bismutite. From
Hérival, Val-d'Ajol, Vosges, Lorraine, France. (Valverde, 2006)
47 | P a g e
Both the structure and the mineral occurrence of preisingerite, Bi3(AsO4)2O(OH),

(Figure 3.2) were first reported from the San Francisco de los Andes, Calingasta, San

Juan, Argentina (Bedlivy and Mereiter, 1982b) where it was found to be associated with

bismuth, bismuthinite, Bi sulfosalts and rooseveltite (Bedlivy et al., 1972; Bedlivy and

Mereiter, 1982b). Other secondary Bi mineral associations reported with preisingerite

include beyerite, bismutite, pucherite and walpurgite (Schlegel et al., 1996),

neustädtelite (Meier and Dünkel, 2010), bismutite, bismutoferrite and neustädtelite

(Meier and Dünkel, 2010), as well as atelestite, bismite, bismoclite, bismutoferrite,

mixite, rooseveltite and walpurgite (Markl, 1990; Walenta, 1992).

Figure 3.2: Preisingerite from the Clara Mine, Rankach valley, Oberwolfach, Wolfach,
Black Forest, Baden-Württemberg, Germany, picture width 2 mm (Wolfsried, 2007).

Rooseveltite, BiAsO4, (Figure 3.3) was first described by Herzenberg (1946)

whereby the specimen analysed was from the Santiaguillo, Macha, Chayanta Province,

Potosí Department, Bolivia (Herzenberg, 1946), but interestingly no other Bi minerals

have been reported from this site. Rooseveltite’s structure was first reported by Bedlivy

48 | P a g e
and Mereiter (1982a). Rooseveltite has had few reported localities (Anthony et al.,

2000) and, as mentioned previously, was reported from Australia at the Elsmore mine,

NSW.

Figure 3.3 : Rooseveltite aggregates (white to pale yellow: ~1mm) and marcasite,FeS2,
(metallic green) from the Lagoa Mine, Estorãos, Ponte de Lima, Viana do Castelo
District, Portugal (Alves, 2010).

Tetrarooseveltite, BiAsO4, has a known structure (Mooney, 1948). The type locality

and only known specimen reported to date is the Moldava deposit, Krušnéhory

Mountains, Bohemia, Czech Republic. The specimen was associated with bayldonite,

malachite, and mimetite in a fluorite-barite-quartz vein (Sejkora and Řídkošil, 1994);

primary Bi phases are bismuth, bismuthinite and Bi sulfosalts; other secondary minerals

are bismutite, mixite, preisingerite, rooseveltite and zavaritskite (Sejkora, 1994; Sejkora

and Řídkošil, 1994).

49 | P a g e
3.3 EXPERIMENTAL

3.3.1 Syntheses

Rooseveltite was synthesised by mixing arsenic acid (50mL, 0.1 M) with

riomarinaite (0.5 g) in an acid digestion bomb, which was sealed and heated at 180 oC

for three days. The mixture was allowed to cool to ambient temperature in the bomb.

The obtained solid was then collected at the pump on GF/F grade filter paper, rinsed

with DI water and acetone, and sucked dry.

Atelestite was synthesised by adding Bi2O3 (0.5013 g, ReagentPlus® 99.9%) and

Na2HAsO4∙7H2O (0.5001 g, ≥98%), mixed in an acid digestion bomb with deionised

water (50 mL). The mixture was buffered to pH 4.39 with 2 M KOH. The bomb was

then placed in an oven (190oC) for eleven days. The bomb was removed and allowed to

cool to ambient temperature and the product was collected as above.

Preisingerite was synthesised by mixing arsenic acid (100.00 mL, 0.1 M) to a

round bottomed flask containing riomarinaite (0.5 g). The mixture was then refluxed for

48 hours and the product was isolated as above.

Powder X-ray diffraction studies of the products were carried-out using a

Philips PW1825/20 powder diffractometer (Ni-filtered CuKα1 radiation, λ = 1.5406 Å,

40 kV, 30 mA). Traces were produced between 5-70o 2θ, with a step size of 0.02 o and a

rate of 1.2 o min–1. Diffraction Technology Data processing software (Traces Version 6)

and JCPDS-ICDD data base files were used to identify the phases produced.

50 | P a g e
3.3.2 Solubility Studies

Solubility studies were undertaken using sealed 250 cm3 conical QuickfitR flasks

maintained at 25.0 0.2oC in a thermostatted water bath inside a controlled temperature

room at the same temperature. Measurements of pH were made using a Radiometer

ION450 apparatus fitted with a combination electrode. The minerals (ca 0.1 g) were

added to a series of conical flasks containing 100 cm3 of standardised HNO3; 0.0981 M.

The flasks were left for 10 weeks during which time the pH of a paired flask was

monitored periodically until no change was detected. The remaining solutions were

collected by filtration through WhatmanR GF/F fibreglass filter papers and stored in

clean PET bottles. Dissolved Bi/As concentrations were determined using ICP-MS by a

NATA-compliant commercial laboratory (mgt | LabMark Environmental Testing

Australia Pty Ltd).

3.4 RESULTS

3.4.1 Syntheses of rooseveltite, tetrarooseveltite, atelestite and preisingerite

High purity samples of atelestite, preisingerite and rooseveltite were obtained in

an essentially quantitative yield. Powder X-ray diffraction did not give any evidence for

any contaminating phase. Unit cell parameters were refined using the program LAPOD

(Langford, 1973), based on indices calculated using PowderCell (Kraus and Nolze,

1996a,b) and gave a = 6.878(1), b =7.159 (2) c =6.734(3) Å for rooseveltite, a = 9.993

(1), 7.404 (1) c = 6.937 (3) Å for preisingerite, and a = 7.000 (1), 7.430(1) c = 10.831

(2) Å for atelestite. These results are in excellent agreement with those reported

elsewhere (Bedlivy and Mereiter, 1982a, 1982b; Mereiter and Preisinger, 1986).

51 | P a g e
120

110

100

90

80
Lin (Counts)

70

60

50

40

30

20

10

9 10 20 30 40 50 60

2-Theta - Scale

Figure 3.4: Powder XRD trace of rooseveltite with peak positions of the ‘type’ pattern from JCPDS file 00-025-0089 (red).

52 | P a g e
280

270

260

250

240

230

220

210

200

190

180

170
Lin (Counts)

160

150

140

130

120

110

100

90

80

70

60

50

40

30

20

10

10 20 30 40 50

2-Theta - Scale

Figure 3.5: Powder XRD trace of preisingerite with peak positions of the ‘type’ pattern from JCPDS file 01-075-1629 (red).

53 | P a g e
120

110

100

90

80

70
Lin (Counts)

60

50

40

30

20

10

11 20 30 40 50 60 70

2-Theta - Scale

Figure 3.6: Powder XRD trace of atelestite with peak positions of the ‘type’ pattern from JCPDS file 00-025-0089 (red).

54 | P a g e
3.4.2 Solution studies

Solubility data for the three minerals are reported in Tables 3.1, 3.2 and 3.3. All

three phases are comparatively insoluble. Rooseveltite and preisingerite dissolve

congruently and atelestite dissolved incongruently in aqueous HNO3. The incongruent

dissolution of atelestite was accounted for with respect to preisingerite. The dissolution

of the minerals is expressed in equations (1), (2) and (3).

BiAsO4 (s) + 2H+(aq) ⇋ Bi3+(aq) + H2AsO4 (aq) (1)

Bi3(AsO4)2O(OH)(s) + 7H+(aq) ⇋ 3Bi3+(aq) + 2H2O(l) + 2HAsO4-(aq) (2)

Bi2(AsO4)O(OH)(s) + 6H+(aq) ⇋ 2Bi3+(aq) + 2H2O(l) + H3AsO40(aq) (3)

The pH at equilibrium (rooseveltite = 1.210, preisingerite =1.080 ; and atelestite

= 1.152,) was used in the determination of ion speciation patterns using the program

COMICS (Perrin and Sayce, 1967). Individual ion activity coefficients were calculated

using the Debye-Hückel equation for 298.15 K, lg  = –0.5085z2((I/(1+I)-0.3I). For

rooseveltite I = 0.098 mol.dm–3, 3 = 0.081, 2 = 0.327 and  = 0.756. For

preisingerite I = 0.090 mol.dm–3, 3 = 0.043, 2 = 0.246 and  = 0.704 and for

atelestite I = 0.102 mol.dm–3, 3 = 0.1079, 2 = 0.372 and  = 0.781. In all instances,

o is taken to be unity. The activity of H+(aq), a(H+), in HNO3 0.0981 M was then

calculated and corresponding activity a(Bi3+) was calculated from the solubility data

(Tables 3.1, 3.2 and 3.3).

55 | P a g e
Table 3.1 Dissolved metal concentrations for rooseveltite

Solution [Bi] ppm [Bi] mol dm-3 [As] ppm [As] mol dm-3
1 0.41 1.96 x 10-06 0.91 9.49 x 10-06
2 0.35 1.67 x 10-06 0.72 7.50 x 10-06
3 0.41 1.96 x 10-06 0.76 7.92 x 10-06
4 0.36 1.72 x 10-06 0.80 8.34 x 10-06
5 0.35 1.67 x 10-06 0.75 7.82 x 10-06
6 0.43 2.06 x 10-06 0.71 7.40 x 10-06
7 0.34 1.63 x 10-06 0.78 8.13 x 10-06
8 0.36 1.72 x 10-06 0.86 8.96 x 10-06
9 0.34 1.63 x 10-06 0.76 7.92 x 10-06
Mean 0.36 1.73 x 10-06 0.85 8.08 x 10-06
Error 0.03 1.52 x 10-07 0.05 5.56 x 10-07

Table 3.2. Dissolved metal concentrations for preisingerite

Solution [Bi] ppm [Bi] mol dm-3 [As] ppm [As] mol dm-3
1 0.59 2.8 x 10-06 1.8 2.4 x 10-05
2 0.52 2.5 x 10-06 2.1 2.8 x 10-05
3 0.58 2.8 x 10-06 1.9 2.5 x 10-05
4 0.61 2.9 x 10-06 1.7 2.3 x 10-05
5 0.37 1.8 x 10-06 3.1 4.1 x 10-05
Mean 0.46 2.2 x 10-06 2.1 2.8 x 10-05
Error 0.16 7.4 x 10-07 2.7 1.3 x 10-05

Table 3.3. Dissolved metal concentrations for atelestite

Solution [Bi] ppm [Bi] mol dm-3


1 150 7.2 x 10-04
2 160 7.7 x 10-04
3 170 8.1 x 10-04
4 180 8.6 x 10-04
5 160 7.7 x 10-04
6 150 7.2 x 10-04
Mean 162 7.7 x 10-04
Error 8.89 4.25 x 10-05

56 | P a g e
The pH at equilibrium for rooseveltite was used in the determination of

individual ion speciation using COMICS (Perrin and Sayce, 1967). Reliable Log K

values for equation (2), (4), (6), (8), (10), (12) and (14) of Log K(298.15 K) = –1.34, –

3.52, -8.44, +10.62, +17.75 and +19.48 at I = 0 mol dm–3 were reported by van Der Lee

and Lomenech (2004) and Wagman et al. (1982). Correction to I = 0.128, mol dm–3 by

the method of Baes and Mesmer (1976), yields lg K(298.15 K) = -1.54, –4.05, –9.07,

10.96, 17.29 and 19.52 Using the relationships shown by equations (5), (7), (9), (11),

(13), and (15) the concentrations of Bi(OH)2+(aq), Bi(OH)2+(aq), Bi(OH)3o(aq),

HAsO42(aq),- and H2AsO42-(aq) are 7.10 x 10-7, 7.60 x 10-8, 2.3 x 10-12, 1.9 x 10-6, 6.2 x

10-6 mol.dm–3 respectively.

Bi3+(aq) + H2O(l) ⇋ Bi(OH)2+(aq) + H+(aq) (4)

[Bi(OH)2+].[H+]/[Bi3+] = 10–1.11 (5)

Bi3+(aq) + 2H2O(l) ⇋ Bi(OH)2+(aq) + 2H+(aq) (6)

[Bi(OH)2+].2[H+]/[Bi3+] = 10–3.30 (7)

Bi3+(aq) + 3H2O(l) ⇋ Bi(OH)30(aq) + 3H+(aq) (8)

[Bi(OH)30].3[H+]/[Bi3+] = 10–8.21 (9)

AsO43-(aq) + H+(aq) ⇋ HAsO4- (aq) (10)

[HAsO4-]/[ AsO43-] = 10+11.60 (11)

AsO43-(aq) + 2H+(aq) ⇋ H2AsO42- (aq) (12)

[H2AsO42-]/[ AsO43-] = 10+18.36 (13)

AsO43-(aq) + 3H+(aq) ⇋ H3AsO40 (aq) (14)

[H3AsO42-]/[ AsO43-] = 10+20.81 (15)

57 | P a g e
Correction to I = 0.128, mol.dm–3 for preisingerite by the method of Baes and

Mesmer (1976), yields lg K(298.15 K) = –1.54, –3.94, –8.85, +10.96, 17.29 and 19.52.

Using the relationships shown by equations (5), (7), (9), (11), (13), and (15) the

concentration of Bi(OH)2+(aq), Bi(OH)2+(aq), Bi(OH)3o(aq), HAsO42(aq),- H2AsO42-(aq)

and H3AsO40(aq) are 6.60 x 10-8, 5.10 x 10-9 7.53 x 10-9, 1.20 x 10-6, 2.8 x 10-12 , and

4.10 x 10-18, mol dm–3 respectively.

Correction to I = 0.128, mol.dm–3 for atelestite by the method of Baes and

Mesmer (1976), yields lg K(298.15 K) = –1.54, –4.05, –9.07, 10.96, 17.29 and 19.52.

Using the relationships shown by equations (5), (7), (9), (11), (13), and (15) the

concentration of Bi(OH)2+(aq), Bi(OH)2+(aq), Bi(OH)3o(aq), HAsO42(aq),- H2AsO42-(aq)

and H3AsO40(aq) are 2.20 x 10-5, 7.20 x 10-7 and 1.30 x 10-10 mol.dm–3 respectively. the

concentration of Bi(OH)2+(aq), Bi(OH)2+(aq) and Bi(OH)3o(aq) are 2.20 x 10-5, 7.20 x

10-7 and 1.30 x 10-10 mol.dm–3 respectively.

This yields values of lg K for the dissolution of rooseveltite, preisingerite and

atelestite for equations (1) –9.65±0.50, (2) –41.43±0.50, and (3) –7.93±0.50

respectively and yields the corresponding ΔGrө and ΔGfө values for equations (1) to (3)

as follows:

(1) ΔGrө(298.15 K) = +55.08. Thus, ΔGfө (BiAsO4, 298.15 K) = –

716.91±1.5 kJ mol–1

(2) ΔGrө(298.15 K) = +7.26. Thus, ΔGfө (Bi3(AsO4)2O(OH), 298.15 K) =

–1823.29 ±2.5 kJ mol–11

(3) ΔGrө(298.15 K) = +45.27. Thus, ΔGfө Bi2(AsO4)O(OH), 298.15 K) =

–1102.58 ±5.6 kJ mol–1

58 | P a g e
The estimated error takes into account the analytical error of the solubility experiments

and errors quoted for the thermochemical data used.

Table 3.4 Thermodynamic quantities used in the calculations (T = 298.15 K).

ΔGfө/kJ mol-1 Reference


bismoclite –322.1 Kaye and Laby (1995)
bismutitea –916.2 ± 7.5 see text
rooseveltite –716.91 ± 0.5 this study
preisingerite –1823.29 ± 0.5 this study
atelestite –1102.58 ± 0.5 this study
H3AsO4o(aq) –766.75 Nordstrom and Archer (2003)
H2AsO4-(aq) –753.65 Nordstrom and Archer (2003)
HAsO42-(aq) –713.73 Nordstrom and Archer (2003)
AsO43-(aq)b –646.36 see text
Bi(OH)3o(aq)c –572.61 see text
Bi(OH)2+(aq)c –363.54 see text
Bi(OH)2+(aq)c –138.94 see text
Bi3+(aq)c +91.82 see text
CO2(g) –394 Robie and Hemingway (1995)
Cl-(aq) –131.2 ± 0.1 Robie and Hemingway (1995)
H2O(l) –237.1 ± 0.1 Cox et al. (1989)
a
Values have been recalculated using values for H2O(l), CO2(g) and Bi2O3(s) at 298.2 K
taken from Robie and Hemingway (1995). bCalculated using ΔGfө(AsO43-) from Barner
and Scheuerman (1978) and dissociation constants of arsenic acid from Martell and
Smith (1982). cCalculated using standard electrode potentials from Lovrecek et al.
(1985) to determine ΔGfө (Bi3+) and dissociation constants from van Der Lee and
Lomenech (2004).

