Вы находитесь на странице: 1из 152

DOCTOR A L T H E S I S

Department of Civil, Environmental and Natural Resources Engineering


Division of Geosciences and Environmental Engineering

Mineralogical Controls on the Recovery of

Friederike Elisabeth Minz Mineralogical Controls on the Recovery of Antimony in Base-Metal Flotation
ISSN 1402-1544
ISBN 978-91-7583-545-7 (print)
ISBN 978-91-7583-546-4 (pdf)
Antimony in Base-Metal Flotation
Luleå University of Technology 2016
– Outlining the Framework of a Geometallurgical Model
for the Rockliden VHMS Deposit, Sweden

Friederike Elisabeth Minz

Ore Geology
Mineralogical Controls on the Recovery of Antimony in Base-Metal Flotation
– Outlining the Framework of a Geometallurgical Model
for the Rockliden VHMS Deposit, Sweden

Friederike Elisabeth Minz

Division of Geosciences and Environmental Engineering

Department of Civil, Environmental and Natural Resources Engineering

Luleå University of Technology

March 2016
Printed by Luleå University of Technology, Graphic Production 2016

ISSN 1402-1544
ISBN 978-91-7583-545-7 (print)
ISBN 978-91-7583-546-4 (pdf)

Luleå 2016

www.ltu.se
ABSTRACT

The results presented in this thesis are a case study on the polymetallic Zn–Cu volcanic-hosted massive sulphide
(VHMS) deposit at Rockliden. The Rockliden mineralisation is located in north-central Sweden, about 150 km south
of the Skellefte district. Exploration at Rockliden started in the 1980’s despite the relative remote location of the
deposit from processing facilities in the Skellefte ore district. During the first exploration period, flotation tests were
conducted which indicated that the locally high Sb grade in the mineralisation would increase the Sb content of the
Cu–Pb concentrate and lower its quality at the smelter. In 2007 exploration was resumed at the Rockliden site and
massive sulphide mineralisation was found to extend toward the depth. The average Sb grade in the mineralisation
decreases with depth; it is approximately 0.13 wt% above 400m below surface and only 0.06 wt% Sb below 400m.

The focus of this PhD study was to evaluate mineralogical controls on the distribution of Sb minerals in base-metal
flotation, which influences the Sb grade and hence the quality of the Cu–Pb concentrates. Modelling the Sb grade of
the Cu–Pb concentrates is expected to be helpful in flowsheet design, e.g. for decision making on when to use
hydrometallurgical treatment to reduce the Sb grade of this product prior to submission to the smelter.

The first part of the study was to characterise the massive sulphides and the immediate host rocks with special
reference to Sb mineralogy. The mineralogy and mineral associations were documented by optical and scanning
electron microscopy (SEM). Micro-analytical tools (such as wavelength-dispersive X-ray spectroscopy at an
electron probe microanalyser) were used to study the mineralogical distribution of major, minor and trace elements
(e.g., Zn, Cu, Fe, Sb). Sphalerite and chalcopyrite are the main base-metal minerals in the massive sulphides.
Antimony (Sb) is a minor to trace element bound in different minerals. The Sb mineralogy is complex; the most
common Sb-bearing phases include gudmundite and Cu- and Pb-bearing Sb sulphosalts; tetrahedrite, bournonite,
and meneghinite. Furthermore, the Sb minerals were found to occur partly locked with base-metal sulphides and
gangue in flotation products from initial flotation tests. Mineralogical parameters such as grain size, degree of
liberation and locking (mineral association in particles) were outlined as potential controls on the distribution of the
Sb minerals.

The second part of this study was to quantify mineralogical parameters influencing the distribution of the Sb
minerals in a laboratory flotation test. For this purpose composites blended from drill core samples were collected
and tested, and the flotation products were analysed with MLA measurement. The measurements were mass-
balanced by a particle tracking technique and the impact of the mineralogical parameters controlling the Sb
distribution was evaluated. The laboratory flotation test showed that Cu- and Pb-bearing Sb sulphosalts show
different flotation behaviour than gudmundite and that locking of gudmundite with Cu- and Pb-bearing minerals
influences its distribution. Further, chemical, bulk mineralogical and particle information was used to build Sb
distribution models based on a first-order kinetic process using HSC Chemistry Sim software. Chemical assays
alone were insufficient to simulate the Sb grades of flotation products, since they do not distinguish between the
various Sb minerals. The simulation indicated that a model using bulk mineralogy (e.g. collected by Scanning
Electron Microscopy based measurements) would be sufficient for estimating the Sb recovery and grade of the Cu–
Pb concentrate in the exploration or scoping stage of the Rockliden deposit.

The following components are required for a complete particle-based geometallurgical model of Rockliden: a
process-adapted geological model, a particle-breakage model, and a unit process model. Implementing full particle
information in the HSC Chemistry Sim software would improve the flotation model for Sb distribution, i.e. unit
process model. Further, suggestions on building a process-adapted geological model and a particle breakage model
are given based on the demands of a corresponding flotation model. The potential of the particle-based approach to
provide a holistic view on the deposit by connecting ore geology, mineral processing and process metallurgy aspects
of the Rockliden deposit was shown in this study.
ACKNOWLEDGEMENT

The study is part of the Work Package 1 (WP1, Geometallurgy and 4D geological modelling) of the Centre of
Advanced Mining and Metallurgy (CAMM1). The following institutions and companies are thanked for their
funding and financial support: CAMM1, Boliden Mineral AB, ProMinNET (research network for Process
Mineralogy and Geometallurgy, funded by Nordforsk), ALS Piteå and the Geometallurgy Laboratory, Technische
Universität Bergakademie Freiberg (TUBAF), Germany.

I would like to thank my supervisors Christina Wanhainen and Pertti Lamberg, for their scientific guidance, patience
throughout the study, careful and constructive corrections and suggestions of ideas; tack så mycket för strukturering
av alla projektetidéer & kiitos paljon lukuisista ideoista väitöskirjaani. Also, I want to thank Jonas Lasskogen
(Exploration Department, Boliden AB) and Nils-Johan Bolin (Division of Process Technology, Boliden AB) for
their supervision, for all assistance, suggestions, and contributions, especially for the assistance during field work
(Jonas) and giving his all for the conduction of the laboratory flotation test (Nils-Johan). Many thanks also to all
those who have initiated, guided and supported the work of this thesis: Pär Weihed (director of CAMM1, assistant
supervisor, LTU) and Jan Rosenkranz (leader of WP1 (CAMM1), LTU), Hein Raat, Benny Mattsson, Rodney Allen,
Pia Fagerström, Andreas Berggren, and Gunnar Agmalm (Boliden AB).