59 | P a g e
3.5 REFERENCES

Alves, P., (2010) Rooseveltite: Bi(AsO4), Marcasite: FeS2, Lagoa Mine, Estorãos, Ponte

de Lima, Viana do Castelo District, Portugal, photograph, viewed 4 January 2013,

http://www.mindat.org/photo-319350.html.

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (2000) Handbook of

Mineralogy Volume 4: Arsenates, Phosphates, Vanadates. Mineral Data

Publishing, Tucson, Arizona.

Baes, C.F. Jr and Mesmer, R.E. (1976) The Hydrolysis of Cations. Plenum Press, New

York.

Barner, H. E. And Scheuerman, R. H. (1978) Handbook of thermochemical data for

compounds and aqueous species. John Wiley and Sons, New York.

Bedlivy, D., Llamblas, E.J. and Astarloa, J.F.H. (1972) Rooseveltit von San Francisco

de los Andes und Cerro Negro de la Aguadita, San Juan, Argentina. Tschermaks

Mineralogische und Petrographische Mitteilungen, 17, 65-75.

Bedlivy, D. and Mereiter, K. (1982a) Structure of α-BiAsO4 (rooseveltite). Acta

Crystallographica, 38, 1559-1561.

Bedlivy, D. and Mereiter, K. (1982b) Preisingerite, Bi3O(OH)(AsO4)2, a new species

from San Juan Province, Argentina: its description and crystal structure. American

Mineralogist, 67, 833-840.

60 | P a g e
Breithaupt, A. (1832) Vollständige Characteristik des Minerals-Systems, Second

edition. Arnoldische Buchhandlung, Dresden, p. 307.

Cox, J.D., Wagman, D.D. and Medvedev, V.A. (1989) CODATA Key Values for

Thermodynamics. Hemisphere Press, New York.

Dunn, P.J., Grice, J.D., Wicks, F.J. and Gault, R.A. (1988) Paulkellerite, a new bismuth

iron phosphate mineral from Schneeberg, Germany. American Mineralogist, 73,

870-872.

Herzenberg, R. (1946) Nuevos minerales de Bolivia. Boletín Técnico de la Técnico

Facultad Nacional Ingeniería, Universidad Técnica Oruro, 1, 10.

Kaye, G.W.C. and Laby, T.H. (1995) Tables of Physical and Chemical Constants. 16th

Edition, Longman, London, Section 3.10.5.

Kraus, W. and Nolze, G. (1996a) PowderCell - a program for the representation and

manipulation of crystal structures and calculation of the resulting X-ray powder

patterns. Journal of Applied Crystallography, 29, 301-303.

Kraus, W. and Nolze, G. (1996b) PowderCell 1.8 and Windows version 1.0. Federal

Institute for Materials Research and Testing (BAM), Berlin.

61 | P a g e
Langford, J.I. (1973) Least-squares refinement of cell dimensions from powder data by

Cohen's method. Journal of Applied Crystallography, 6, 190-196.

Lovreček, B., Mekjavić, I. and Metikoš-Huković, M. (1985) Bismuth. In: Bard, A.J.,

Parsons, R. and Jordan, J. (Eds.) Standard Potentials in Aqueous Solution.

International Union of Pure and Applied Chemistry and Marcel Dekker, New York,

180-187.

Markl, G. (1990) Mineralogie der Grube Anton bei Schiltach im Heubachtal,

Schwarzwald. Lapis, 15(5), 11-20, 50.

Martell , A. E. and Smith, R. M. (1982) Critical Stability Constants Volume 5: First

Supplement. Plenum Press, New York.

Meier, S. and Dünkel, B. (2010) Mineralien von der Dorschenmühle bei Lichtenberg,

Franken. Lapis, 35(10), 60-64, 90.

Mereiter, K. and Preisinger, A. (1986) Kristallstrukturdaten der Wismutminerale

Atelestit, Mixit und Pucherit. Anzeiger Österreichischen Akademie der

Wissenschaften mathematisch - naturwissenschaftlichen Klasse, 123, 79-81.

Mooney, R.C.L. (1948) Crystal structure of tetragonal bismuth arsenate. Acta

Crystallographica, 1, 163-165.

62 | P a g e
Nordstrom, D.K. and Archer, D.G. (2003) Arsenic thermodynamic data and

environmental geochemistry. In Welch, A.H. and Stollenwerk, K.G. (Eds) Arsenic

in Ground Water. Kluwer Academic Publishers, Boston, pp. 1-25.

Palache, C., Berman, H. and Frondel, C. (1951) Dana’s System of Mineralogy, seventh

edition. John Wiley and Sons, New York.

Perrin, D.D. and Sayce, I.G. (1967) Computer calculation of equilibrium concentrations

in mixtures of metal ions and complexing species. Talanta, 14, 833-842.

Rankin, J., Lawrence, L.J., Sharpe, J.L. and Williams, P.A. (2002) Rare secondary

bismuth, tungsten and molybdenum minerals from Elsmore, New England district

of New South Wales. Australian Journal of Mineralogy, 8, 55-60.

Rankin, J., Sharpe, J.L. and Williams, P.A. (2013) Unpublished survey of secondary Bi,

Mo and W minerals in Australian deposits.

Robie, R.A. and Hemingway, B.S. (1995) Thermodynamic properties of minerals and

related substances at 298.15K and 1 bar (105 Pascals) pressure and at higher

temperatures. United States Geological Survey Bulletin, 2131.

Schlegel, L., Kleeberg, R. and Meisser, N. (1996) Sekundäre Wismutminerale und

weitere Neufunde aus Schneeberg/Sachsen, 1992-1995. Lapis, 21(9), 37-41.

63 | P a g e
Sejkora, J. (1994) Minerály ložiska Moldava v Krušných horách. Bulletin

Mineralogicko-petrografického oddělení Národního muzea v Praze, 2, 110-116.

Sejkora, J. and Řídkošil, T. (1994) Tetrarooseveltite, Bi(AsO4), a new mineral species

from Moldava deposit, the Krušnéhory Mountains, northwestern Bohemia, Czech

Republic. Neues Jahrbuch für Mineralogie, Monatshefte, 179-184.

Valverde, J., (2006) Atelestite:Bi2(AsO4)O(OH) Hérival, Val-d'Ajol,

Vosges, Lorraine, France, photograph, viewed 4 January 2013,

http://www.mindat.org/photo-50845.html.

van der Lee, J. and Lomenech, C. (2004) Towards a common thermodynamic database

for speciation models. Radiochimica Acta, 92, 811-818.

Wagman, D. D., Evans, W. H., Parker, V. B, Schumm, R. H., Halow, I., Bailey, S.

M., Churney, K. L., and Nuttall, R. L. (1982) The NBS tables of chemical

thermodynamic properties: selected values for inorganic and C1 and C2 organic

substances in SI units. Journal of Physical and Chemical Reference Data, 11,

Supplement 2, 1-392.

Walenta, K., (1972) Die Sekundärmineralien der Co-Ni-Ag-Bi-U-Erzgänge im Gebiet

von Wittichen im mittleren Schwarzwald. Aufschluss, 23, 279-329.

Walenta, K., (1992) Die Mineralien des Schwarzwaldes. Weise, Munich

64 | P a g e
Walenta, K., (1995) Sphaerobismoite, a new mineral of the composition Bi2O3 from the

Black Forest. Der Aufschluss, 46, 245-248.

Weiss, S., Green, D.H., Hooper, J.J., and Elton, N. (2007) Erzgänge und “Alpine”

Klüfte: Die Mineralien des Hingston quarry, Calstock, Cornwall. Lapis, 32(11), 13-

25.

Wolfsried, S., (2007) Preisingerite:Bi3(AsO4)2O(OH) Clara Mine, Rankach valley,

Oberwolfach, Wolfach, Black Forest, Baden-Württemberg Germany, photograph,

viewed 4 January 2013, http://www.mindat.org/photo-102499.html.

65 | P a g e
CHAPTER 4

RUSSELLITE, KOECHLINITE AND SARDIGNAITE

66 | P a g e
4.1 ABSTRACT

Bismuth is concentrated in a wide variety of ore types, principally associated

with high-temperature acid intrusives, related skarn deposits and epithermal systems

(Angio and Long, 1979; Baker et al., 2005; Lang and Baker, 2001). It is, together with

Te, W, Sn and other accessory elements, often associated with gold mineralisation

(Angio and Long, 1979; Baker et al., 2005). The association of Bi minerals at the Mo-

Bi deposits of Kingsgate and surrounding deposits found in the New England Orogen,

New South Wales, commonly give rise to koechlinite, Bi2MoO6, and russellite,

Bi2WO6, in the oxidised zone (Rankin et al., 2002; Sharpe and Williams, 2004;

Clissold, 2007). The Gibbs free energy of formation values for russellite, koechlinite,

and sardignaite, BiMo2O7(OH)∙2H2O, are derived in this chapter ΔGfө (298.15 K) = –

1318.93±1.5 kJ mol–1, –1270.78±1.5 kJ mol–1 and –2176.94±1.56 kJ mol–1 respectively

and are used to determine the roles they play in the immobilisation of bismuth in the

oxidised environment.

4.2 INTRODUCTION

Molybdenum possesses exotic supergene mineralogy. Thirty different species

are currently recognised by the IMA and most of them are rare mineral phases (Anthony

et al., 1990, 1995, 1997, 2000, 2003; Gaines et al., 1997). The molybdate salts

powellite, CaMoO4, and wulfenite, PbMoO4, are the most common, and have been both

reported in small amounts from Kingsgate (England, 1985). Basic Cu(II) molybdates

occur in certain porphyry Mo deposits and several uranyl molybdates are associated

with oxidised, U-rich ores. No carbonates are known and the simple oxides molybdite,

MoO3, and sidwillite, MoO3·2H2O, are very rare. This study has recognised that much

material identified in the past as molybdite, or ferrimolybdite, Fe2(MoO4)·8H2O is in

67 | P a g e
fact more often koechlinite Bi2MoO6. Ferrimolybdite can only form at pH values less

than about 3 and is thus characteristic of acid oxidising conditions (Sarafian and

Furbish, 1965). Between about pH 3 and 3.5, the more basic phase bamfordite,

FeMo2O6(OH)3·H2O, is formed but the narrow pH range associated with its stability

(Sarafian and Furbish, 1965) limits its distribution in nature; to date bamfordite is

known only from its type locality, the Bamford Hill W-Mo-Bi deposit, Bamford,

Tablelands Region, Queensland, Australia (Birch et al., 1998). These observations,

together with knowledge of the occurrence of polymolybdate species, are important

geochemical clues with respect to the dispersion of Mo in the supergene zone.

Like tungstate (VI), molybdate, MoO42-, polymerises under acid conditions if

sufficiently high concentrations are reached (Baes and Mesmer, 1976) and molybdate

can be dispersed under basic conditions in ground waters. However, when molybdenite

oxidises together with bismuth or tungsten minerals, the very insoluble Aurivillius

phases koechlinite, Bi2MoO6, and russellite, Bi2WO6, are formed. These are very

common secondary minerals in the New England Bi deposits. It has hitherto been

supposed (England, 1985) that ferrimolybdite is the common oxidation product of

molybdenite, MoS2, at Kingsgate, but this is not the case; most of the yellow “molybdic

ochres” have been shown to be koechlinite, Bi2MoO6. This too is significant in

understanding the geochemistry of the oxidised zone in this setting and its bearing on

the dispersion of Bi and Mo under weathering conditions.

In general, the supergene zone at Kingsgate is acidic. The Bi-Mo deposits of

Kingsgate are virtually devoid of carbonates, aside from the very insoluble Bi species

bismutite. The sole pipe reported to carry primary carbonate mineralisation is

68 | P a g e
Goodwin’s Pipe, in which calcite formed the cement for fractured quartz and occurred

as crystals a few cm long lining vughs (England, 1985). Drusy films of calcite on or

filling joints between quartz crystals and fragments may still be collected on the dumps.

No other carbonate minerals have been found elsewhere on the field. Given the above

data, and the geochemical considerations below, the occurrence of bismite at Kingsgate,

and elsewhere, must be due to its crystallisation in local micro-environments that were

CO2-depleted; alternatively, it may form as a metastable phase. In any event, bismite is

a rare mineral at Kingsgate (England, 1985; Sharpe and Williams, 2004). Oxidation of

molybdenite in a quartz-dominated gangue free of carbonate inevitably leads to very

acidic conditions. This is evident from an inspection of the stoichiometry of the process

as shown in equation (1).

MoS2(s) + 4O2(g) + 4H2O(l) → MoO42-(aq) + 2SO42-(aq) + 8H+(aq) (1)

Oxidation of accessory pyrrhotite and arsenopyrite, in particular, followed by

hydrolysis of Fe3+(aq), will further acidify the system. In the absence of a carbonate

gangue, protons will either be flushed from the supergene zone or react with feldspars to

yield clays, which are conspicuous constituents of the upper sections of the Kingsgate

pipes. In related deposits, such as at Elsmore, acid-catalysed polymerisation of MoO42-

gives rise to Ca-Mg betpakdalite, [Mg(H2O)6]Ca2(H2O)13[Mo8As2Fe3O36(OH)](H2O)4

(Rankin et al., 2001, 2002). At Kingsgate, this species has not been observed, but

conditions were certainly acid enough to give rise to ferrimolybdite in the oxidised

zones of Bi-poor pipes.

69 | P a g e
The structure of koechlinite, Bi2MoO6, was originally reported by Zemann

(1956). This was subsequently refined by van den Elzen and Rieck (1973) and Teller et

al. (1984). The type specimen was from St Daniel mine, Neustädtel, Schneeberg

District, Saxony, Germany (Schaller, 1916), where it was discovered in a quartz vein

with bismuth and smaltite. Associated primary minerals include bismuthinite and

secondary minerals include bismutite, preisingerite and neustädtelite (Wittern, 2001).

Another report on koechlinite (Yang et al., 1989), coated by chiluite occurs at an

unnamed molybdenite deposit at Chilu, Fujian, China. Molybdenite, bismuthinite,

joseite and cassiterite are present in the associated primary assemblage. In an Australian

context, koechlinite is also common in various sites of the New England region, NSW,

Australia as mentioned in Chapter 1.

Figure 4.1: Koechlinite completely replacing bismuthinite needles, FOV 10mm. From
Quartz vein outcrops, Knöttel area (Knötel; Knödel; Knödlberg), Krupka

70 | P a g e
(Graupen), Krušné Hory Mts (Erzgebirge), Ústí Region, Bohemia (Böhmen;
Boehmen), Czech Republic (Fuchs, 2012).

The structure of sardignaite, BiMo2O7(OH)·2H2O, was reported by Orlandi et al.

(2010) on the type specimen from Punta de Su Seinargiu, Sarroch, Cagliari Province,

Sardinia, Italy (Orlandi et al., 2010), at this site bismuth and molybdenite are primary

phases; and sardignaite was associated with gelosaite. Another reported occurrence is

from Elsmore Hill, Elsmore, associated with a host of secondary Bi and Mo minerals,

particularly koechlinite (Rankin et al., 2013, unpublished data).

Figure 4.2: Elongated tabular crystals of sardignaite on quartz, with colourless fluorite
group size crystals 1mm. from Punta de Su Seinargiu, Sarroch, Cagliari Province,
Sardinia, Italy (Ambrino, 2013).