To the following people and their co-workers and personnel at stated laboratories I want to express my thanks for
their help with all practical, analytical and laboratory work involved in this thesis: Ulla-Britt Lundström and
personnel (Boliden drill core archive, Sweden), Rolf Danielsson, Peter Markström, Carina Andersson, Iris
Wunderlich and personnel (Boliden TMP laboratorium, Sweden), colleagues at the SEM laboratorium (Luleå
University of Technology, Sweden), Saku Varpenius and Leena Palmu (Center of Microscopy and Nanotechnology,
Oulu University, Finland), Kurt Aasly and Kjetil Eriksen (Department of Materials Science and Engineering, NTNU
Trondheim, Norway), Kai Bachmann and Bernhard Schulz (Geometallurgy Laboratory, TUBAF), Andreas Bartzsch
and the personal of the Erzlabor (Helmholtz Institute Freiberg for Resource Technology), David Berglund, Tony
Ökvist, Pekka Parvinen and Marke Oinonen (ALS Piteå, Sweden, and Labtium, Finnland).

Further, I want to thank all my colleagues at LTU, for all their help and the insight into their disciplines which was
needed for this study. In particular I want to thank: Tobias Bauer, Glenn Bark, Cecilia Lund, Mehdi Parian, Abdul
Mwanga, Pierre-Henri Koch and Samuel Awe. Milan Vnuk is thanked for help in formatting of articles (Paper II and
IV) and Riia Chmielowski for her readiness and her fast support through the language check of almost all written
work belonging to this thesis and Stephen Mayowa Famurewa for additional language corrections by the end of this
thesis. Also, comments and corrections of the anonymous reviewers to the published articles are acknowledged.

With a thankful heart I want to express gratitude to my parents and brothers for their support on my way to Luleå
and their patience with an often busy daughter and sister. Especially all my brothers and sisters in our Lord Jesus
Christ, I want to thank you for your accompany and encouragements.

Above all I want to thank God for the opportunity to start and finish third-cycle studies at LTU, for the
understanding He gave and all progress on the way; to know His strength, faithfulness, loving-kindness, grace and
peace that also took me through challenging times of this work. “The Lord is my rock and my fortress and my
deliverer; my God, my strength, in whom I will trust; my buckler, and the horn of my salvation, and my high tower”
(Psalm 18:2). To Him is all my praise!

Friederike Minz,

Luleå, March 2016


Appended papers in support of this thesis

I. “Detailed Characterisation of Antimony Mineralogy in a Geometallurgical Context at the Rockliden Ore


Deposit, North-Central Sweden” (Minz, F., Bolin, N-J., Lamberg, P., Wanhainen, C.), Minerals
Engineering (2013), Vol. 52, pp. 95-103.

II. “Lithology and mineralisation types of the Rockliden Zn–Cu massive sulphide deposit, north-central
Sweden: Implications for ore processing” (Minz, F. E., Lasskogen, J., Wanhainen, C., Lamberg, P.),
Applied Earth Science: Transactions of the Institutions of Mining and Metallurgy: Section B (2014), Vol.
123 (1), pp. 2-17.

III. “Distribution of Sb minerals in the Cu and Zn flotation of Rockliden massive sulphide ore in north-central
Sweden” (Minz, F. E., Bolin, N-J., Lamberg, P., Bachmann, K., Gutzmer, J. Wanhainen, C.), Minerals
Engineering (2015), Vol. 82, pp. 125–135.

IV. “Particle-based Sb distribution model for Cu–Pb flotation as part of geometallurgical modelling at the
Rockliden deposit, north-central Sweden” (Minz, F. E., Bolin, N-J., Lamberg, P., Wanhainen, C.,
Bachmann, K., Gutzmer, J.), manuscript to be submitted (to: Mineral Processing and Extractive
Metallurgy: Transactions of the Institutions of Mining and Metallurgy: Section C).
Related contributions not included in the thesis

x “Detailed Characterisation of the Antimony Mineralogy in a Geometallurgical Context at the Rockliden


Ore Deposit, Northern Sweden.” (Minz, F., Bolin, N-J., Lamberg, P., Wanhainen, C.), extended abstract
Process Mineralogy ’12 conference (Cape Town/ZA, 7th – 9th November 2012).

x “Geological Background and Qualitative Ore Characterisation for the Geometallurgical Project at
Rockliden, North-Central Sweden” (Minz, F., Wanhainen, C., Lamberg, P., Lasskogen, J., Raat, H.),
extended abstract & poster, SGA biennial meeting 2013 (Uppsala/SE, 12th – 15th August 2013).

x “Building a Geometallurgical Model in Iron Ores using a Mineralogical Approach with Liberation Data”
(Lamberg, P., Rosenkranz, J., Wanhainen, C., Lund, C., Minz, F., Mwanga, A., Amiri Parian, M.), paper
presented at GeoMet 2013 conference (Second AusIMM International Geometallurgy Conference 2013,
Brisbane/AUS).

x “Distribution of Sb-bearing minerals in the Cu and Zn flotation of Rockliden massive sulphide ore in north-
central Sweden.” (Minz, F. E., Bolin, N-J., Lamberg, P., Bachmann, K., Gutzmer, J., Wanhainen, C.),
extended abstract Process Mineralogy ’14 conference (Cape Town /ZA, 17th – 19th November 2014).

x “Mineralogical controls on the distribution of antimony in a base-metal flotation test at the Rockliden
massive sulphide deposit, north-central Sweden” (Minz, F. E., Lamberg, P., Wanhainen, C., Bolin, N-J.,
Bachmann, K., Gutzmer, J.), extended abstract & poster, SGA biennial meeting 2014 (Nancy/FR, 24th to
27th August 2015).

x “Mineralogical Characterisation of the Rockliden Antimony-Bearing Volcanic-Hosted Massive Sulphide


Deposit, Sweden” (Minz, F. E.), Licentiate thesis: Luleå University of Technology (2013).

Distribution of the work and author’s contribution

The following list gives the distribution of the work carried out for this thesis and contribution of the authors to the
papers appended in support of the thesis:

x Field work and data collection: F.E. Minz, J. Lasskogen, N-J. Bolin, personal at Boliden Mineral AB, C.
Wanhainen,

x Experimental and analytical work: N-J. Bolin, F.E. Minz, personal at different laboratories (cf. section 3),

x Data compilation and modelling: F.E. Minz, P. Lamberg, colleagues at the division of Minerals and
Metallurgical Engineering (LTU),

x Concept and writing of articles: F.E. Minz, P. Lamberg and C. Wanhainen,

x Review of manuscripts: C. Wanhainen, P. Lamberg, co-authors of the respective papers, R. Chmielowski


and anonymous reviewers.
General Abbreviations and Definitions

AAS – Atomic absorption VHMS – volcanic-hosted Cst – cassiterite


spectroscopy massive sulphide deposit
Dol – dolomite group
BSE – Back Scattered WDS – wavelength-dispersive
Electrons X-ray spectroscopy Epi – epidote group

EDS – energy-dispersive X-ray Fl – fluorite


spectroscopy
Mineral abbreviations Fsp – feldspar minerals
EPMA – electron probe
Gd – gudmundite
microanalyser

Amp – amphibole minerals Il – ilmenite


FESEM – Field Emission
Scanning Electron Microscope Mene – meneghinite
Anh – anhydrite
MASS – massive sulphides Mgt – magnetite
Ap – apatite
MLA – Mineral Liberation Phsl – phyllosilicates
Apy – arsenopyrite
Analyser
Bt – biotite Po – pyrrhotite
PCA – Principal Component
Analysis Boul – boulangerite Py – pyrite

RSQ – Pearson product- Bour – bournonite Qtz – quartz


moment correlation coefficient
Cal – calcite Ser – sericite
SEM – Scanning Electron
Microscope Carb – carbonate minerals Sp – sphalerite

XRD – X-ray diffraction Ccp – chalcopyrite Stnn – stannite

XRF – X-ray fluorescence Zn-Chr – Zn-rich chromite Ttn – titanite

Chl – chlorite group Ttr – tetrahedrite.

The term texture includes all parameters which are collectively found under petrographic terms: fabric, micro-
structure and texture (cf. Lund 2013). A distinction of scalar and vector mineral information is not made (cf.
Minz 2013). In this study emphasis is taken on process relevant parameters such as mineral association,
distribution, orientation, grain boundary relationship, grain size and shape (e.g. Butcher 2010; Walters and
Kojovic 2006).

The term ore comprises the massive sulphides and immediate host rocks forming the Rockliden deposit, which
potentially would be mined together depending on the size of theoretical mining blocks or smallest mineable
units. However, economic aspects are not further considered and the usage of the term ore does not strictly imply
a profitable extraction of valuable metals in this study.
TABLE OF CONTENT

1 Introduction................................................................................................................................................................ 1

1.1 Geometallurgical Context of the Study............................................................................................................... 1

1.2 Process Mineralogy ............................................................................................................................................ 3

1.3 Geometallurgy and Process Mineralogy of Base-Metal Deposits ...................................................................... 4

1.3.1 Copper deposits ........................................................................................................................................... 4

1.3.2 Zinc-lead(-copper) deposits (including VHMS deposits)............................................................................ 5

1.4 Aim and Scope of the Thesis .............................................................................................................................. 7

2 Background – Mineralogical Characterisation of the Rockliden Deposit .................................................................. 9

2.1 Ore Characterisation ........................................................................................................................................... 9

2.2 Process Mineralogy .......................................................................................................................................... 11

2.2.1 Distribution of Sb minerals in the ore body............................................................................................... 11

2.2.2 Expected distribution of Sb minerals in flotation process ......................................................................... 13

2.2.3 Implications of ore characterisation for automated mineralogy ................................................................ 15

3 Materials and Methods............................................................................................................................................. 16

3.1 Sampling, assays and microscopy..................................................................................................................... 16

3.1.1 Note on chemical assays as basis for element to mineral conversion........................................................ 18

3.2 Automated Mineralogy, Particle Tracking and Simulation .............................................................................. 18

3.3 Evaluation of Error Sources.............................................................................................................................. 19

3.3.1 Notes on sampling errors........................................................................................................................... 20

3.3.2 Notes on analytical errors.......................................................................................................................... 21

3.3.3 Data reconciliation and error sources in MLA measurements................................................................... 22

3.3.4 Notes on experimental and modelling errors............................................................................................. 24

4 Results – Quantification and Modelling of the Flotation Distribution of Sb ........................................................... 26

4.1 Flotation Test – Distribution of Sb Minerals .................................................................................................... 26

4.1.1 Characteristics of the composites .............................................................................................................. 26

4.1.2 Characteristics of the flotation products .................................................................................................... 33

4.2 Sb Distribution Model for Flotation ................................................................................................................. 37

4.2.1 Modelling of flotation rates and testing of different simulation options ................................................... 37

4.2.2 Distinction of Sb minerals in flotation feed............................................................................................... 40


5 Summary and Discussion ......................................................................................................................................... 42

5.1 Summary of Results.......................................................................................................................................... 42

5.2 Components for a Particle-Based Geometallurgical Model.............................................................................. 46

5.2.1 Process-related geological sub-model ....................................................................................................... 47

5.2.2 Flotation-related comminution model ....................................................................................................... 50

5.3 Model Alternatives ........................................................................................................................................... 53

5.3.1 Outlining alternative approaches ............................................................................................................... 53

5.3.2 Evaluation of increased precision in modelling using mineralogical data................................................. 54

6 Conclusions and Recommendations for Future Work.............................................................................................. 56

References................................................................................................................................................................... 59
1 Introduction