Gelosaite, BiMo6+(2-5x)Mo5+6xO7(OH)·H2O, was first reported by Orlandi et al

(2011) along with its structure. The type locality was the Punta de Su Seinargiu,

71 | P a g e
Sarroch, Cagliari Province, Sardinia, Italy (Orlandi et al., 2011), where bismuth and

molybdenite are primary phases and it was associated with sardignaite. In the same

report (Orlandi et al., 2011) another occurrence is mentioned from the number 25 Pipe,

Kingsgate, New South Wales, Australia where it was associated with a host of

secondary Mo and Bi minerals, and intimate associates are bismutite and bismite, but it

usually occurs alone in vughs in quartz (Orlandi et al., 2011).

Figure 4.3: Cluster of gelosaite crystals 2mm across. Colour varies from greyish green
to deep blue-green. from Old 25 Pipe, Kingsgate, Gough Co., New South Wales,
Australia (Haupt, 2011)
The structure of russellite, Bi2WO6, (Figure 4.4) was re-determined from a

neutron powder diffraction study of synthetic Bi2WO6 and reported by Knight (1992).

The type locality is Castle-an-Dinas mine, Castle-an-Dinas, St Columb Major,

Cornwall, UK, where wolframite is listed as a primary phase and chrysocolla is the only

reported secondary phase (Hey and Bannister, 1938) .

72 | P a g e
Figure 4.4: Yellow microcrystalline masses of russellite in cavities in quartz vein.
Specimen size is 9.5x6x3 cm from Kara-Oba W deposit, Betpakdal Desert (Bet-Pak-Dal
Desert), Karagandy Province (Qaragandy Oblysy; Karaganda Oblast'), Kazakhstan
(Pavel, 2008)

4.3 EXPERIMENTAL

4.3.1 Syntheses of russellite, koechlinite and sardignaite

Russellite was obtained as follows. A mixture of Bi(NO3)3∙5H2O (1.0045 g,

ACS reagent ≥ 98%), Na2WO4∙2H2O (0.3564 g, ACS reagent ≥ 98%) and HNO3 (10

cm3, 0.1 M, ACS reagent 70%) was reacted in an acid digestion bomb at 180oC for 48

hours. The product obtained was a pale yellow, crystalline powder. The product was

suspended in aqueous 0.1 M HNO3 at room temperature for 24 hours to dissolve any

residual starting materials. It was then collected at the pump, washed with water then

acetone, and vacuum dried.

73 | P a g e
Koechlinite was synthesised in a similar way. A mixture of Bi(NO3)3∙5H2O

(1.0054 g, ACS reagent ≥ 98%), (NH4)6Mo7O24∙4H2O (0.18274 g, ACS reagent ≥ 98%)

and HNO3 (10 cm3, 0.1 M, ACS reagent 70%) was reacted in an acid digestion bomb at

180oC for 48 hours. The colour of the crystalline powder obtained was off-white to

yellow, and was isolated as described above.

Sardignaite was prepared by mixing Bi(NO3)3∙5H2O (0.5001 g),

(NH4)6Mo7O24∙4H2O (0.1999 g,) and HNO3 (20 cm3, 0.1 M, ACS reagent 70%). The

mixture was placed in an acid digestion bomb at 180oC for 48 hours. The product

obtained was off-white in colour, and a crystalline powder was isolated as described

above.

Powder X-ray diffraction studies of the products were carried-out using a Philips

PW1825/20 powder diffractometer (Ni-filtered CuKα1 radiation, λ = 1.5406 Å, 40 kV,

30 mA). Traces were produced between 5-70o 2θ, with a step size of 0.02 o and a rate of

1.2 o min–1. Diffraction Technology Data processing software (Traces Version 6) and

JCPDS-ICDD data base files were used to identify the phases produced.

4.3.2 Solubility studies

Solubility studies were undertaken using sealed 250 cm3 conical QuickfitR flasks

maintained at 25.0 0.2oC in a thermostatted water bath inside a controlled temperature

room at the same temperature. Measurements of pH were made using a Radiometer

ION450 apparatus fitted with a combination electrode. Subsequent analyses showed that

all three minerals dissolve congruently in aqueous HNO3. The minerals (ca 0.1 g) were

added to a series of conical flasks containing 100 cm3 of standardised HNO3; 0.09928

M for koechlinite and russellite and 0.0981 M for sardignaite. The flasks were left for

74 | P a g e
10 weeks, during which time the pH of paired flasks was monitored periodically until

no change was detected. In all cases, equilibrium was achieved after 6 weeks. Resulting

solutions were filtered through a WhatmanR GF/F fibreglass filter and collected in clean

PET bottles. Dissolved Bi/Mo and Bi/W concentrations were determined using ICP-MS

by a NATA-compliant commercial laboratory (mgt | LabMark Environmental Testing

Australia Pty Ltd).

4.4 RESULTS

4.4.1 Syntheses of russellite, koechlinite and sardignaite

High purity, samples of russellite, koechlinite and sardignaite were obtained in

essentially quantitative yield. Powder X-ray diffraction did not give any evidence for

any contaminating phase. Unit cell parameters were refined using the program LAPOD

(Langford, 1973) and gave a = 5.439(2), b = 16.422(2), c = 5.45338) Å for russellite, a

= 5.493(3), b = 16.200(6) c = 5.507(3) Å for koechlinite, and a = 5.794(7), b = 11.56(2)

c = 6.357(7) Å for sardignaite. These results are in excellent agreement with those

reported elsewhere (Teller et al., 1984; Knight, 1992; and Orlandi et al., 2010).

75 | P a g e
140

130

120

110

100

90
Lin (Counts)

80

70

60

50

40

30

20

10

11 20 30 40 50 60

2-Theta - Scale

Figure 4.5: Powder XRD trace of russellite with peak positions of the ‘type’ pattern from JCPDS file 00-039-0256 (red).

76 | P a g e
220

210

200

190

180

170

160

150

140

130
Lin (Counts)

120

110

100

90

80

70

60

50

40

30

20

10

5 10 20 30 40 50 60 70

2-Theta - Scale

Figure 4.6: Powder XRD trace of koechlinite with peak positions of the ‘type’ pattern from JCPDS file 01-072-1524 (red).

77 | P a g e
160

150

140

130

120

110

100
Lin (Counts)

90

80

70

60

50

40

30

20

10

14 20 30 40 50

2-Theta - Scale

Figure 4.7: Powder XRD trace of sardignaite with peak positions of the ‘type’ pattern from JCPDS file 01-072-1524 (red).

78 | P a g e
4.4.2 Solution studies

Solubility data for the three minerals are reported in Tables 4.1, 4.2, and 4.3. All

three phases are comparatively insoluble and dissolve congruently in aqueous HNO3.

Bi2WO6(s) + 4H+(aq) ⇋ 2Bi3+(aq) + WO42-(aq) +2H2O(l) (2)

Bi2MoO6(s) + 4H+(aq) ⇋ 2Bi3+(aq) + MoO42-(aq) +2H2O(l) (3)

BiMo2O7(OH)∙2H2O(s) + H2O ⇋ Bi3+(aq) + 2MoO42-(aq) + 3H2O + H+(aq) (4)

The pH at equilibrium (1.082, russellite; 1.071 koechlinite; 1.090, sardignaite)

was used in the determination of ion speciation patterns using the program COMICS

(Perrin and Sayce, 1967). Individual ion activity coefficients were calculated using the

Davis extension of the Debye-Hückel equation for 298.15 K, lg  = –

0.5085z2((I/(1+I)-0.3I), for russellite I = 0.090 mol.dm–3, 3 = 0.042 and 2 = 0.245,

for koechlinite I = 0.092 mol.dm–3, 3 = 0.041 and 2 = 0.242, and for sardignaite I =

0.090 mol.dm–3, 3 = 0.043 and 2 = 0.2460, In all instances, o is taken to be unity.

The activity of H+(aq), a(H+), in HNO3; 0.0981 M was then calculated and

corresponding activities a(Bi3+), a(WO42-) and a (MoO42-), were calculated from the

solubility data (Tables 4.1, 4.2, and 4.3).

79 | P a g e
Table 4.1 Dissolved metal concentrations for russellite

Solution [Bi] ppm [Bi] mol dm-3


1 5.6 2.7 x 10-5

2 5.6 2.7 x 10-5

3 5.2 2.5 x 10-5

4 5.9 2.8 x 10-5

5 5.8 2.8 x 10-5

6 5.1 2.4 x 10-5

Mean 5.5 2.6 x 10-5

Error 0.30 1.4 x 10-6

Table 4.2. Dissolved metal concentrations for koechlinite

Solution [Bi] ppm [Bi] mol dm-3 [Mo] ppm [Mo] mol dm-3
1 10 4.8 x 10-5 1.5 1.6 x 10-5

2 10 4.8 x 10-5 1.5 1.6 x 10-5

3 13 6.2 x 10-5 1.6 1.7 x 10-5

4 10 4.8 x 10-5 1.6 1.7 x 10-5

5 15 7.2 x 10-5 1.7 1.8 x 10-5

Mean 12 5.7 x 10-5 1.6 1.7 x 10-5

Error 2.0 9.6 x 10-5 0.1 5.2 x 10-7

Table 4.3. Dissolved metal concentrations for sardignaite

Solution [Bi] ppm [Bi] mol dm-3 [Mo] ppm [Mo] mol dm-3
1 0.59 2.8 x 10-6 1.8 2.4 x 10-5
2 0.52 2.5 x 10-6 2.1 2.8 x 10-5
3 0.58 2.8 x 10-6 1.9 2.5 x 10-5
4 0.61 2.9 x 10-6 1.7 2.3 x 10-5
Mean 0.58 2.75 x 10-6 1.9 2.5 x 10-5
Error 0.16 7.4 x 10-7 0.2 3.0 x 10-6

80 | P a g e
The pH at equilibrium for russellite was used in the determination of individual

ion speciation using COMICS (Perrin and Sayce, 1967). Reliable lg K values for the

equations (6), (8), (10), (12), (14), (16), and (18) at I = 0 mol dm–3 were reported by van

Der Lee and Lomenech (2004) and Baes and Mesmer (1976). Correction to I = 0.090,

mol dm–3 yields lg K(298.15 K) = –2.63, –5.74, –10.97, –11.42, and –7.12, using the

relationships shown by equations (7), (9), (11), (13), and (15). This gave the

concentration of Bi(OH)2+(aq), Bi(OH)2+(aq), Bi(OH)3o(aq), HWO4-(aq), and H2WO4-

(aq) as 3.7 x 10-8, 3.5 x 10-10 2.6 x 10-14, 2.00 x 10-19, and 3.30 x 10-16 mol.dm–3

respectively.

Bi3+(aq) + H2O(l) ⇋ Bi(OH)2+(aq) + H+(aq) (6)

[Bi(OH)2+].[H+]/[Bi3+] = 10–1.11 (7)

Bi3+(aq) + 2H2O(l) ⇋ Bi(OH)2+(aq) + 2H+(aq) (8)

[Bi(OH)2+].2[H+]/[Bi3+] = 10–3.30 (9)

Bi3+(aq) + 3H2O(l) ⇋ Bi(OH)30(aq) + 3H+(aq) (10)

[Bi(OH)30].3[H+]/[Bi3+] = 10–8.21 (11)

WO42-(aq) + H+(aq) ⇋ HWO4- (aq) (12)

[HWO4-]/[ WO42-][H+] = 10+3.5 (13)

WO42-(aq) + 2H+(aq) ⇋ H2WO4- (aq) (14)

[H2 HWO4--]/[ WO42--]2[H+] = 10+8.1 (15)

MoO42-(aq) + H+(aq) ⇋ HMoO4- (aq) (16)

[H MoO4-]/[ MoO42-][H+] = 10+3.89 (17)

81 | P a g e
MoO42-(aq) + 2H+(aq) ⇋ H2MoO4- (aq) (18)

[H MoO4-]/[ MoO42-][H+] = 10+7.50 (19)

Correction to I = 0.092, mol dm–3 for koechlinite by the method of Baes and

Mesmer (1976), yields lg K(298.15 K) = –1.63, –4.13, –9.14, –9.3 and –7.56, using the

relationships shown by equations (7), (9), (11), (13), and (15). This gave the

concentration of Bi(OH)2+(aq), Bi(OH)2+(aq), Bi(OH)3o(aq), HWO4-(aq), and H2MoO4-

(aq) as 5.10 x 10-7, 1.9 x 10-8, 2.2 x 10--12, 6.10 x 10-17, and 2.90 x 10-16 mol.dm–3

respectively.

Correction to I = 0.090, mol dm–3 for sardignaite by the method of Baes and

Mesmer (1976), yields lg K(298.15 K) = –1.87, –4.52, –9.58, –9.45 and –7.76, using the

relationships shown by equations (7), (9), (11), (13), and (15). This gave the

concentration of Bi(OH)2+(aq), Bi(OH)2+(aq), Bi(OH)3o(aq), HWO4-(aq), and H2MoO4-

(aq) as 3.00 x 10-7, 8.10 x 10-9, 8.7 x 10--13, 7.50 x 10-17, and 3.00 x 10-16 mol.dm–3

respectively.

The above then yield values of lg K for russellite, –16.99±1.5, koechlinite –

24.83±1.5 and sardignaite–20.59±1.50, equations (2), (3) and (4) respectively and the

following ΔGfө values can be calculated:

ΔGrө(298.15 K) = +55.08. Thus, ΔGfө (Bi2WO6, 298.15 K) = –1318.93 ±1.5 kJ mol–1

ΔGrө(298.15 K) = +7.26. Thus, ΔGfө (Bi2MoO6,), 298.15 K) = –1270.78±1.5 kJ mol–1

ΔGrө(298.15 K) = +45.27. Thus, ΔGfөB2MoO7(OH)·2H2O, 298.15 K) = –2176.94 ±1.56

kJ mol–1

82 | P a g e
The estimated error takes into account the analytical error of the solubility

experiments and errors quoted for the thermochemical data used.

Table 4.4 Thermodynamic quantities used in the calculations (T = 298.15 K).

ΔGfө /kJ.mol-1 Reference


bismoclite –322.1 Kaye and Laby (1995)
bismutitea –916.2 ±7.5 see text
russellite –1318.93 ±1.5 this study
koechlinite –1270.78 ±1.5 this study
sardignaite –2176.94 ±1.56 this study
H2WO4o(aq)b –964.51 see text
HWO4- (aq)b –951.95 see text
WO42-(aq)b –931.4 see text
H2MoO4o(aq)b –877.64 see text
HMoO4-(aq)b –867.10 see text
MoO42-(aq)b –838.50 see text
Bi(OH)3o(aq)c –572.61 see text
Bi(OH)2+(aq)c –363.54 see text
Bi(OH)2+(aq)C –138.94 see text
Bi3+(aq)c +91.82 see text
CO2(g) –394 Robie and Hemingway (1995)
Cl-(aq) –131.2±0.1 Robie and Hemingway (1995)
H2O(l) –237.1±0.1 Cox et al. (1989)
a
Values have been recalculated using values for H2O(l), CO2(g) and Bi2O3(s) at 298.2 K
taken from Robie and Hemingway (1995).bCalculated from Baes and Mesmer (1976)
and Martell and Smith (1982). cCalculated using standard electrode potentials from
Lovrecek et al. (1985) to determine ΔGfө (Bi3+) and dissociation constants from van Der
Lee and Lomenech (2004).

83 | P a g e
4.5 REFERENCES

Ambrino, P., (2013) Sardignaite: BiMo2O7(OH)·2H2O Punta de Su Seinargiu, Sarroch,

Cagliari Province, Sardinia, Italy, photograph, viewed 10 January 2014, <

http://www.mindat.org/photo-566105.html>.

Angio, E.E. and Long, D.T. (Eds) (1979) Geochemistry of Bismuth. Dowden,

Hutchinson and Ross, Inc., Stroudsburg, PA, USA.

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (1990) Handbook of

Mineralogy Volume 1: Elements, Sulfides, Sulfosalts. Mineral Data Publishing,

Tucson, Arizona.

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (1995) Handbook of

Mineralogy Volume 2: Silica, Silicates. Mineral Data publishing, Tucson, Arizona.