1.1 Geometallurgical Context of the Study

The term geometallurgy has been used in the literature for almost 50 years and a multitude of definitions and
descriptions are given (Jackson et al. 2011). A common concept in geometallurgy is the cooperation of the
disciplines involved in exploiting of mineral deposits (Jackson et al. 2011). Before the 20th century the cross-
disciplinary approach in mining and metallurgy was common practice (Jackson et al. 2011). For example, optical
microscopy was used in mineral processing studies (Chayes 1944). In the first half of the 20th century, the
management and organisation of large-scale mining has led into specialisation and separation between disciplines
(Jackson et al. 2011 and references therein). Moreover, another aspect of the decline in cross-disciplinary
communication is the fact that the usage of analytical tools, such as (ore) microscopy, is labour-intensive and
seldom quantitative (Walters and Kojovic 2006). However, mineralogical variations are critical to all steps of ore
processing, irrespective of the scale of mining (Baum 2014). The increase of mine sizes and production had required
automation of mineral analytical techniques. The development of SEM-based automated tools, capable of providing
detailed quantitative mineralogical information, started about 40 years ago (Lotter et al. 2014 and references
therein). In addition to the change to large-scale mining operations, the mining of complex and refractory ores has
increased (Lorenzen and Barnard 2011). Particularly for the processing of complex or refractory ores, the
determination of the complexity in the process response of such ores is helpful in evaluating the economic
exploitability of the deposit (Dunham and Vann 2007).

In the late 20th century, the advances of a cross-disciplinary approach were rediscovered (Jackson et al 2011 and
references therein). The potential benefits of a geometallurgical approach are numerous as described by Lamberg
(2011). For a geometallurgical approach, the ore bodies are outlined based on forecasted metallurgical performance
(Lamberg 2011). This allows the estimation of the value of an ore body spatially and more accurately. For example,
the variation in the plant feed can be controlled and minimised by more strategically blending or adjustment of plant
operation, including considerations about unconventional process solutions (Alruiz et al. 2009; Lamberg 2011;
Suazo et al. 2010). Using integrated measures of processing attributes in a spatial model gives opportunities to
optimise mine planning, to improve metallurgical performance, to reduce operational risk, and to make the overall
production more sustainable at the same time as socio-economic factors can be considered through increased
resource efficiency (Dunham and Vann 2007; Lund and Lamberg 2014; Parian 2015; Walters and Kojovic 2006).
Given the benefits of geometallurgy, geometallurgical programs were developed as industrial applications to cover
the stages from the early exploration through feasibility studies to the operational phase (Jackson et al. 2011;
Lamberg 2011).

The narrowest definitions of geometallurgy pertain to the subject of ore characterisation and process mineralogy, i.e.
studies without explicit spatial relation or coverage of the ore body (e.g. Helle et al. 2005; Minz 2013; Schouwstra et
al. 2010; Tungpalan et al. 2015). More advanced studies include geometallurgical mapping and modelling (e.g.
Alruiz et al. 2009; Jackson et al. 2011; Lamberg 2011; Walters and Kojovic 2006; Walters 2011). The broadest view
on the cross-disciplinary approach of geometallurgy incorporates also the non-technical aspects such as risk
management and economic evaluation, environmental and sustainability issues (Jackson et al. 2011).

Bulled and McInnes (2005) have defined the geometallurgical modelling as “an approach that measures
metallurgical variability within an ore body and quantifies the effect of this variability on the comminution and
metallurgical response of the ore”. Geometallurgical modelling requires that processing parameters are available
throughout the spatial extent, meaning a large number of samples need to be analysed (Keeney et al. 2011; Lamberg
2011; Parian 2015). This can be achieved by developing small-scale geometallurgical tests applicable to drill core
samples (e.g. Bulled and McInnes 2005; Mwanga 2014; Mwanga et al. 2015a; Walters and Kojovic 2006 and
references therein; Walters 2011). A statistical approach may be chosen for geometallurgical modelling, e.g. by
deriving formulas for throughput, grade-recovery curves or tailing quality, through regression functions (e.g.

1
Dunham and Vann 2007). Another option is to define geometallurgical domains of ore bodies using principal
component analysis (PCA) of geological databases and assigning processing parameters to the domains (e.g. Alruiz
et al. 2009; Keeney et al. 2011). Besides standard laboratory tests for measuring variability in process response of
geometallurgical domains, Lotter et al. (2011) recommended conducting tests on smaller samples throughout a
geometallurgical unit in order to measure the range in process response within a unit and thus to verify the validity
of the definition of units.

In the geometallurgical research group at Luleå University of Technology the focus is on practical amalgamation of
ore geology, process mineralogy and minerals processing (Lamberg 2011; Lund 2013; Mwanga 2014) in order to
build a model which can be used in production planning and management (Lishchuk et al. 2015a; Lishchuk et al.
2015b; Parian 2015).

The scope of a geometallurgical program may be defined on the framework or background of an individual project
(Jackson et al. 2011). In the Rockliden case, the model should have the capability to forecast the product quality in
terms of grades and recovery of penalty elements, in particular of antimony (Sb) in the Cu–Pb concentrate, over the
spatial extent of the Rockliden deposit (Fig. 1.1). Two aspects were considered: (1) the occurrence and variability of
Sb in minerals (including textural factors such as mineral association), and (2) the distribution of Sb minerals in the
flotation process. These aspects focus on two of the geometallurgical program points suggested by Lamberg (2011):
first, ore characterisation for sample selection and, second, laboratory testing of selected samples in order to extract
process model parameters (see also Table 5.1).

Initial exploration started in the 1980s at the Rockliden site and comprised of a drilling campaign and flotation tests.
Due to the remote location of Rockliden and the metallurgical challenges, high investment costs were expected and
the exploration activities were stopped (Raat and Årebäck 2009). Despite these facts, a new exploration campaign
started in 2007, with the metallurgical challenges remaining. Relatively high Sb grades had been noticed in parts of
the Rockliden massive sulphide deposit and in the Cu concentrate (Minz 2013). Generally, antimony (Sb) grades are
known to affect the smelting and refining process. For example, the occurrence of Sb in the matte has negative
effects on the copper electro-refining process e.g. through contamination of the copper cathodes by floating slimes
(Awe et al. 2012 and references therein; Larouche 2001; Navarro and Alguacil 2002; Wang 2004). Also, Sb affects
the properties of metallic copper, e.g. by lowering the ductility of refined copper %DOiåDQG $FKLPRYLþRYi
and references therein; Knorr and Breedis 1986). Due to these unwanted effects, Sb is referred to as a penalty
element constraining the quality of Cu concentrates and grades higher than ca. 0.2 wt% Sb can cause penalties at Cu
smelters (Larouche 2001). In case of a high Sb content in the Cu concentrate, unconventional process solutions are
considered, including hydro- and pyrometallurgical treatment. Compared to arsenic (As), Sb shows a lower
distribution into offgas during smelting but the distribution of Sb into matte phases is double to three times higher
than for As (Awe 2013and references therein; Björnberg et al. 1986). This indicates the removal of Sb will be less
effective than the removal of As by pyrometallurgical treatment. The removal of Sb by hydrometallurgical treatment
of the Cu concentrates, e.g. alkaline sulphide leaching, has been shown to be promising, since it is capable to
dissolve all Sb minerals in Cu–Pb concentrates from Rockliden (Awe et al. 2012).