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (1997) Handbook of

Mineralogy Volume 3: Halides, Hydroxides, Oxides. Mineral Data publishing,

Tucson, Arizona.

Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (2000) Handbook of

Mineralogy Volume 4: Arsenates, Phosphates, Vanadates. Mineral Data

Publishing, Tucson, Arizona.

84 | P a g e
Anthony, J.W., Bideaux, R.A., Bladh, K.W. and Nichols, M.C. (2003) Handbook of

Mineralogy Volume 5: Borates, Carbonates, Sulfates. Mineral Data Publishing,

Tucson, Arizona.

Baes, C.F. Jr and Mesmer, R.E. (1976) The Hydrolysis of Cations. Plenum Press, New

York.

Baker, T., Pollard, P.J., Mustard, R., Mark, G. and Graham, J.L. (2005) A comparison

of granite-related tin, tungsten and gold–bismuth deposits: implications for

exploration. Society of Exploration Geologists Newsletter, 61, 5–17.

Birch, W.D., Pring, A., McBriar, E.M. Gatehouse, B.M. and McCammon, C.A. (1998)

Bamfordite, Fe3+Mo2O6(OH)3·H2O, a new hydrated iron molybdenum

oxyhydroxide from Queensland, Australia: description and crystal chemistry.

American Mineralogist, 83, 172-177.

Clissold, M.E. (2007) Aspects of the supergene geochemistry of copper nickel and

bismuth. PhD thesis unpublished, University of Western Sydney.

Cox, J.D., Wagman, D.D. and Medvedev, V.A. (1989) CODATA Key Values for
Thermodynamics. Hemisphere Press, New York.

England, B.M. (1985) Famous mineral localities: the Kingsgate mines. The

Mineralogical Record, 16, 265-289.

Fuchs, P., (2012) Koechlinite: Bi2MoO6 Quartz vein outcrops, Knöttel area (Knötel;

Knödel; Knödlberg), Krupka (Graupen), Krušné Hory Mts (Erzgebirge), Ústí

85 | P a g e
Region, Bohemia (Böhmen; Boehmen), Czech Republic, photograph, viewed 10

January 2013, < http://www.mindat.org/photo-477077.htmll>.

Gaines, R.V., Skinner, H.C.W., Foord, E.E., Mason, B. and Rosenzweig, A. (1997)

Dana’s New Mineralogy. Eight edition, John Wiley and Sons, New York.

Haupt, J., (2011) Gelosaite: BiMo6+(2-5x)Mo5+6xO7(OH)·H2O (0 ≤ x ≤ 0.4) Old 25 Pipe,

Kingsgate, Gough Co., New South Wales, Australia, photograph, viewed 10

January 2013, < http://www.mindat.org/photo-383852.html>.

Hey, M.H. and Bannister, F.A. (1938) Russellite, a new British mineral, with a note on

the occurrence and the accompanying minerals by Arthur Russell. Mineralogical

Magazine, 25, 41-56.

Knight, K.S. (1992) The crystal structure of russellite; a re-determination using neutron

powder diffraction of synthetic Bi2WO6. Mineralogical Magazine, 56, 399-409.

Lang, J.R. and Baker, T. 2001: Intrusion-related gold systems: the present level of

understanding. Mineralium Deposita, 36, 477-489.

Lovreček, B., Mekjavić, I. and Metikoš-Huković, M. (1985) Bismuth. In: Bard, A.J.,

Parsons, R. and Jordan, J. (Eds) Standard Potentials in Aqueous Solution.

International Union of Pure and Applied Chemistry and Marcel Dekker, New York,

180-187.

86 | P a g e
Martell , A. E. And Smith, R. M. (1982) Critical Stability Constants Volume 5: First

Supplement. Plenum Press, New York.

Orlandi, P., Demartin, F., Pasero, M., Leverett, P., Williams, P.A. and Hibbs, D.E.

(2011) Gelosaite, BiMo6+(2-5x)Mo5+6xO7(OH)·H2O (0 ≤ x ≤ 0.4), a new mineral from

Su Senargiu (Ca), Sardinia, Italy, and a second occurrence from Kingsgate, New

England, Australia. American Mineralogist, 96, 268-273.

Pavel, M.., (2008) Russellite: Bi2WO6 Kara-Oba W deposit, Betpakdala Desert,

Karagandy Province, Kazakhstan, photograph, viewed 10 January 2013, <

http://www.mindat.org/photo-148658.html>.

Perrin, D.D. and Sayce, I.G. (1967) Computer calculation of equilibrium concentrations

in mixtures of metal ions and complexing species. Talanta, 14, 833-842.

Rankin, J., Lawrence, L.J., Sharpe, J.L. and Williams, P.A. (2002) Rare secondary

bismuth, tungsten and molybdenum minerals from Elsmore, New England district

of New South Wales. Australian Journal of Mineralogy, 8, 55-60.

Rankin, J., Sharpe, J.L. and Williams, P.A. (2001) Betpakdalite from the tin deposits of

Elsmore, New England district of New South Wales. Australian Journal of

Mineralogy, 7, 15-17.

Rankin, J., Sharpe, J.L. and Williams, P.A. (2013) Unpublished survey of secondary Bi,

Mo and W minerals in Australian deposits.

87 | P a g e
Robie, R.A. and Hemingway, B.S. (1995) Thermodynamic properties of minerals and

related substances at 298.15K and 1 bar (105 Pascals) pressure and at higher

temperatures. United States Geological Survey Bulletin, 2131.

Sharpe, J.L. and Williams (2004) Secondary bismuth and molybdenum minerals from

Kingsgate, New England district of New South Wales. Australian Journal of

Mineralogy, 10, 7-12

Sarafian, P.G. and Furbish, W.J. (1965) Solubilities of natural and synthetic

ferrimolybdite. American Mineralogist, 50, 223-226.

Schaller, W.T. (1916) Koechlinite (bismuth molybdate), a new mineral. United States

Geological Survey Bulletin, 610, 10-34.

Teller, R.G., Brazdil, J.F., Grasselli, R.K. and Jorgensen, J.D. (1984) The structure of

bismuth molybdate, Bi2MoO6, by powder neutron diffraction. Acta

Crystallographica, C40, 2001-2005.

van den Elzen, A.F. and Rieck, G.D. (1973) Redetermination of the structure of

Bi2MoO6, koechlinite. Acta Crystallographica, B29, 2436-2438.

van der Lee, J. and Lomenech, C. (2004) Towards a common thermodynamic database
for speciation models. Radiochimica Acta, 92, 811-818.

Wittern, A. (2001) Mineralfundorte in Deutschland. Schweizerbart, Stuttgart.

88 | P a g e
Zemann, J. (1956) Die Kristallstruktur von Koechlinit, Bi2MoO6. Contributions to

Mineralogy and Petrology, 5, 139-145.

89 | P a g e
CHAPTER 5

SMIRNITE AND CHEKHOVICHITE

90 | P a g e
5.1 ABSTRACT

The Secondary Bi-Te minerals chekhovichite, Bi2Te4O11, and smirnite, Bi2TeO5,

were investigated to determine their role as path finder minerals in exploration

geochemistry due to the association of Bi-Te with precious metal deposits (McPhail,

1995). Syntheses and stability studies of the two minerals have been undertaken.

Solubilities in aqueous HNO3 were determined at 298.2 K and the data obtained used to

calculate values of ΔGfө at the same temperature. The derived ΔGfө (298.2 K) values for

chekhovichite, (–1457.23 ± 0.5 kJ mol–1), and smirnite (–820.92 ± 1.0 kJ mol–1) have

been used in subsequent calculations to determine their relative stabilities to each other

and their dispersion in the supergene zone.

5.2 INTRODUCTION

The occurrence of native Te is observed in some deposits, but Te most

commonly occurs in precious metal and sulfosalt minerals (Afifi et al., 1988; Jaireth,

1991). Of the known secondary species containing essential Bi and Te (Chapter 2, Table

1), the only phases that have any reasonable distribution include chekhovichite,

Bi2Te4O11, smirnite Bi2TeO5, and to a lesser extent pingguite Bi6Te2O13. These three

minerals form by oxidation of Bi-Te sulfosalts, which are common accessories in a

number of deposits. Chekhovichite was originally reported by Spiridonov et al. (1987)

where it was found at the Zod Mine (Sotk deposit), Vardenis, Gelark’unik’, Armenia,

the Northern Aksu deposit, Stepnogorsk, Akmolinskaya Oblast’, Kazakhstan and

Zhana-Tyube Au deposit, Akmolinskaya Oblast’, Kazakhstan. Spiridonov et al., (1987)

reported that chekhovicihite was associated with a number of Bi and Te sulfosalts and

known secondary phases, tetradymite, tellurobismuthite, plumbotellurite, PbTeO3,

91 | P a g e
Figure 5.1: Chekhovichite from the Zod Mine (Sotk deposit), Vardenis, Gelark'unik'
Province (Geghark'unik'), Specimen size 3.5 x 3.5 x 3 mm (Bosi, 2012)

“emmonsite” and tripuhyite. Rossell et al. (1992) have reported a structure for

chekhovichite, and a synthesis was reported by Szaller et al. (1996). The crystal

structure of smirnite, Bi2TeO5, was reported in Mercurio et al. (1983) and was found

from the same type locality as chekhovichite (Spiridonov et al., 1987). Smirnite is

associated with tellurobismuthite, tetradymite and volynskite and also has a known

synthesis (Ok, 2001).

92 | P a g e
Figure 5.2: Greyish to olive-green crusts of smirnite over quartz and metallic-grey
tsumoite from the Shilovo-Isetskii mine, Ekaterinburg, Sverdlovskaya Oblast', Middle
Urals, Urals Region, Russia. FOV is ~12x10 mm (Pavel, 2007).

There are other known bismuth tellurite minerals. The mineral “montanite”,

Bi2TeO6·2H2O or Bi2(OH)4TeO4, is listed by the IMA as doubtful despite 19

occurrences to date as its structure is yet to be confirmed (Palache et al., 1951;

Kazachenko et al., 1980). Another species which contains essential molybdenum is the

mineral chiluite, Bi6Te2Mo2O21, discovered from the Chilu deposit, Fu’an County,

Ningde Prefecture, China (Yang et al., 1989; Deng and Yang, 1993). The mineral is rare

and occurs as a replacement of joséite inclusions in bismuthinite and as coatings on

93 | P a g e
koechlinite, with molybdenite reported to be present in the primary assemblage. Despite

a known type locality, the mineral’s structure is uncertain. The synthetic analogue was

prepared, but the mineral requires further study as the cell constants are not well-

defined. The mineral was originally supposed to be hexagonal (Yang et al., 1989), but

subsequent work on the synthetic analogue shows it to be orthorhombic (Deng and

Yang, 1993).

5.3 EXPERIMENTAL

5.3.1 Syntheses of smirnite and chekhovichite

The synthesis of chekhovichite, Bi2Te4O11, was adapted from the synthesis

reported by Szaller et al. (1996). Bi2O3 and TeO3 were both initially annealed in air

(450oC) and then ground and sieved (-63 μm), the smaller fraction was used. A Bi2O3:

4TeO3 mixture was sealed in a quartz tube and heated at 720oC for 2 hours. The charge

was cooled slowly (50oC/hour) to ambient temperature then heated again (500oC) for

two days.

The synthesis of smirnite, Bi2TeO5, was adapted from the method reported by

Ok et al. (2001). A stoichiometric mixture of Bi2O3 (1.72 mmol) and TeO2 (1.72 mmol)

was thoroughly ground and sealed in a quartz tube and placed in a furnace (690oC) for

36 hours. The mixture was cooled to ambient temperature and the product collected as

above.

94 | P a g e
5.3.2 Solubility Studies

Solubility studies of the Te minerals were undertaken using sealed 250 cm3

conical QuickfitR flasks maintained at 25.0  0.2oC in a thermostatted water bath inside

a controlled temperature room at the same temperature. Measurements of pH were made

using a Radiometer ION450 apparatus fitted with a combination electrode. The minerals

(ca 0.1 g) were added to a series of conical flasks containing 100 cm3 of standardised

HNO3; 0.1065 M. The flasks were left for a period of time until the pH of a paired flask,

which was monitored periodically, reached stability. Resulting solutions were filtered

through a WhatmanR GF/F fibreglass filter and collected in clean PET bottles.

Dissolved metal concentrations were determined using ICP-MS by a NATA-compliant

commercial laboratory (mgt | LabMark Environmental Testing Australia Pty Ltd).

5.4 RESULTS

5.4.1 Syntheses of smirnite and chekhovichite

High purity, single phase samples of chekhovichite and smirnite were obtained by the

solid state methods mentioned above and were characterised by powder X-ray

diffraction. Unit cell parameters were refined in LAPOD (Langford, 1973) and obtained

values for chekhovichite of a = 18.897(3), b = 7.952(2), and c = 6.991(1) Å, and for

smirrnite of a = 16.442(8), b = 5.511(2) and c = 11.576(5) Å, these unit cell dimensions

agree well with those reported by Rossell et al. (1992) and Mercurio et al. (1983).

95 | P a g e
600

500

400
Lin (Counts)

300

200

100

5 10 20 30 40 50 60 70

2-Theta - Scale
Figure 5.3: Powder XRD trace of checkhovichite with peak positions of the ‘type’ pattern from JCPDS file 01-081-1330 (red).

96 | P a g e
1000

900

800

700

600
Lin (Counts)

500

400

300

200

100

5 10 20 30 40 50 60 70

2-Theta - Scale
Figure 5.4: Powder XRD trace of smirnite with peak positions of the ‘type’ pattern from JCPDS file 00-038-04200 (red).

97 | P a g e
5.4.2 Solubility Studies

The dissolution of chekhovichite, and smirnite are described in equations (1) and (2)

respectively.

Bi2Te4O11(s) + H2O(l) + 10H+ ⇋ 2Bi3+ + 4H3TeO3+ (1)

Bi2TeO5(s) + 7H+ ⇋ 2Bi3+ + H3TeO3+ 2H2O(l) (2)

Individual ion activity coefficients were calculated using the Davis extension of

the Debye-Hückel equation for 298.15 K, lg  = –0.5085z2(I/(1+I) – 0.3I). For

chekhovichite, I = 0.1069 mol dm–3, 3 = 0.1045 2 = 0.3665 and  = 0.778. For

smirnite I = 0.1068 mol dm–3, 3 = 0.1046, 2 = 0.3666 and  = 0.7781. In all

instances, o is taken to be unity. The activity of H+(aq), a(H+), in HNO3; 0.1065 M was

then calculated and corresponding activities a(Bi3+) and a(H3TeO3+) calculated from the

solubility data in Tables 5.1 and 5.2 .

The pH at equilibrium (1.087; chekhovichite) was used in the determination of

individual ion speciation using COMICS (Perrin and Sayce, 1967). A reliable lg K value

for equation (3) of lg K(298.15 K) = –2.83 and for equation (5) of lg K(298.15 K) = –

9.21 at I = 0 mol dm–3 was reported by McPhail (1995). Correction to I = 0.1069, mol

dm–3 by the method of Baes and Mesmer (1976), yields lg K(298.15 K) = –2.94 and –

9.21 respectively. Using the relation shown in equation (4), the concentration of

H2TeO3o(aq) is 1.663 x 10-8 and HTeO3-(aq) = 7.31 x 10–16 mol.dm–3.

98 | P a g e
Table 5.1. Dissolved metal concentrations for chekhovichite
Solution [Bi] ppm [Bi] mol dm-3
1 18 8.613 x 10-5
2 19 9.092 x 10-5
3 18 8.613 x 10-5
4 19 9.092 x 10-5
5 19 9.092 x 10-5
6 18 8.613 x 10-5
Mean 18.5 8.853 x 10-5
Error ±0.5 ±2.393 x 10-5

Table 5.2. Dissolved metal concentrations for smirnite


Solution [Bi] ppm [Bi] mol dm-3
1 12 5.742 x 10-5
2 12 5.742 x 10-5
3 11 5.264 x 10-5
4 12 5.742 x 10-5
5 12 5.742 x 10-5
6 13 6.221 x 10-5
Mean 12 5.742 x 10-5
Error ±1.0 ±4.79 x 10-6

H3TeO3+(aq) ⇋ H2TeO3o(aq) + H+(aq) (3)

[H2TeO3o] [H+]/[H3TeO3+] (4)

H3TeO3+(aq) ⇋ HTeO3-(aq) + H+(aq) (5)

[HTeO3-] 2.[H+]/[H3TeO3+] (6)

99 | P a g e
Bi3+(aq) + H2O(l) ⇋ Bi(OH)2+(aq) + H+(aq) (7)

[Bi(OH)2+].[H+]/[Bi3+] (8)

Bi3+(aq) + 2H2O(l) ⇋ Bi(OH)2+(aq) + 2H+(aq) (9)

[Bi(OH)2+].2[H+]/[Bi3+] (10)

Bi3+(aq) + 3H2O(l) ⇋ Bi(OH)30(aq) + 3H+(aq) (11)

[Bi(OH)30].3[H+]/[Bi3+] (12)

Reliable lg K values for equation (7), (9) and (11) of lg K(298.15 K) = –1.11, –

3.30 and –8.21 at I = 0 mol dm–3 were reported by van Der Lee and Lomenech (2004).

Correction to I = 0.1069, mol dm–3 by the method of Baes and Mesmer (1976), yields lg

K(298.15 K) = –1.65, –4.17 and –9.19. Using the relationships shown by equations (8),

(10) and (12), the concentration of Bi(OH)2+(aq), Bi(OH)2+(aq) and Bi(OH)3o(aq) are

1.886 x 10-5, 6.958 x 10-7 and 8.12 x 10-11 mol dm–3, respectively.

The pH at equilibrium (1.109; smirnite) was used in the determination of

individual ion speciation using COMICS (Perrin and Sayce, 1967). Correction to I =

0.1068, mol dm–3 by the method of Baes and Mesmer (1976), yields lg K(298.15 K) = –

2.94 for equation (3) and lg K(298.15 K) = –9.21 for equation (4). Using the relation

shown in equation (4), the concentration of H2TeO3o(aq) is 1.047 x 10-8 and HTeO3-(aq)

= 4.466 x 10–16 mol dm–3. Correction to I = 0.1068, mol dm–3 by the method of Baes and

Mesmer (1976) for equations (8), (10) and (12), yields lg K(298.15 K) = –1.66, –4.17

and –9.19. Using the relationships shown by equations (8), (10) and (12), the

100 | P a g e
concentration of Bi(OH)2+(aq), Bi(OH)2+(aq) and Bi(OH)3o(aq) are 1.23 x 10-5, 4.784 x

10-7 and 5.752 x 10-11 mol.dm–3 respectively.

Table 5.3. Thermodynamic quantities used in the calculations (T = 298.15 K).

ΔGfө /kJ.mol-1 Reference


chekhovichite Bi2Te4O11(s) –1457.23 ± 0.5 this study

smirnite Bi2TeO5(s) –820.92 ± 1.0 this study

Bi(OH)3o(aq) –572.61 ± 0.1 van Der Lee and Lomenech (2004)

Bi(OH)2+(aq) –363.54 ± 0.1 van Der Lee and Lomenech (2004)

Bi(OH)2+(aq) –138.94 ± 0.1 van Der Lee and Lomenech (2004)


H3TeO3+(aq) -490.7 McPhail (1995)

H2TeO30(aq) -474.6 McPhail (1995)

HTeO3-(aq) -438.2 McPhail (1995)

TeO32-(aq) -384 McPhail (1995)

Cl-(aq) –131.2 ± 0.1 Robie and Hemingway (1995)


H2O(l) –237.1 ± 0.1 Cox et al. (1989)

The activity of H+(aq), a(H+), in 0.1084 and 0.0104 M HNO3 was then

calculated and corresponding activities a(Bi3+), a(H3TeO3+), a(H2TeO3o), a(HTeO3-),

a(Bi(OH)2+), a(Bi(OH)2+), and a(Bi(OH)3o), were calculated from the solubility data.

This yields values of lg K for equations (1) and (2) of –14.86 ± 0.10 and -6.95 ± 0.12

respectively. Use of the appropriate data in Table 5.3 yields corresponding values for

ΔGrө(298.15 K) = +84.83 and +39.66 kJ mol–1. Thus, ΔGfө(Bi2Te4O11, s, 298.15 K) = –

1457.23 ± 0.5 kJ mol–1 and ΔGfө(Bi2TeO5, s, 298.15 K) = –820.92 ± 1.0 kJ mol–1. The

101 | P a g e
estimated error takes into account the analytical error of the solubility experiments and

errors quoted for the thermochemical data used.

5.5 REFERENCES

Afifi, A., Kelly, M., and Essene, E.J. (1988) Phase relationships among tellurides,

sulphides and oxides: I. Thermochemical data and calculated equilibria. Economic

Geology, 83, 337-394.

Baes, C.F. Jr and Mesmer, R.E. (1976) The Hydrolysis of Cations. Plenum Press, New

York.

Bosi, R. (2012) Chekhovichite: Bi2Te4+4O11 Zod Mine (Sotk deposit), Vardenis,

Gelark'unik' Province (Geghark'unik'), Armenia, photograph, viewed 6 March

2014, < http://www.mindat.org/photo-492561.html l>

Cox, J.D., Wagman, D.D. and Medvedev, V.A. (1989) CODATA Key Values for

Thermodynamics. Hemisphere Press, New York.

Deng, M. and Yang, X. (1993) Bi6Te2Mo2O21 - a new synthetic crystal and its physical

properties. Journal of Crystal Growth, 128, 876-879.

Jaireth, S. (1991) Hydrothermal geochemistry of Te, Ag2Te, and AuTe2 in epithermal

precious metal deposits. Economic Geology, 37. James Cook University of North

Queensland, Australia.

102 | P a g e
Kazachenko, V.T., Fat’yanov, I.I. and Chubanov. (1980) Discovery of a lead-containing

variety of montanite. Doklady Academii Nauk SSSR., 225 , 968-971

Langford, J.I. (1973) Least-squares refinement of cell dimensions from powder data by

Cohen's method. Journal of Applied Crystallography, 6, 190-196.

McPhail, D.C. (1995) Thermodynamic properties of aqueous tellurium species between

25 and 350°C. Geochimica et Cosmochimica Acta, 59, 5, 851-866.

Mercurio, D., El Farissi, M., Frit, B. and Goursat, P. (1983) Etude structurale et

densification d'un nouveau materiaux piezoelectrique: Bi2TeO5. Materials

Chemistry and Physics, 9, 467-476.

Ok, K.M., Bhucanesh, N.S.P. and Halasyamani, P.S. (2001) Bi2TeO5: synthesis,

structure and powder second harmonic generation properties, Inorganic Chemistry,

40, 1978-1980.

Palache, C., Berman, H. and Frondel, C. (1951) Dana’s System of Mineralogy, 2, 636-

637

Pavel, M. (2007) Smirnite : Bi2Te4+O5, Tsumoite : BiTe Shilovo-Isetskii mine,

Ekaterinburg, Sverdlovskaya Oblast', Middle Urals, Urals Region, Russia,

photograph, viewed 6 March 2014, <http://www.mindat.org/photo-134969.html>

103 | P a g e
Perrin, D.D. and Sayce, I.G. (1967) Computer calculation of equilibrium concentrations

in mixtures of metal ions and complexing species. Talanta, 14, 833-842.

Robie, R.A. and Hemingway, B.S. (1995) Thermodynamic properties of minerals and

related substances at 298.15K and 1 bar (105 Pascals) pressure and at higher

temperatures. United States Geological Survey Bulletin, 2131.

Rossell, H.J., Leblanc, M., F´erey, G., Bevan, D.J.M., Simpson, D.J. and Taylor, M.R.

(1992) On the crystal structure of Bi2Te4O11. Australian Journal of Chemistry, 45,

1415-1425.

Spiridonov, E.M., Petrova, I.V., Demina, L.A., Dolgikh, V.I. and Antonyan, G.M.

(1987) The new mineral chekhovichite (Bi2Te4O11). Moscow University Geology

Bulletin, 42, 71-75.

Szaller, Z., Pöppl, L., Lovas, G., and Dódony, I. (1996) Study of the formation of

Bi2Te4O11, Journal of Solid State Chemistry, 121, 251-261.

van der Lee, J. and Lomenech, C. (2004) Towards a common thermodynamic database

for speciation models. Radiochimica Acta, 92, 811-818.

Yang, X., Li, D., Wang, G., Deng, M., Chen, N. and Wang, S. (1989) A study of

chiluite - a new mineral found in Chilu, Fujian, China. Acta Mineralogica Sinica, 9,

9-14; American Mineralogist, 76, 666 (1990).

104 | P a g e
CHAPTER 6

Conclusions

105 | P a g e
6.1 ABSTRACT

A model of the dispersion of bismuth in the supergene zone has been developed

in this work. This was done by using the free energy of formation values and equations

derived in this thesis for cannonite, rooseveltite, koechlinite, russellite and smirnite

were compared to the two most common bismuth minerals bismoclite and bismutite. To

make this determination and provide a clear and reasonable understanding of the

dispersion of bismuth, an assessment of the Cobar region and Kingsgate region of NSW

was undertaken. This was done by reviewing the respective environments, the known

mineralogy and the reported geochemical data, as well as drawing upon previous

chapter’s data and conclusions to develop a reverse water solubility model, which has

major implications on the understanding of bismuth in the supergene zone.

6.2 INTRODUCTION

Every mineral deposit has, in some respects, a unique surface geochemical

“signature” due to differences in geological, geomorphological and environmental

settings. On the contrary many similarities in regards to elemental dispersion

characteristics may be displayed over extensive regions, either generically, for many

commodities, or more specifically for one commodity. A prerequisite for mineral

exploration in the supergene environment is the understanding and knowledge of the

geochemical “signature” of particular mineralisation (Leverett et al., 2004). The

dispersion of elements in the environment is heterogeneous and particular elements may

be widely dispersed compared to others, therefore it is necessary to have some

understanding of the chemical as well as physical controls that govern this dispersion.

Therefore, in the case of this study to draw conclusions on a proper understanding of the

106 | P a g e
dispersion of Bi in the environment will depend on appreciation of its low-temperature

aqueous chemistry and knowledge of the secondary Bi minerals that serve to buffer the

element between the solid and solution states.

Weathering, in a geochemical exploration context, causes the destruction of

primary ore deposits and the dispersion of ore and pathfinder elements in the

surrounding regolith. Conversely, it may also result in the supergene enrichment of

some deposits and promote the formation of secondary orebodies. To understand the

history and potential mechanisms and pathways of migration of ore and pathfinder

elements in regolith, it is necessary to unravel the complex superposition of events that

may have occurred during regolith-landscape evolution. According to McKinnon et al.

(2005) the ability of minerals to resist chemical weathering (i.e., dissolution) is

paramount when considering the hydromorphic dispersion of elements from an orebody

and, inter alia, the formation of geochemical anomalies. If appropriate geochemical

parameters can be estimated, and with an assumption that the mineral phases of interest

are in thermodynamic equilibrium with the groundwater, it is possible to evaluate the

volume of groundwater needed to completely dissolve a known quantity of a mineral.

This approach can be used to calculate the volume of proxy groundwater solution

required to dissolve one kilogram of bismutite (for example) at various pH values. This

is achieved by dividing the number of moles of bismuth in one kilogram of bismutite by

the total dissolved concentration of bismuth at any particular pH. The same calculation

can be used for other mineral species and the calculated volumes are estimated, as

factors such as changes in groundwater compositions during dissolution, the time

required for each system to reach equilibrium, and even variations in temperature will

serve to alter them. Even with these factors considered, the magnitude of differences in

107 | P a g e
results implies that some species are relatively resistant to chemical weathering, while

others will dissolve in quite small quantities of groundwater.

6.3 Discussion of Data

6.3.1 Chapter 2 Discussion

The work completed in chapter two can be used to determine the relationships

which control bismuth dispersion. The two basic carbonates bismutite, Bi2O2CO3, and a

synthetic phase of composition Bi4O4CO3(OH)2 must be taken into account. Taylor et

al. (1984) determined thermochemical data for both of these phases, relative to that of

bismite, α-Bi2O3. ΔGfө values have been recalculated using values for H2O(l), CO2(g)

and Bi2O3(s) at 298.2 K taken from Robie and Hemingway (1995) in order to establish

consistency and to eliminate error between data sets. This yields values of

ΔGfө(Bi2O2CO3, bismutite, 298.2 K) and ΔGfө(Bi4O4CO3(OH)2, 298.2K) of -916.2 ± 7.5

and -1649.9 ± 9.0 kJ mol-1, respectively. It is immediately apparent that Bi4O4CO3(OH)2

forms only at p(CO2) values below ambient (10-3.5) and that bismite is

thermodynamically unstable with respect to Bi4O4CO3(OH)2 under ambient conditions

when H2O is present or with respect to bismutite whenever p(CO2) > 10-5.5. Bismutite is

thus the stable phase under ambient conditions with respect both to Bi4O4CO3(OH)2 and

bismite. This is in line with the observation that Bi2O3, when exposed to air, undergoes

surface carbonation and that the preparation of pure Bi2O3 requires heating the samples

in an air stream at elevated temperatures (Barreca et al., 2001; Levin and Roth, 1964). A

first approach to an understanding of the behaviour of Bi in the supergene environment

can be based on bismutite, the most common secondary Bi mineral, bismoclite, BiOCl

and cannonite (Figure 6.1). A value for ΔGfө (BiOCl, bismoclite, 298.2 K), of -322.1 kJ

108 | P a g e
mol-1, was taken from Kaye and Laby (1995). The equilibrium conditions for equation

(1), (3) and (5) was then calculated using the data listed above, with ΔGfө (Cl-(aq), 298.2

K) = -131.2 kJ mol-1 (Robie and Hemingway, 1995), and ΔGfө (SO42-(aq),298.2 K) = -

744.0± 0.4 kJ mol-1 (Cox et al., 1989) this gives equation (2), (4) and (6) which are used

in Figure 6.1.

Bi2O2CO3(s) + 2H+(aq) + 2Cl-(aq) ⇋ 2BiOCl(s) + CO2(g) + H2O(l) (1)

pH = 8.50 – ½lg p(CO2) + lg a(Cl-) (2)

Bi2(SO4)O(OH)2(s) + 2Cl- (aq) ⇋ 2BiOCl(s) + SO42-(aq) + H2O(l) (3)

lg a(Cl-) = {-8.219 + lg a(SO42-)}/2 (4)

Bi2O2CO3(s) + SO42-(aq) + 2H+(aq) ⇋ Bi2(SO4)O(OH)2(s) + 2CO2 (g) (5)

pH = {8.727 –lg a(pCO2) + lg a(SO42-)}/2 (6)

pH
0 1 2 3 4 5 6 7 8 9 10 11
0

-1

-2
log activit of Cl-

-3 Bismoclite
-4

-5
Bismutite
-6

-7
Cannonite
-8

109 | P a g e
Figure 6.1: Relationship of cannonite, with common bismutite and bismoclite. a(SO42-)

= 10-2. Dashed and solid lines represent p(CO2) = 10-3.5 and p(CO2) = 10-2.5 (solid line),

respectively.

It is apparent that bismoclite dominates the stability field even at high a(SO42-),

thus for cannonite to persist in natural settings requires the depletion of the [Cl-] and

low pH. The formation of bismutite also requires almost complete exhaustion of both

Cl-(aq) as shown by Figure 6.1. It is clear that cannonite has a contribution to the

immobilisation of Bi, but not as great as that of bismoclite. Thus, the possibility for the

significantly smaller number of occurrences of cannonite as compared to bismoclite

reflects the persistent acidic conditions at localities where the a(Cl-) is slow enough or

expelled, without hosting bismoclite. Furthermore, the occurrence of riomarinaite is

limited to the minerals’ contact time with ground waters due to it being metastable.

6.3.2 Chapter 3 Discussion

The following equations describe the relationships which are considered to

control the stability of rooseveltite, preisingerite and atelestite. The equilibrium

conditions for equation (7, 9 and 11) were calculated using the data listed in Table 3.4,

which gives equations (8), (10) and (11).