Given the above considerations, the following disciplines are involved in the Rockliden study: ore geology, process
mineralogy, mineral processing and metallurgy. Knowledge from these disciplines forms the main academic
framework of this study.

2
Fig. 1.1. Simplified process chain including unconventional process solutions (for a more detailed flowsheet see figure 3.3).

1.2 Process Mineralogy

Process mineralogy is a branch of applied mineralogy which focus on studying mineralogical and petrographic
variations in the ore feed (also referred to as ore characterisation) and the effect of this variation on the metallurgical
performance (Baum 2014; Jackson et al. 2011; Petruk 2000). Generally, metallurgical parameters are functions of a
variety of geological and mineralogical parameters, such as (mineral) grades, ore textures, the association of ore and
gangue minerals, mineral chemistry, and element deportment (e.g. Ashley and Callow 2000; Lorenzen and Barnard
2011). In addition, petrophysical properties, such as rock strength (hardness, abrasion, etc.), grain size, ‘dustability’,
grindability and liberability, and specific gravity of ore and host rocks and smelting characteristics, might be
included in process mineralogy studies, depending on the framework of an individual study (Dunham and Vann
2007; Mwanga et al. 2015b; Walters and Kojovic 2006).

Process mineralogy studies are important in designing a process and to optimise it (Baum 2014; Lorenzen and
Barnard 2011; Lotter et al. 2014; Petruk 2000). The process diagnosis or problem-solving approach for poor or
unexpected metallurgical outcomes is a key aspect of process mineralogy (Lotter et al. 2011). It should be noted that
process mineralogy does not provide a direct spatial context to an ore deposit (Jackson et al. 2011). A spatial context
to the results of a process mineralogical study could be inferred if results are coupled to an ore type definition of a
geological model. However, process mineralogy can have economic benefits in itself, for example by determining

3
the optimal grinding size for flowsheet design and improving the forecast of process models based on quantitative
mineralogical data (Baum et al. 2004; Gu et al. 2014; Lotter et al. 2011; Lotter et al. 2014).

A range of analytical techniques are available to determine mineralogical parameters. Automated X-ray diffraction
(XRD) and automated mineralogy based on scanning-electron microscopy (such as MLA, QEM*SEM/QEMSCAN
and INCA Feature analysis) are used in the mining industry (Baum 2014; Gottlieb et al. 2000; Gu et al. 2014; Liipo
et al. 2012; Parian et al. 2015; Petruk 2000). For the selection of appropriate analytical methods and (laboratory)
tests, several aspects might need to be considered such as cost-efficiency, time for data acquisition and processing,
reproducibility and reliability of measurements on small sample sizes (e.g. Baum 2014; Parian et al. 2015; Walters
and Kojovic 2006). The economic aspects become critical especially when the geometallurgical context is
considered, since a large number of samples might need to be analysed or tested to achieve a link between
(mineralogical) parameters and the metallurgical response, considering the mineralogical variability of an ore
deposit (Lamberg 2011; Parian et al. 2015 and references therein).

1.3 Geometallurgy and Process Mineralogy of Base-Metal Deposits

The following chapter gives a summary for case studies which consider different aspects of ore characterisation and
process mineralogy for Cu, Zn and Pb ore deposits and some of which have a more extensive geometallurgical
context. As this thesis presents results of a case study for the Rockliden Zn–Cu volcanic-hosted massive sulphide
(VHMS) deposit, this literature review is intended to give an insight into research approaches chosen in other studies
and it positions the Rockliden study in the respective research area.

Studies concerning geometallurgical modelling which cover the whole mining and processing chain are not so
common, and mostly pertain to large scale operations, e.g. for Cu porphyry deposits (e.g. Tungpalan et al. 2015).
Process mineralogical studies focusing on mineralogical complex massive sulphide ore are more common (e.g.
Lastra 2007; Petruk and Hughson 1977). Additional to deposit case studies, research on development of analytical
approaches to record parameters relevant to mineral processing is presented in few studies (e.g. Hunt et al. 2011).

1.3.1 Copper deposits

Porphyry deposits are the most important sources for Cu, accounting for about 60% of the world Cu production. Cu
porphyry deposits are low grade – high tonnage deposits with a typical Cu grade of about 0.2 wt% (Cropp et al.
2013 and references therein). Cropp et al. (2013) describe general mineralogical and textural parameters which can
influence the flotation of Cu minerals. Since Cu porphyries are low grade deposits the gangue mineralogy influences
to a high degree the flotation performance, but also the comminution characteristics such as throughput (Cropp et al.
2013). Alruiz et al. (2009) proposed a comminution model based on characteristics of geometallurgical units, such
as ore grindability (derived from comminution tests on a large number of samples), to simulate the throughput of the
Collahuasi grinding circuit. Suazo et al. (2010) built a model to calculate the metallurgical results of the Collahuasi
flotation circuit using inherent floatability as a characteristic of geometallurgical units. In a case study for a
porphyry and a porphyry associated skarn deposit, the application of modelling software for ”integrated simulation
of comminution, classification and flotation” was demonstrated for quick estimation of product quality of a flotation
plant (Tungpalan et al. 2015). Finally, Helle et al. (2005) showed the relation between mineralogical characteristics
of geometallurgical units and Cu recovery by acid leaching. The above list of geometallurgical research covers
almost all processing aspects of Cu porphyry deposits.