3Bi2(AsO4)O(OH)(s) +H3AsO40(aq) ⇋ 2Bi3(AsO4)2O(OH)(s) + 2H2O(l) (7)

ΔGrө= –46.29. Thus, Log K = 8.108

Log K = 1/ a(H3AsO40) (8)

2BiAsO4(s) + 2H2O(l) ⇋ Bi2(AsO4)O(OH)(s) + H3AsO40(aq) (9)

110 | P a g e
ΔGrө= +38.69 Thus, Log K = –6.777

Log K = a(H3AsO40) (10)

3BiAsO4(s) + 2H2O(l) ⇋ Bi3(AsO4)2O(OH)(s) + H3AsO40(aq) (11)

ΔGrө= +34.84. Thus, Log K = –6.111

Log K = a(H3AsO40). (12)

The relative stability of the three phases depends only on the a(H3AsO40). Thus

rooseveltite is the favourable bismuth arsenate formed in secondary oxidised

environments, despite having fewer reported localities than preisingerite and atelestite.

As previously stated, the numerical values reported from mindat.org are only a guide.

Below are the equations for the relationship between rooseveltite and bismoclite, which

are shown in equations (13) and (14) where ΔGrө = -3.64 and log K = 0.64; therefore

bismoclite is favourable in an environment with an appreciable a(Cl-).

BiAsO4(s) + Cl-(aq) + H+(aq) + H2O(l) ⇋ BiOCl(s) + H3AsO40(aq) (13)

log K = log a(H3AsO40) – log a(Cl-) + pH (14)

This work has determined the conditions under which the rare minerals

rooseveltite, preisingerite and atelestite can form under ambient conditions in the

oxidised zones of Bi-bearing deposits. Therefore solution conditions which contain

H3AsO40 must be observed for the formation and stability of the basic Bi-As minerals.

Furthermore in these environments, it is possible that they affect other secondary Bi

species in regards to the immobilisation of Bi in the supergene environment. This

111 | P a g e
conclusion is significant as it provides a more refined focus on other secondary Bi

phases that play important roles in buffering dissolved Bi species in solution.

6.3.3Chapter 4 Discussion

Figure 6.2 and Figure 6.3 describe the relationships which are considered to

control the stability of koechlinite and russellite with bismoclite and bismutite. The

equilibrium conditions were calculated using the data listed in Table 4.4, with respect to

the various speciation of MoO42- and WO42- across the range of pH 0-11. It is apparent

from this work that we can conclude that the dominant bismuth secondaries are

koechlinite and russellite. Even when the activity of either MoO42- or WO42- is low, in

this case 10-6 both koechlinite and russellite dominate the stability field.

pH
0 1 2 3 4 5 6 7 8 9 10 11
0
-1
-2
bismoclite
Log activity of Cl-

-3
-4
-5
-6 bismutite
-7
-8 koechlinite
-9
-10

Figure 6.2: Stability relationships of koechlinite with other common Bi minerals.


The activity of total dissolved Mo(VI) = 10-6 and pCO2 = 10-3.5 (dashed line) and 10-2.5
(solid line). The breaks in the boundary between koechlinite and bismoclite correspond

112 | P a g e
to those points where a(MoO42-) = a(HMoO4-) at about pH 5 and where a(HMoO4-) =
a(H2MoO4O) at about pH 2.

pH
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
0
-1 bismoclite
-2
Log activity of Cl-

-3
-4 bismutite
-5
-6
-7 russellite
-8
-9
-10

Figure 6.3: Stability relationships of russellite with other common Bi minerals. The
activity of total dissolved W(VI) = 10-6 and pCO2 = 10-3.5 (dashed line) and 10-2.5 (solid
line). The break in the boundary between russellite and bismoclite corresponds to the
points where a(WO42-) = a(H2WO4O) at about pH 3. No data exist in the literature for
the first ionisation constant of H2WO40(aq).

Equation (15) describes the relationship between koechlinite and sardignaite.

Where ΔGrө = +63.30, therefore the formation of koechlinite is more favourable to that

of sardignaite

Bi2MoO6(s) + H2MoO40(aq) +3H+(aq) ⇋ BiMo2O7(OH)∙2H2O(s) + Bi3+(aq) (15)

113 | P a g e
This chapter has established that the common yellow coloured secondary

minerals of the Kingsgate region of NSW formerly assumed to be ferrimolybdite, are in

fact koechlinte or russellite and as such reported specimens need to be examined by X-

ray diffraction or Scanning Electron Microscopy to determine the difference,

furthermore the minerals koechlinite and russellite dominate the site (Chapter 1

Appendix A). Also established are the conditions in which russellite, koechlinite and

sardignaite can form under ambient conditions in the oxidised zones of Bi-Mo and Bi-

W deposits. Both koechlinite and russellite are stable phases and play a significant

effect, in relation to other secondary Bi species, and persist even at low [Mo] and [W],

in the immobilisation of Bi in the supergene environment of the Kingsgate Region of

NSW. This conclusion is important as it has refined our understanding of secondary Bi-

Mo and Bi-W phases which control the buffering dissolved Bi species in solution. This

is discussed further in this chapter in section 6.6

6.3.4.Chpater 5 Discussion

Equation (16) describes the relationship between smirnite and chekhovichite.

Bi2TeO5 (s) + 3 H3TeO3+ (aq) ⇋ Bi2Te4O11 (s) + 3H2O(l) + 3H+(aq) (16)

Where ΔGrө = +124.49, the formation of smirnite is more favourable to that of

chekhovichite as shown in Figure 6.4

114 | P a g e
pH
0 2 4 6 8 10
0
-2
Log activity of H3TeO3+

-4
smirnite
-6
-8
-10
-12
-14
-16
chekovichite
-18
-20

Figure 6.4: Stability field diagram of smirnite and chekhovichite.

Equation (17) and (18) describes the relationship between smirnite and

bismoclite for the species H3TeO3+(aq) and H2TeO30(aq)where ΔGrө = –51.58 and –

35.48 respectively.

Bi2TeO5 (s) + 2Cl-(aq) + 3H+(aq) ⇋ 2BiOCl(s) + H3TeO3+(aq) (17)

Bi2TeO5(s) + 2Cl-(aq) + 2H+(aq) ⇋ 2BiOCl(s) + H2TeO30(aq) (18)

This study has established the conditions in which smirnite and chekovichite can

form under ambient conditions in the oxidised zones of Bi-Te bearing deposits. Smirnite

is the more favourable phase and plays a significant effect on the immobilisation and

115 | P a g e
dispersion of bismuth in oxidising environments as seen in Figure 6.4 even though Bi-

Te mineralogy is rare. This conclusion has further refined our understanding of

secondary Bi-Te phases which control the buffering dissolved Bi species in ground

water solutions and can be used with previous work and data to make a fair assessment

on the dispersion of bismuth in the oxidised environment.

6.4 Reverse Ground Water Modelling

For the chosen conditions of the model, an equation was derived for every

conceivable complex and hydrolysed species of Bi3+ which are listed in Table 6.1

whereby state terms have been omitted for simplicity and the ΔGfө values used for the

minerals modelled are those listed in previous chapters of this thesis. Calculated

activities of dissolved species at equilibrium over the pH range from 0 to 11 for each

mineral of interest were then calculated and summed to yield a total dissolved [Bi3+] in

each case. Model conditions considered were for a(Cl–) = 10-1 (comparatively saline

conditions) and 10–4 (simulating rain water-flushed conditions), and p(CO2) = 10–3.5

(atmospheric) and 10–2.5 (simulating soil atmospheric levels) and a(SO42–)= 10-4. No

correction for ionic activity coefficients were made as this does not affect the

conclusions that may be drawn.

116 | P a g e
Table 6.1: Data used in species distributions in solutions at equilibrium with

solid phases. State terms have been omitted for simplicity. Data used in species

distributions in solutions at equilibrium with solid phases.

Equation lg K Ref
Bi + H2O ⇋ BiOH + H
3+ 2+ +
-1.11 1
Bi + 2H2O ⇋ Bi(OH)2 + 2H
3+ + +
-3.30 1
Bi + 3H2O ⇋ Bi(OH)3 + 3H
3+ 0 +
-8.21 1
Bi3+ + 4H2O ⇋ Bi(OH)4- + H+ -21.11 1
6Bi + 12H2O ⇋ Bi6(OH)22 + 12H
3+ 6+ +
-0.33 ±0.1 2
1.5Bi6(OH)22 + 2H2O ⇋ Bi9(OH)20 + 2H
6+ 7+ +
-3.5 ±0.1 2
Bi9(OH)207+ + H2O ⇋ Bi9(OH)216+ + H+ -3.2 ±0.1 2
Bi3+ + Cl- ⇋ BiCl2+ 2.4 2, 3
Bi3+ + 2Cl- ⇋ BiCl2+ 3.5 2, 3
Bi + 3Cl ⇋ BiCl3
3+ -
5.4 2, 3
Bi + 4Cl ⇋ BiCl4
3+ - -
6.1 2, 3
Bi3+ + 5Cl- ⇋ BiCl52- 6.7 2, 3
Bi + 6Cl ⇋
3+ -
BiCl63- 6.6 2, 3
3+
Bi + SO42- ⇋ BiSO4+ 2.01 4
Bi3+ + 2SO42- ⇋ Bi(SO4)- 3.41 4
3+
Bi + 3SO42- ⇋ Bi(SO4)33- 4.6 4
Bi3+ + 4SO4 ⇋ Bi(SO4)4
2- 5-
4.88 4
1: van Der Lee and Lomenech (2004); 2: Lovereček et al., (1985); 3: Smith and Martell
(1976); 4: Fedorov et al., (1971)

6.5 BISMUTH MODEL FOR THE COBAR REGION

The Cobar region is a major metallogenic province located within the Lachlan

old Belt and has been explored and mined since the late 1800’s. Exploration in this

region is difficult due to the complex regolith, deep weathering and extensive

transported cover components, as well as by the narrow and deep style and geometry of

the deposits. Due to the large ore deposits of the region, the geology is relatively well-

117 | P a g e
known and has been reported and updated (Andrews, 1911; Rayner, 1969; McClatchie,

1971; Suppel, 1982; Gilligan and Byrnes, 1994; Stegman and Pocock, 1996; Stegman,

2001; Stegman and Reynolds, 2005). Also, the completion of CRC LEME (2001-2008)

covers much of the information on Cobar-style deposits and as such it is not assessed in

this body of work.

However in regards to the role of bismuth, a proper assessment is still to be

made. Berthelsen (2004) noted that there is notable gold–bismuth and gold–copper

correlation in the deposits at Cobar (i.e. New Cobar, Peak, Perseverance and the New

Occidental mines). Minor primary bismuth mineralisation is a feature of Cobar-style

mineralisation (Andrews, 1911; Rayner, 1969; McClatchie, 1971; Suppel, 1982;

Gilligan and Byrnes, 1994; Stegman and Pocock, 1996; Reynolds, 1998; Stegman,

2001; Stegman and Reynolds, 2005). Native bismuth and bismuthinite have been

discovered along with ikunolite, Bi4(S,Se)4, maldonite, Au2Bi, guanajuatite, Bi2(Se,S)3,

volynskite, AgBiTe2, and lilianite, Pb3Bi2(S,Se)6 (Stegman and Pocock, 1996;

Reynolds, 1998; Stegman and Reynolds, 2005). Reported secondary bismuth

mineralisation in the Cobar Basin include bismutite (Bi2O2CO3), identified at the New

Cobar deposit (McKinnon, 2003) and the C.S.A. deposit (Stegman and Reynolds,

2005), as well as bismutite, bismite (Bi2O3) and bismoclite (BiOCl) from the Mount

Allen mine (Suppel and Gilligan, 1993).

Taking into consideration the common and rare secondary bismuth minerals and

know bismuth sulfosalts mentioned above from the Cobar region, an understanding of

the behaviour of Bi in the supergene environment can be made. If we consider the work

done in Chapter 2 which assessed the solubility and stability of the minerals bismutite,

118 | P a g e
bismoclite, and cannonite, we can determine a reasonable background concentration for

the area. For this purpose, the phase diagram shown in Chapter 2 Figure 2.1 can be

referenced.

Taking the ΔGfө(298.2 K) values for bismutite, bismoclite and cannonite, found in

Chapter 2, and modelling them in equilibrium with every conceivable dissolved species

then total dissolved Bi3+ can be evaluated. For this purpose, the equilibrium data of

Table 6.1 has been taken from the literature. In the calculations, a value for ΔGfө

(Bi3+,aq,298.2K) is needed. Lovreček et al., (1985) give the standard electrode potential,

Eө, at 298.2 K for equation (4) as 0.3172 ±0.0006 V; this corresponds to ΔGfө

(Bi3+,aq,298.2 K) = +91.8 ± 0.2 kJ.mol-1 and this is the value adopted here. It should be

noted that uncertainties in the value cancel out in calculations of dissolved Bi3+ species.

The following is given as an example calculation for bismoclite. Lg K (298.2 K) for

equation (1) is derived from ΔGfө data for constituent species given above and equation

(4) is the sum of (2) and (3).

Bi3+(aq) + 3e- ⇋ Bi(s) (1)

BiOCl(s) + 2H+(aq) ⇋ Bi3+(aq) + Cl-(aq) + H2O(l) lg K = -7.99 (2)

Bi3+(aq) + H2O(l) ⇋ BiOH2+(aq) + H+(aq) lg K = -1.11 (3)

BiOCl(s) + H+(aq) ⇋ BiOH2+(aq) + Cl-(aq) lg K = -9.10 (4)

119 | P a g e
Therefore, we can start to model specific sites based on the reported minerals

from specific regional or localised areas to define total dissolved bismuth values in any

given area.

Inspection of the data from Tables 6.2 to 6.5 shows that the maximum Bi

concentration in groundwater in equilibrium with the secondary Bi minerals at pH 5 is

about 10-8 M (ignoring activity corrections), when negligible Cl-(aq) ion is present, and

bismutite is the thermodynamically stable phase. This in turn corresponds to a total

solution Bi load of only 2.0 ppb. As a(Cl-) increases, total solution Bi is depressed

further. The presence of less soluble phases will have an impact and also serve to limit

Bi solution levels. In addition, soil gas concentrations of CO2 can reach partial pressures

an order of magnitude higher than those considered here, as the result of biological

activity (Appelo and Postma, 1993). This would result in an extension of the stability

field for bismutite and further lowering of solution Bi concentrations. Put simply, the

rate of dissolution of Bi minerals and chemical dispersion of the element in the regolith

is predicted to be very slow in comparison with most base metals.