Besides Cu porphyry deposits, IOCG, Cu-Ni and Cu-shale deposits are also relevant Cu sources. Examples of
process mineralogical and geometallurgical research for Cu shale deposits are given by Lotter et al. (2014) and Van
den Boogaart et al. (2011). Hunt et al. (2011) demonstrated the usage of artificial segmentation algorithms on

4
mineral maps in order to determine potential liberation behaviour of samples from an IOCG (iron oxide copper
gold) deposit. Bonnici (2012) investigated and adapted microscopic techniques, using high resolution digital images
and image analysis, as tools to quantify mineralogical and textural attributes relevant to processing behaviour of
minerals exemplified for IOCG and Cu porphyry deposits. Lotter et al. (2011) conducted process mineralogical
research at the Cu–Ni Nickel Rim South deposit (Sudbury district, Canada) and the related Raglan concentrator. The
applicability of process mineralogy, including benchmark plant surveys using QEMSCAN data, was demonstrated
in the development of blending, grinding and flotation strategies (Lotter et al. 2011).

1.3.2 Zinc-lead(-copper) deposits (including VHMS deposits)

The following studies highlight process mineralogical and geometallurgical aspects of Zn-Pb (Cu) massive and
semi-massive sulphide deposits.

McClung and Viljoen (2012) underlined the importance of mineralogical and geometallurgical investigations in
mineralogical complex ore at the Zn–Pb Broken Hill deposit, South Africa. The significance of mineralogy (and
mineral chemistry) for the explanation of the quality of the flotation products was demonstrated by an element
deportment study at the Gamsberg Zn deposit (South Africa); comprising QEMSCAN analysis, microprobe data on
sphalerite and Time of Flight Secondary Ion Mass Spectrometer (ToF-SIMS) (Schouwstra et al. 2010). Leaching
characteristics of a Gamsberg pyrite-quartz-sphalerite ore sample were studied by Ghorbani et al. (2013) using X-
ray tomography and QEMSCAN measurements. A connection between mineral association and mineral chemistry
was shown (Ghorbani et al. 2013). These three studies present results on different process-mineralogical and
geometallurgical aspects of the Zn–Pb deposits, however, give no direct proposal for a complete geometallurgical
model.

A detailed ore characterisation and process mineralogy study was conducted for the George Fisher Zn–Pb sediment-
hosted sulphide ore deposit close to the Mount Isa deposit, Australia (Bojcevski et al. 1998; Bojcevski 2004).
Mineral textures were defined on meso-scale (centimeter level, usable for drill core logging) and on micro-scale (by
using optical microscopy to describe mineral association, grain sizes, etc.). Incorporating meso-texture information,
in addition to chemical assays, into an empirical flotation model allowed more precise simulation of the recovery of
base-metals related to improved correlation. Assuming metallurgical characteristics of the outlined ore textures were
found consistent throughout the ore deposit, it was suggested to infer metallurgical characteristics of ore types from
meso-textures of a geological model, e.g. for mine planning (Bojcevski 2004). The study conducted by Bojcevski
(2004) provided the most complete concept of a geometallurgical model for a Zn–Pb ore deposit. In contrast to the
empirical (or statistical) approach connecting ore characteristics with flotation results, Gorain et al. (2000) used
liberation analysis (derived from QEM*SEM analysis) to measure ore characteristics of the flotation feed and to
model chalcopyrite recovery from Mount Isa ore.

Bolin et al. (2003) described the gradual improvement of the Garpenberg concentrator treating ore from Zn–Pb–Ag–
(Cu–Au) sulphide deposits in central Sweden. A mineralogical study on flotation products helped to find options for
improvement of the existing operation, e.g. suggestions were given to increase sphalerite liberation by careful
regrinding in order to avoid coating of sphalerite with oxyhydroxides.

Petruk and Hughson (1977) conducted a process mineralogical study on ore from the Zn–Pb–Cu volcanic-hosted
massive sulphide (VHMS) deposits at the Matagami and Brunswick mines, Canada. Petruk and Schnarr (1981)
studied Brunswick concentrator products. Chemical assays, XRF analysis, optical microscopy and image analysis
with Quantimet 720 Image Analyzing system were used for mineralogical and textural characterisation and for
comparison of flotation products. Regrinding was recommended to improve sphalerite, chalcopyrite, galena and
tetrahedrite liberation and recovery (Petruk and Schnarr 1981). However, advanced regrinding on the other hand
caused sliming of galena in the Brunswick ore (Petruk and Hughson 1977). In his book “Applied Mineralogy in the
Mining Industry”, Petruk (2000) summarised the ore characterisation of several VHMS deposits Flin Flon-Snow
Lake areas, Manitoba, Canada. Descriptions of ore texture were mostly qualitative and based on ore microscopy.

5
Microprobe analyses were conducted for e.g. trace minerals. By combining ore characterisation with chemical
analysis of the flotation products, suggestions for improvement in processing were made.

Benzaazoua et al. (2002) conducted a detailed EPMA/WDS study on all minerals in the Cu-rich feed of the VHMS
deposit of Neves Corvo, Iberian Pyrite Belt, Portugal. The detailed mineral chemistry (EPMA analysis) served as
basis for element to mineral conversion. High mineral liberation was assumed. A laboratory test was conducted
using different flotation reagents and the bulk mineralogy of all products derived from chemical assays. Some key
findings of this process mineralogical study were that: firstly, a differential flotation behaviour for fahlores
(tetrahedrite–tennantite series) was found; secondly, principal component analysis (PCA) showed that a higher
mineral separation could be achieved; and thirdly, a correlation between major elements or minerals and trace
elements was established and was found potentially useful for prediction of trace element content of products from
XRF analysis in slurries in the processing plant. Beyond the traditional approach to processing of ore material, Hunt
et al. (2016) studied the mineralogy of gossanous waste rock material from the polymetallic VHMS deposits of the
Iberian Pyrite Belt, Portugal and Spain. Information from the mineralogical studies was used to give advice on the
extraction of precious metals from this material in order to increase the exploitability of the resources, and at the
same time taking into consideration environmental issues.