120 | P a g e
Table 6.2: Total bismuth in solution at equilibrium with cannonite, bismoclite and bismutite Saline Water p(CO2) = 10-3.5, activity of Cl- = 10-1
and activity of SO42- = 10--3 (Saline) (absence of data refers to cases when the calculated activity is less than 10-13).
pH
Species 0 1 2 3 4 5 6 7 8 9 10 11
2+
BiOH 7.78E-09 7.78E-10 7.78E-11 7.78E-12 7.78E-13
Bi(OH)2+ 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 2.90E-12 2.90E-13
Bi(OH)3 3.57E-07 3.57E-07
-
Bi(OH)4 7.92E-13 7.92E-11 4.50E-10 4.50E-09
Bi6(OH)226+
Bi9(OH)207+
Bi9(OH)216+
BiCl2+ 2.56E-06 2.56E-08 2.56E-10 2.56E-12
+
BiCl2 3.23E-06 3.23E-08 3.23E-10 3.23E-12
BiCl3 2.57E-04 2.57E-06 2.57E-08 2.57E-10 2.57E-12
-
BiCl4 1.28E-05 1.28E-07 1.28E-09 1.28E-11 1.28E-13
2-
BiCl5 5.12E-06 5.12E-08 5.12E-10 5.12E-12
BiCl63- 4.07E-07 4.07E-09 4.07E-11 4.07E-13
BiSO4+ 1.02E-10 1.02E-12
Bi(SO4)2- 2.62E-10 2.62E-12
Bi(SO4)33- 4.15E-10 4.15E-12
Bi(SO4)45- 1.99E-11 1.99E-13
3+
Bi 1.02E-07 1.02E-09 1.02E-11 1.02E-13
Σ 2.81E-04 2.81E-06 2.82E-08 3.40E-10 5.49E-11 5.14E-11 5.13E-11 5.13E-11 5.21E-11 1.31E-10 3.58E-07 3.62E-07

121 | P a g e
Table 6.3: Total bismuth in solution at equilibrium with cannonite, bismoclite and bismutite Saline Water p(CO2) = 10-2.5, activity of Cl- = 10-1
and activity of SO42- = 10-3 (Saline) (absence of data refers to cases when the calculated activity is less than 10-13).
pH
Species 0 1 2 3 4 5 6 7 8 9 10 11
2+
BiOH 7.78E-09 7.78E-10 7.78E-11 7.78E-12 7.78E-13
Bi(OH)2+ 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 2.90E-11 2.90E-12 2.90E-13
Bi(OH)3 3.57E-07 3.57E-07 3.57E-07
-
Bi(OH)4 7.92E-13 4.50E-11 4.50E-10 4.50E-09
Bi6(OH)226+
Bi9(OH)207+
Bi9(OH)216+
BiCl2+ 2.56E-06 2.56E-08 2.56E-10 2.56E-12
+
BiCl2 3.23E-06 3.23E-08 3.23E-10 3.23E-12
BiCl3 2.57E-04 2.57E-06 2.57E-08 2.57E-10 2.57E-12
-
BiCl4 1.28E-05 1.28E-07 1.28E-09 1.28E-11 1.28E-13
2-
BiCl5 5.12E-06 5.12E-08 5.12E-10 5.12E-12
BiCl63- 4.07E-07 4.07E-09 4.07E-11 4.07E-13
BiSO4+ 1.02E-10 1.02E-12
-
Bi(SO4)2 2.62E-10 2.62E-12
Bi(SO4)33- 4.15E-10 4.15E-12
Bi(SO4)45- 1.99E-11 1.99E-13
3+
Bi 1.02E-07 1.02E-09 1.02E-11 1.02E-13
Σ 2.81E-04 2.81E-06 2.82E-08 3.40E-10 5.49E-11 5.14E-11 5.13E-11 5.13E-11 5.21E-11 3.57E-07 3.58E-07 3.62E-07

122 | P a g e
Table 6.4: Total bismuth in solution at equilibrium with cannonite, bismoclite and bismutite Saline Water p(CO2) = 10-3.5, activity of Cl- = 10-4
and activity of SO42- = 10-3 (fresh) (absence of data refers to cases when the calculated activity is less than 10-13).
pH
Species 0 1 2 3 4 5 6 7 8 9 10 11
BiOH2+ 7.78E-06 7.78E-07 7.78E-08 7.78E-09 7.78E-10 7.78E-11 7.78E-12 1.37E-13
+
Bi(OH)2 5.13E-08 5.13E-08 5.13E-08 5.13E-08 5.13E-08 5.13E-08 5.13E-08 9.18E-09 9.18E-10 9.18E-11 9.18E-12 9.18E-13
Bi(OH)3 6.29E-13 1.13E-06 1.13E-06 1.13E-06 1.13E-06 1.13E-06
-
Bi(OH)4 1.42E-12 1.42E-11 1.42E-10 1.42E-09 1.42E-08
Bi6(OH)226+
Bi9(OH)207+
Bi9(OH)216+
BiCl2+ 2.56E-06 2.56E-08 2.56E-10 2.56E-12
BiCl2+ 3.23E-09 3.23E-11 3.23E-13
BiCl3 2.57E-10 2.57E-12
BiCl4-
BiCl52-
BiCl63-
BiSO4+ 1.02E-07 1.02E-09 1.02E-11 1.02E-13
-
Bi(SO4)2 2.62E-07 2.62E-09 2.62E-11 2.62E-13
Bi(SO4)33- 4.15E-07 4.15E-09 4.15E-11 4.15E-13
Bi(SO4)45- 1.99E-08 1.99E-10 1.99E-12
Bi3+ 1.02E-04 1.02E-06 1.02E-08 1.02E-10 1.02E-12
Σ 1.13E-04 1.88E-06 1.40E-07 5.92E-08 5.21E-08 5.14E-08 5.13E-08 1.14E-06 1.13E-06 1.13E-06 1.13E-06 1.14E-06

123 | P a g e
Table 6.5: Total bismuth in solution at equilibrium with cannonite, bismoclite and bismutite Saline Water p(CO2) = 10-2.5, activity of Cl- = 10-4
and activity of SO42- = 10-3 (fresh) (absence of data refers to cases when the calculated activity is less than 10-13).
pH
Species 0 1 2 3 4 5 6 7 8 9 10 11
BiOH2+ 7.78E-06 7.78E-07 7.78E-08 7.78E-09 7.78E-10 7.78E-11 4.33E-12
+
Bi(OH)2 5.13E-08 5.13E-08 5.13E-08 5.13E-08 5.13E-08 5.13E-08 2.90E-08 2.90E-09 2.90E-10 2.90E-11 2.90E-12 2.90E-13
Bi(OH)3 6.29E-13 3.57E-07 3.57E-07 3.57E-07 3.57E-07 3.57E-07 3.57E-07
-
Bi(OH)4 4.50E-13 4.50E-12 4.50E-11 4.50E-10 4.50E-09
Bi6(OH)226+
Bi9(OH)207+
Bi9(OH)216+
BiCl2+ 2.56E-06 2.56E-08 2.56E-10 2.56E-12
BiCl2+ 3.23E-09 3.23E-11 3.23E-13
BiCl3 2.57E-10 2.57E-12
BiCl4-
BiCl52-
BiCl63-
BiSO4+ 1.02E-07 1.02E-09 1.02E-11 1.02E-13
-
Bi(SO4)2 2.62E-07 2.62E-09 2.62E-11 2.62E-13
3-
Bi(SO4)3 4.15E-07 4.15E-09 4.15E-11 4.15E-13
Bi(SO4)45- 1.99E-08 1.99E-10 1.99E-12
Bi3+ 1.02E-04 1.02E-06 1.02E-08 1.02E-10 1.02E-12
Σ 1.13E-04 1.88E-06 1.40E-07 5.92E-08 5.21E-08 5.14E-08 3.86E-07 3.60E-07 3.58E-07 3.57E-07 3.58E-07 3.62E-07

124 | P a g e
6.6 BISMUTH MODEL FOR THE NEW ENGLAND REGION

This thesis has previously mentioned the Kingsgate region of Northern NSW

and its reported mineralogy has been outlined in Chapter 1, Appendix A and Chapter 4.

Therefore, in this example we would need to consider and impose a number of limits to

provide an appropriate summation on the solubility of Bi3+. For the pH conditions found

in the Kingsgate regolith we find that pH is expected to vary between about 4 and 6.5,

with a median of about 5. Soils covering granites of the New England region are known

to have pH values of less than 5.5 in all horizons (Dolling et al., 2001). Also, the soils

covering mineralised areas, in this case they are predominantly shallow, with an umbric

A horizon formed over acid siliceous rocks.

Inspection of the data from Tables 6.6 to 6.9 shows that the maximum Bi

concentration in groundwater in equilibrium with the secondary Bi minerals koechlinite,

bismoclite and bismutite at pH 5 is about 10-9 M (ignoring activity corrections). This in

turn corresponds to a total solution Bi load of only 0.2 ppb. As a(Cl-) increases, total

solution Bi is depressed further at 10-11M and is limited to 2.0 ppt. Put simply, the rate

of dissolution of Bi minerals in the New England region and chemical dispersion of the

element in the regolith is predicted to be very slow in comparison with most base

metals. Therefore we can conclude that bismuth is not very mobile in the supergene

environment. The mineralogical evidence and experimental solubility evidence

demonstrates this and, in the context of exploration geochemistry of eastern Australia,

has largely been untested as the use of Bi and Mo in these environment as pathfinders

are of little to no use on a large-scale regional pathfinder. However, Bi would be useful

as a localised pathfinder.

125 | P a g e
Table 6.6: Total bismuth in solution at equilibrium with koechlinite, bismoclite and bismutite Saline Water p(CO2) = 10-2.5, activity of Cl- = 10-1
activity of SO42- = 10-4 and activity of WO42- = 10-6 (absence of data refers to cases when the calculated activity is less than 10-13).
pH
Species 0 1 2 3 4 5 6 7 8 9 10 11
2+
BiOH 7.78E-09 7.78E-10 7.78E-11 7.78E-12 7.78E-13
+
Bi(OH)2 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 2.90E-11 2.90E-12 2.90E-13
Bi(OH)3 3.57E-07 3.57E-07 3.57E-07
Bi(OH)4- 7.92E-13 4.50E-11 4.50E-10 4.50E-09
6+
Bi6(OH)22
Bi9(OH)207+
Bi9(OH)216+
BiCl2+ 2.56E-06 2.56E-08 2.56E-10 2.56E-12
BiCl2+ 3.23E-06 3.23E-08 3.23E-10 3.23E-12
BiCl3 2.57E-04 2.57E-06 2.57E-08 2.57E-10 2.57E-12
BiCl4- 1.28E-05 1.28E-07 1.28E-09 1.28E-11 1.28E-13
BiCl52- 5.12E-06 5.12E-08 5.12E-10 5.12E-12
BiCl63- 4.07E-07 4.07E-09 4.07E-11 4.07E-13
BiSO4+ 1.02E-10 1.02E-12

Bi(SO4)2- 2.62E-10 2.62E-12

Bi(SO4)33- 4.15E-10 4.15E-12

Bi(SO4)45- 1.99E-11 1.99E-13

Bi3+ 1.02E-07 1.02E-09 1.02E-11 1.02E-13


Σ 2.81E-04 2.81E-06 2.82E-08 3.40E-10 5.49E-11 5.14E-11 5.13E-11 5.13E-11 5.21E-11 3.57E-07 3.58E-07 3.62E-07

126 | P a g e
Table 6.7: Total bismuth in solution at equilibrium with koechlinite, bismoclite and bismutite Saline Water p(CO2) = 10-3.5, activity of Cl- = 10-1
activity of SO42- = 10-4 and activity of WO42- = 10-6(absence of data refers to cases when the calculated activity is less than 10-13).
pH
Species 0 1 2 3 4 5 6 7 8 9 10 11
BiOH2+ 7.78E-09 7.78E-10 7.78E-11 7.78E-12 7.78E-13
+
Bi(OH)2 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 5.13E-11 9.18E-12 9.18E-13
Bi(OH)3 1.13E-06 1.13E-06
-
Bi(OH)4 7.92E-13 7.92E-11 1.42E-09 1.42E-08
Bi6(OH)226+
Bi9(OH)207+
Bi9(OH)216+
BiCl2+ 2.56E-06 2.56E-08 2.56E-10 2.56E-12
BiCl2+ 3.23E-06 3.23E-08 3.23E-10 3.23E-12
BiCl3 2.57E-04 2.57E-06 2.57E-08 2.57E-10 2.57E-12
BiCl4- 1.28E-05 1.28E-07 1.28E-09 1.28E-11 1.28E-13
2-
BiCl5 5.12E-06 5.12E-08 5.12E-10 5.12E-12
BiCl63- 4.07E-07 4.07E-09 4.07E-11 4.07E-13
BiSO4+ 1.02E-10 1.02E-12
-
Bi(SO4)2 2.62E-10 2.62E-12
3-
Bi(SO4)3 4.15E-10 4.15E-12
Bi(SO4)45- 1.99E-11 1.99E-13
Bi3+ 1.02E-07 1.02E-09 1.02E-11 1.02E-13
Σ 2.81E-04 2.81E-06 2.82E-08 3.40E-10 5.49E-11 5.14E-11 5.13E-11 5.13E-11 5.21E-11 1.31E-10 1.13E-06 1.14E-06

127 | P a g e
Table 6.8: Total bismuth in solution at equilibrium with koechlinite, bismoclite and bismutite Fresh Water p(CO2) = 10-2.5, activity of Cl- = 10-4
activity of SO42- = 10-4 and activity of WO42- = 10-6 (absence of data refers to cases when the calculated activity is less than 10-13).
pH
Species 0 1 2 3 4 5 6 7 8 9 10 11
BiOH2+ 7.78E-06 7.78E-07 7.78E-08 5.60951E-09 1.77388E-10 5.54607E-12 5.54607E-13
+
Bi(OH)2 5.13E-08 5.13E-08 5.13E-08 3.68383E-08 1.16493E-08 3.64216E-09 3.64216E-09 2.90E-09 2.90E-10 2.90E-11 2.90E-12 2.90E-13
Bi(OH)3 6.29E-13 1.43307E-11 1.43307E-11 4.48051E-09 4.48051E-08 3.57E-07 3.57E-07 3.57E-07 3.57E-07 3.57E-07
-
Bi(OH)4 4.50E-13 4.50E-12 4.50E-11 4.50E-10 4.50E-09
Bi6(OH)226+
Bi9(OH)207+
Bi9(OH)216+
BiCl2+ 2.56E-06 2.56E-08 2.56E-10 1.84651E-12
BiCl2+ 3.23E-09 3.23E-11 3.23E-13
BiCl3 2.57E-10 2.57E-12
BiCl4-
BiCl52-
BiCl63-
BiSO4+ 1.02E-07 1.02E-09 1.02E-11 2.37531E-13
-
Bi(SO4)2 2.62E-07 2.62E-09 2.62E-11 1.8851E-13
3-
Bi(SO4)3 4.15E-07 4.15E-09 4.15E-11 2.9902E-13
Bi(SO4)45- 1.99E-08 1.99E-10 1.99E-12
Bi3+ 1.02E-04 1.02E-06 1.02E-08 7.35383E-11 2.32548E-13
Σ 1.13E-04 1.88E-06 1.40E-07 4.25382E-08 1.18412E-08 8.12822E-09 4.84478E-08 3.60E-07 3.58E-07 3.57E-07 3.58E-07 3.62E-07

128 | P a g e
Table 6.9: Total bismuth in solution at equilibrium with koechlinite, bismoclite and bismutite Fresh Water p(CO2) = 10-.3.5, activity of Cl- = 10-4
activity of SO42- = 10-4 and activity of WO42- = 10-6 (absence of data refers to cases when the calculated activity is less than 10-13).
pH
Species 0 1 2 3 4 5 6 7 8 9 10 11
2+
BiOH 7.78E-06 7.78E-07 7.78E-08 2.49041E-10 2.49041E-11 2.49041E-12 2.49041E-13

Bi(OH)2+ 5.13E-08 5.13E-08 5.13E-08 1.60931E-09 1.60931E-09 1.60931E-09 1.60931E-09 1.60931E-09 9.18E-10 9.18E-11 9.18E-12 9.18E-13

Bi(OH)3 6.29E-13 1.97973E-11 1.97973E-10 1.97973E-09 1.97973E-08 1.97973E-07 1.13E-06 1.13E-06 1.13E-06 1.13E-06

Bi(OH)4- 2.49485E-13 1.42E-11 1.42E-10 1.42E-09 1.42E-08

Bi6(OH)226+
Bi9(OH)207+
Bi9(OH)216+
BiCl2+ 2.56E-06 2.56E-08 2.56E-10

BiCl2+ 3.23E-09 3.23E-11 3.23E-13

BiCl3 2.57E-10 2.57E-12

BiCl4-
BiCl52-
BiCl63-
BiSO4+ 1.02E-07 1.02E-09 1.02E-11 3.28141E-13
-
Bi(SO4)2 2.62E-07 2.62E-09 2.62E-11

Bi(SO4)33- 4.15E-07 4.15E-09 4.15E-11

Bi(SO4)45- 1.99E-08 1.99E-10 1.99E-12


3+
Bi 1.02E-04 1.02E-06 1.02E-08

Σ 1.13E-04 1.88E-06 1.40E-07 1.87858E-09 1.83219E-09 3.59153E-09 2.14069E-08 1.99583E-07 1.13E-06 1.13E-06 1.13E-06 1.14E-06

129 | P a g e
6.7 CONCLUSION

In conclusion, the chemical rationale for bismuth’s geochemically immobile nature

can be understood in the weathering environment. This work explains the differential

leaching and mobility of Bi the Kingsgate and Cobar deposits. The model outlined above

explains the limited mobility at Kingsgate and suggests that much of the New England

Orogen remains untested as the values obtained in this work are found to be lower than the

reported arbitrary values found in the literature (MacDuff and Snow, 1971; MacDuff and

Zerwick 1971a; MacDuff and Zerwick 1971b; MacDuff and Zerwick 1972; Siegal, 1974;

Rose et al., 1979; Levinson, 1980; Plant et al., 1989; Plant et al.,1991; Filella, 2010). Also,

the use of conventional Atomic Absorption Spectroscopy (AAS) is not a useful analytical

method for the determination of Bi at low levels. This is important given the data above

concerning the distribution of the element in rocks, soils and natural waters. Even though

modern ICP-OES or ICP-MS and other methods are both accurate and precise to much lower

levels (Filella, 2010; Gholivand and Romiani, 2006) and hydride generation methods have

greatly increased the sensitivity of Bi determinations using AAS (Campbell, 1992) these

techniques would still be inadequate in some circumstances.