A detailed ore characterisation and geological modelling work was done on the Hellyer Zn–Pb–Cu VHMS deposit,
Tasmania (McArthur 1988; McArthur 1996). McArthur (1988) adapted kriging technique, using variography within
stratigraphic layers, to model the metal grade distribution for early mine planning studies. In the PhD thesis,
McArthur (1996) described micro-textures in detail and established a coding system based on optical microscopy.
Microprobe analyses were conducted to identify trace minerals and to determine the chemistry of the main sulphide
minerals and tetrahedrite over the spatial extent of the Hellyer deposit (McArthur 1996). It was suggested to embed
mineralogical parameters, such as grain size and mineral association, in a block model, which could then assist in
mine planning and forecast process performance (Walters and Kojovic 2006). For example, attributes such as rock
hardness, mineral grade, and textures, are expected to have an impact on mill operation (Walters and Kojovic 2006).

Knight et al. (2011) built a geometallurgical model for an unnamed Zn–Cu VHMS deposit in the prefeasibility stage
of a mining project. A uniform mineralogy and mineral chemistry was assumed for the deposit. Two locked cycle
flotation tests were used to estimate the Cd or Sb+As content of the flotation products. Based on this information
iso-surfaces were constructed and areas where these contents were estimated to be above the threshold for smelter
penalties were outlined. The variability in flotation response within the margins of iso-surface remained unknown
since the precision of this approach had not been shown (cf. Lotter et al. 2011). Despite these shortcomings, Knight
et al. (2011) provided the most comprehensive study on geometallurgical modelling of a VHMS ore deposit.

In addition to the research on Sb minerals in VHMS deposits presented by Benzaazoua et al. (2002), Lager (1985)
conducted a mineralogical study distinguishing several Sb minerals at the Rakkejaur massive sulphide deposit,
Skellefte district (northern Sweden). The behaviour of the Sb minerals was studied in a flotation test at a pilot plant
(Lager and Forssberg 1990), and a SEM-based mineralogical study was done to record the Sb mineralogy of the
products. Additional mineralogical factors such as degree of liberation were not investigated (Lager and Forssberg
1990). However, it was shown that the grain size of Sb minerals can be as low as 30 μm and complex intergrowths
were documented in the ore (Lager 1985), which would require very fine grinding to completely liberate the Sb
minerals. Advances in the analytical techniques make it possible to record mineralogical factors in addition to the
bulk mineralogy, such as the degree of liberation for particles in flotation products in the Rockliden case study
(Paper III).

In summary it can be stated that massive sulphide deposits in particular have a complex mineralogy and vast variety
of textures, which are found to influence the processing behaviour. Furthermore, it is known that these deposits can
host a range of deleterious trace minerals (cf. Minz 2013). Geometallurgical or process mineralogical studies on the
trace mineralogy of massive sulphide deposits and the implications for mining and quality of processing products
are few (e.g. Benzaazoua et al. 2002; Lager and Forssberg 1990). However, the usage of detailed mineralogical
information in modelling of flotation processes is known for complex PGE deposits, requiring consistent analysis
with SEM-based automated tools such as MLA analysis (e.g. Chetty et al. 2009).

6
1.4 Aim and Scope of the Thesis

The main aim of this study is to describe the mineralogical variation of the Rockliden deposit, with focus on Sb
minerals, and with the purpose of outlining and extracting rock-intrinsic parameters influencing the distribution of
the Sb minerals during flotation. Further, it was aimed to determine whether (and which) mineralogical and particle
information can be used to model the Sb grade of the Cu–Pb concentrate precisely enough for implementation into a
geometallurgical model in the exploration stage of a deposit. Suggestions are given on how quantitative
mineralogical information – focusing on the Sb mineralogy – can be added to a resource model in the current stage
of the Rockliden exploration project.

The project is divided into two parts, which are described in the following:

Part 1: Ore characterisation and Sb mineralogy (Paper I & II, Minz 2013, summarised in section 2)

Ore characterisation has been defined as a fundamental component in the geometallurgical approach used at LTU
(e.g. Lund 2013). The Rockliden ore characterisation study comprises petrographic description of selected drill core
sections and samples in order to classify host rocks and massive sulphides using geochemical and mineralogical
classification tools. This includes also a detailed characterisation of the complex mineralogy, both the main and Sb
mineralogy in the ore and flotation products, using qualitative microscopic studies (SEM/BSE images) and
microanalysis (SEM/EDS and EPMA/WDS). Results from the ore characterisation study (part 1) assisted in the
sample collection for the laboratory flotation test and provided basic information for automated mineralogy analysis
in the continuation of the study (part 2, cf. Lotter et al. 2003).

Part 2: Extraction test and Sb distribution model (Paper III & IV, section 3 to 6)

Variability testing is used to extract parameters needed for modelling of the process (Lamberg 2011).
Mineralogically distinct composites (end-member samples in terms of main and Sb mineralogy) were used as feed to
a simplified laboratory flotation test to evaluate variability in process response (cf. Lamberg 2011). Automated
SEM-based measurements were conducted to document the mineralogy of the flotation products using the Mineral
Liberation Analyser (Fandrich et al. 2007). A particle tracking technique was applied to automated SEM-based
measurements in order to study the distribution of locked and unlocked Sb minerals in the flotation products
(Lamberg and Vianna 2007). Using the quantitative mineralogical data, the flotation distribution of Sb was
simulated. Different levels of mineralogical information were used for simulation. The aim was to study the depth of
mineralogical information needed, i.e. which rock-intrinsic parameters are necessary to simulate the Sb distribution
in different flotation products accurately enough to be usable for a geometallurgical model in the exploration stage
of the Rockliden deposit. Based on the relevance of parameters, suggestions were made on how to collect (spatial)
quantitative mineralogical information and how to implement such information in a process-adapted geological
model for the Rockliden volcanic-hosted massive sulphide (VHMS) deposit (Fig. 1.2, Fig. 5.2).

Figure 1.2 gives the general outline of this study and Table 1.1 discloses the objectives of the two parts of this study.
Results presented in this thesis focus on the second part the Rockliden PhD study.

7
Fig. 1.2. Flow chart showing the structure of the Rockliden PhD project. Grey outlined are the part 1 and black outlined the
second part of the study.

Table 1.1

Objectives within the aim of this study.