This work also highlights the persistence of secondary Bi minerals in the highly

weathered gossans of Cobar-style ores (Leverett et al., 2003, 2004, 2005a,b). The model

provides the confidence that Bi can be successfully used as a localised pathfinder element in

geochemical exploration when it is present in significant amounts in primary ores. Cobar-

style mineralisation is known to contain considerable amounts of Bi (Stegman and Reynolds,

2005). Refractory elements such as Sn in these deposits have been suggested as being useful

exploration guides (Leverett et al., 2004); Bi may also be useful, especially in Au-poor

deposits such as Elura, to the north of Cobar. Stable isotope measurements of some ground

130 | P a g e
waters in the region (Leverett et al., 2003, 2004) indicate that they are quite old and serve

well as proxies for those that had an input to the oxidised zones, despite a very long

weathering history (Pillans et al., 1999). This conclusion applies to exploration programs

carried-out in the past with the same kinds of analytical constraints. The model developed

above requires that the results be revaluated and this model should be taken into

consideration for future planning and development of geochemical exploration strategies

along the East Coast of Australia.

6.7 REFERENCES

Andrews, E.C. (1911) Cobar Copper and Gold Field. Part I. New South Wales Department

of Mines, Sydney.

Barreca, D., Morazzoni, F., Rizzi, G.A, Scotti, R. and Tondello, E. (2001) Molecular oxygen

interaction with Bi2O=: a spectroscopic and spectromagnetic investigation. Physical

Chemistry and Chemical Physics, 3, 1743-1749.

Berthelsen, R.R., (2004) Exploration in the Cobar Gold Field: A 2004 Perspective. In:

McQueen, K.G. and Scott, K.M. (2004) Exploration Field Workshop Cobar Region

2004,Proceedings. CRC LEME Report, Cooperative Research Centre for Landscape

Environments and Mineral Exploration, Perth WA. 5-9.

Campbell, A.D. (1992) A critical survey of hydride generation techniques in atomic

spectroscopy. Pure and Applied Chemistry, 64, 227-244.

131 | P a g e
Cox, J.D., Wagman, D.D. and Medvedev, V.A. (1989) CODATA Key Values for

Thermodynamics. Hemisphere Press, New York.

Dolling, P.J., Moody, P., Noble, A., Helyar, K., Hughes, B., Reuter, D. and Sparrow, L.

(2001) Soil Acidity and Acidification. Final report to National Land and Water Resources

Audit for Project 5.4C.

Fedorov, V.A., Kalosh, T.N., Chernikova, G.E. and Mironov, V.E. (1971) Sulphato-

complexes of bismuth(III). Russian Journal of Physical Chemistry, 45, 106.

Filella, M. (2010) How reliable are environmental data on 'orphan' elements? The case of

bismuth concentrations in surface waters. Journal of Environmental Monitoring, 12, 90-

109.

Gholivand, M.B. and Romiani, A.A. (2006) Highly sensitive and selective measurement of

bismuth in seawater and drug with 1,2-phenylenedioxydiacetic acid by cathodic

adsorptive stripping voltammetry. Electroanalysis, 18, 730-734.

Gilligan, L.B. & Byrnes, J.G. (1994) Cobar 1:250 000 Metallogenic Map SH/55-14: Metal

Study and Mineral Deposit Data Sheets. Geological Survey of New South Wales,

Sydney

Kaye, G.W.C. and Laby, T.H. (1995) Tables of Physical and Chemical Constants. 16th

Edition, Longman, London, Section 3.10.5.

132 | P a g e
Leverett, P., McKinnon, A.R. and Williams, P.A. (2003) Mineralogy of the oxidised zone of

the New Cobar orebody. In: I.C. Roach (Ed.) Advances in Regolith, CRC LEME,

Canberra, 267-270.

Leverett, P., McKinnon, A.R. and Williams, P.A. (2004) A supergene exploration model for

Cobar style deposits. In: McQueen, K.G. and Scott, K.M. (Eds.) Exploration Field

Workshop Cobar Region 2004, Proceedings, CRC LEME, Perth, 46-50

Leverett, P., McKinnon, A.R. and Williams, P.A. (2005a) Supergene geochemistry of the

Endeavor ore body, Cobar, NSW, and relationships to other deposits in the Cobar Basin.

In: I.C. Roach (Ed.) Regolith 2005 – Ten Years of CRC LEME, CRC LEME, 191-194.

Leverett, P., McKinnon, A.R., Sharpe, J.L. and Williams, P.A. (2005b) Secondary minerals

from the central Cobar mines. Australian Journal of Mineralogy, 11, 75-82.

Levin, E.M. and Roth, R.S. (1964) Polymorphism of bismuth sesquioxide. I. Pure Bi 2O3.

Journal of Research of the National Bureau of Standards, 68A, 189-195.

Levinson, A.A., (1980) Introduction to Exploration Geochemistry: The Supplement, Applied

Publishing Ltd, Wilmette, USA, 1980, 615-924.

Lovreček, B., Mekjavić, I. and Metikoš-Huković, M. (1985) Bismuth. In: Bard, A.J., Parsons,

R. and Jordan, J. (Eds) Standard Potentials in Aqueous Solution. International Union of

Pure and Applied Chemistry and Marcel Dekker, New York, 180-187.

133 | P a g e
Pillans, B., Tonui, E. and Indurm, M. (1999) Paleomagnetic dating and weathered regolith.

In: Taylor, G. and Pain, C. (Eds) Approaches to an Old Continent, Proceedings of the

Regolith 98 Conference, Kalgoorlie, CRC LEME, Perth, 237-242.

Plant, J.A., Breward, N., Forrest, M.D. and Smith, R.T. (1989) The gold pathfinder elements

As, Sb and Bi - their distribution and significance in the southwest Highlands of Scotland.

Transactions of the Institution of Mining and Metallurgy, 98, B91-101.

Plant, J.A., Cooper, D.C., Green, P.M., Reedman, A.J. and Simpson, P.R. (1991) Regional

distribution of As, Sb and Bi in the Grampian Highlands of Scotland and English Lake

District: implications for gold metallogeny. Transactions of the Institution of Mining and

Metallurgy, 100, B135-147.

MacDuff, R. and Snow, A. (1971) Quarterly Report Period to 8th January 1971, EL208,

AOG Minerals PL. New South Wales Department of Primary Industries (Mineral

Resources) Open File Reports, No. R00028059.

MacDuff, R. and Zerwick, J. (1971a) EL208, Quarterly Report Period to 8th July 1971, AOG

Minerals PL. New South Wales Department of Primary Industries (Mineral Resources)

Open File Reports, No. R00007895.

MacDuff, R. and Zerwick, J. (1971b) EL208, Quarterly Report Period to 8th October 1971,

AOG Minerals PL. New South Wales Department of Primary Industries (Mineral

Resources) Open File Reports, No. R00007896.

134 | P a g e
MacDuff, R. and Zerwick, J. (1972) EL208, Quarterly Report Period to 8th January 1972,

AOG Minerals PL. New South Wales Department of Primary Industries (Mineral

Resources) Open File Reports, No. R00007897.

McClatchie, L. (1971) Base metal mineralisation at Mineral Hill, central Western New South

Wales. Geological Survey of New South Wales, Sydney.

McKinnon, A.R., Sharpe, J.L. & Williams, P.A. (2005) Other notable mineral occurrences in

the Cobar region. Australian Journal of Mineralogy, 11, 117-121.

Rayner, E.O. (1969) The copper ores of the Cobar region, New South Wales Geological

Survey of New South Wales, Sydney.

Robie, R.A. and Hemingway, B.S. (1995) Thermodynamic properties of minerals and related

substances at 298.15K and 1 bar (105 Pascals) pressure and at higher temperatures.

United States Geological Survey Bulletin, 2131.

Rose, A.W., Hawkes, H.E. and Webb, J.S., 1979. Geochemistry in Mineral Exploration

Academic Press, New York, N.Y., pp. 490--517.

Smith, R.M. and Martell, A.E., 1976: Critical Stability Constants. 4. Inorganic Complexes.

Plenum Press, New York.

Siegel, F. R., 1974, Applied geochemistry: New York, John Wiley and Sons, 353p.

135 | P a g e
Stegman, C.L. (2001) Cobar deposits: Still defying classification. SEG Newsletter, 44.

Stegman, C.L. and Pocock, J.A. (1996) The Cobar gold field - a geological perspective. In

Cook, W.G., Ford, A.J.H., McDermot, J.J., Standish, P.N., Stegman, C.L. and Pocock,

J.A. (Eds) The Cobar Mineral Field – A 1996 Perspective, Australasian Institute of

Mining and Metallurgy, Melbourne, pp. 229‐264.

Stegman, C.L. & Reynolds, I. (2005) Primary mineralisation in Cobar deposits, with

emphasis on gold deposits of the Cobar Goldfield. Australian Journal of Mineralogy, 11,

63-72.

Suppel, D.W. (1982) The geology and mineral deposits of the Mount Hope area, New South

Wales. Geological Survey of New South Wales Report, GS1982/476.

Suppel, D.W. & Gilligan, L.B. (1993) Nymagee Metallogenic Map 1:250 000 SI/55- 2:

Metallogenic Study and Mineral Deposit Data Sheets. Geological Survey of New South

Wales, Sydney.

Taylor, P., Sunder, S. and Lopata, V.J. (1984) Structure, spectra, and stability of solid

bismuth carbonates. Canadian Journal of Chemistry, 62, 2863-2873.

van der Lee, J. and Lomenech, C. (2004) Towards a common thermodynamic database for

speciation models. Radiochimica Acta, 92, 811-818.

136 | P a g e
Appendix A

Secondary Bi and Mo minerals identified in selected Australian depositsa.

NEW SOUTH WALES

Bald Nob
D30997b Bismutite

Captains Flat
M15216 Tetragonal bismuth oxide (new mineral) with bismutite, koechlinite,
russellite

Duckmaloi
D30994 Bismutite

Deepwater
D21997 Bismutite, koechlinite, chekhovichite (Bi2Te4O11)
D22221 Koechlinite, bismutite
D22232 Bismutite, koechlinite.
D28072 Bismutite
D37006 Koechlinite

Dundee
D24224 Koechlinite
D28509 Koechlinite, quartz
D28510 Bismite, bismutite
D28511 Koechlinite, bismutite
D28512 As for D28511
D35450 Zavaritskite, bismite, bismutite, koechlinite
D35450 Bismutite.

Elsmore
D15631 Bismoclite
D28050 Bismoclite

Emmaville
D18210 Russellite, bismutite

137 | P a g e
Glen Eden
D35884 Bismutite, koechlinite

Glen Elgin
D31404 Koechlinite, bismutite

Gumble Flat
D31051 Bismutite

Kingsgate
D2969 Bismutite, koechlinite
D2970 Koechlinite
D2971 Bismutite, koechlinite.
D2973 Koechlinite
D2974 Bismutite, bismite, koechlinite
D2975 Preisingerite, bismutite.
D7194 Koechlinite, bismutite
D22004 Bismutite
D30956 Bismutite, koechlinite.
D31339 Koechlinite
D31340 Wulfenite, koechlinite, bismutite
D31341 Wulfenite, koechlinite, bismutite
D31343 Bismutite, preisingerite
D31420 Bismutite.
D38036 Bismutite, chekhovichite, smrkovecite
M30035 Bismutite, bismite
M47205 Ferrimolybdite

Nanima
D18931 Bismutite
D25033 Bismutite.

Mt Gipps
D15637 Bismutite, kettnerite

138 | P a g e
Mt Razorback
D35482 Bismutite

Murrumbateman
D30998 Bismutite

Torrington
D18209 Bismutite, russellite
D28513 Bismutite
D29744 Bismutite, russellite
D29745 Bismutite, russellite
D29746 Russellite

Whipstick
D18930 Koechlinite
D31052 Koechlinite

Unknown
D36340 Bismutite, russellite
NORTHERN TERRITORY

Alice Springs
D30600 Bismutite

Hatches Creek
M18636 Bismutite

Tennant Creek
M48721 Bismutite; XRDMV
M49700 Bismutite, bismite

Wolfram Hill
D51120 Bismutite, russellite
QUEENSLAND

Bamford
D30972 Bismutite
D30991 Koechlinite

139 | P a g e
M7425 Ferrimolybdite; XRDMV

Biggenden
D31014 Bismutite, preisingerite
D31016 Bismutite, preisingerite

Chillagoe
D18284 Bismutite
D18284 Bismutite
D18285 Bismutite
D31026 Bismutite, preisingerite
D31027 Bismutite
M49613 Russellite; XRDMV

Ukalunda
D31017 Bismutite

Halifax Bay
D31012 Bismutite, koechlinite
D31013 Bismutite, minor bismoclite

Kaboonga
D31045 Bismutite

Malbon
D30593 Bismutite

Mt Hastings
D31454 Bismutite

Mt Shamrock
D31048 Bismoclite

Mt Wyatt
D31035 Bismutite with minor quartz and amorphous iron oxides.

Wolfram Camp
D31451 Koechlinite
D31452 Koechlinite, bismutite

140 | P a g e
D31453 Russellite
D51077 Bismutite, koechlinite
M40932 Russellite
SOUTH AUSTRALIA

Balhanna mine
M30249 Bismutite
TASMANIA

Forth River Gorge


M47771 Ferrimolybdite; XRDMV

Shepherd and Murphy mine


D17161 Bismutite
D31041 Bismutite, cannonite
D31042 Bismutite, bismoclite
D31043 Bismutite
D31448 Bismutite
VICTORIA

Beechworth
M15403 Bismutite; XRDMV
M43565 Bismutite; XRDMV

Bendoc
D21624 Bismutite, zairite

Everton
M38367 Ferrimolybdite
M40860 Ferrimolybdite; XRDMV
M42790 Ferrimolybdite
M47686 Mendozavilite

Gombianbar Creek
M274 Bismutite
Maldon

141 | P a g e
M7552 Bismutite
M15402 Bismutite

Mt Moliagul
M39540 Ferrimolybdite
M40723 As for M39540

Mt Stanley
M41176 Ferrimolybdite

Pittong
M47373 Koechlinite, bismoclite; XRDMV
M47621 Koechlinite; XRDMV
M47622 Bismoclite; XRDMV

Thologolong
M48714 Koechlinite; XRDMV

Wangrabell
M46176 Ferrimolybdite; XRDMV
WESTERN AUSTRALIA

Leinster
M47710 Bismoclite

Mt Magnet
D36109 Bismutite
D36110 Bismutite

Poona
M49672 Russellite, bismutite; XRDMV

Warriedar
M35505 Ferrimolybdite
a
These were determined by XRD during the course of this study and are in addition to
those previously reported by Rankin (2001, 2002) Sharpe and Williams (2004).
b
Specimen numbers beginning with D are from the Australian Museum and those with
M from Museum Victoria. With respect to the latter, the note XRDMV refers to
identification by powder XRD by W.D. Birch (personal communication).

142 | P a g e

Você também pode gostar