Article Objectives

Paper I x Detailed description of the Sb mineralogy and the mode of occurrence of Sb minerals in the
Rockliden ore,
x Qualitative characterisation of flotation products to outline mineralogical (rock-intrinsic) parameters,
potentially controlling Sb flotation distribution,

Paper II x Compilation of knowledge of the local geology and mineralogy of the Rockliden deposit,
x Detailed petrographic and mineralogical description of the different rock and mineralisation types,
x Qualitative classification of the massive sulphide ore based on the grade of the major sulphide
minerals and micro-textures,

Paper III x Quantification of the Sb mineralogy in the ore and in the products of a laboratory flotation test,
x Defining mineralogical factors controlling the distribution of Sb between different flotation products,

Paper IV x Simulation of the Sb grade and recovery in the Cu–Pb concentrate based on Sb mineralogy of the
feed and flotation products,
x Testing whether a particle-based model (i.e. using particle-tracked MLA measurements) provides
more precise simulations than using bulk mineralogy or chemical assays.

8
2 Background – Mineralogical Characterisation of the Rockliden Deposit

2.1 Ore Characterisation

The inferred mineral resource of the Rockliden massive sulphide ore is 9.2 Mt with 4.0 wt% Zn, 1.8 wt% Cu,
0.4 wt% Pb and 48 g/t Ag (New Boliden 2015). Besides bonus trace elements such as Ag, the Rockliden deposit
contains deleterious trace elements such as As, Hg and Sb. An average Sb content was estimated from drill core
assays to be 0.1 wt%. In this study the characterisation of the massive sulphide ore bodies and the immediate host
rock is based on drill core logging, assaying, optical microscopy and SEM (scanning electron microscopy). This
chapter summarises the results of Paper I and II.

The main rock types of Rockliden are metasedimentary and volcanic rocks enveloping the steeply dipping massive
sulphide ore. Mafic dykes are cross-cutting the deposit. A summary of all rock and massive sulphide types is given
in Table 2.1. Based on the total sulphide content (TS) of drill core samples, the following broad distinction can be
made: disseminated (generally < 15 wt% TS), semi-massive (ca. 15 to 50 wt% TS), and massive sulphides (> 50
wt% TS). The drill core samples with disseminated sulphides comprise the silicate-dominated host rock, which may
contain minor amounts of sulphide minerals when it is altered. An extensive alteration halo is present in the volcanic
rocks, which transgress through semi-massive into massive sulphides in structurally undisturbed drill core profiles.
The main sulphide minerals are pyrite (Py), sphalerite (Sp), chalcopyrite (Ccp) and pyrrhotite (Po). Based on the
main mineralogy six massive sulphide types have been distinguished (Table 2.1). These six types may be grouped
based on their chemical assays (the only quantitative information systematically collected for drill core intervals)
into two classes: one Cu-dominated (Cu/Cu+Zn > 0.5), the other Zn-dominated (Cu/Cu+Zn < 0.5). The Cu-
dominated class comprise largely massive sulphides of type 5 and 6 while the Zn-dominated class contains the
massive sulphides of type 2, 3 and 4 (Table 2.1).

Table 2.1

Characterics of host rocks and massive sulphides (MS) types. Mineral abbreviations: Amp – amphibole minerals, And –
andalusite, Anh – anhydrite, Ap – apatite, Apy – arsenopyrite, Bt – biotite, Cal – calcite, Carb – carbonate minerals, Ccp –
chalcopyrite, Zn-Chr – Zn-rich chromite, Phsl – phyllosilicates, Chl – chlorite group, Cst – cassiterite, Epi – epidote group, Fl –
fluorite, Fsp – feldspar minerals, Il – ilmenite, Mgt – magnetite, Po – pyrrhotite, Py – pyrite, Qtz – quartz, Rt – rutil, Ser –
sericite, Sp – sphalerite, Stnn – stannite, Ttn – titanite, Ttr – tetrahedrite.

Rock type Mineralogy Texture

Turbiditic Phsl, Qtz. Alternations of shale and fine-grained


sediments and sandstone; locally silicification in contact with
greywackes massive sulphides.

Felsic volcanic Fsp, Qtz, Ser (minor and trace minerals in altered Porphyritic with Fsp and Qtz phenocrysts,
rocks volcanics: Phsl (Chl), Fe-ghanite, Epi, And, Carb, foliation defined by alignment of
Anh, Fl; Py, Po, Sp, Ccp), phyllosilicates; alteration types: sericite and
geochemical classification: rhyolitic–dacitic. sericite-quartz alteration, chlorite alteration.

Mafic Rocks Amp, Fsp, Phsl, Bt, (minor and trace minerals: Carb, Mafic dykes: Amp (composition of hornblende,
(mafic dykes) Qtz, Zn-Chr, Rt, Il, Ttn, Ap, Cal), grain size: 50 to 500 μm), Phsl, Bt, and Plg
trace chemistry of mafic dykes: Cr > 300 ppm and V forming a fine-grained groundmass (no distinct
> 150 ppm (Zn-Chr grains surrounded by Cr–V-rich chilled margins in contrast to dolerite dykes (cf.
Mgt rims or Il rims). Raat et al. 2011), locally with Cal-Qtz-Chl
veinlets.

9
Table 2.1 (cont.)

Massive sul- Mineralogy Texture


phides (MS)

MS – Type 1 Po–Sp–Ccp (non-opaque minerals: Carb, Qtz, Ser, Locally, with isolated grains of pyrite (up to 500
Phsl, Fsp, Amp), mm diameter).
high grades (> 1000 ppm) of Sn (tin) locally in clast-
rich, isolated (max. 2 m width), massive sulphide
intersections (Sn mineralogy: Cst and Stnn).

MS – Type 2 Py (non-opaque minerals: Qtz, Carb, Phsl), 50–500 μm grain size variation of pyrite, locally
Ȉ &X=Q FRQWHQWZW. high density packing of pyrite grains,
irregular-shaped Ccp±Po meshes if Cu > 2 wt%,
Mgt grains up to 200 μm in diameter

MS – Type 3 Py–Sp (non-opaque minerals: Carb, Qtz, Phsl, Ser, Py grains (200 to 500 μm, max 1000 μm
Bt), diameter) often show embayments of Sp, Py
Zn > 5 wt% (ranging from 7 – 15 wt%), Cu < 2 wt%. with inclusions of Sp (locally also Gn and Ttr)
Po in diablastic intergrowth in the Sp
groundmass

MS – Type 4 Py–Sp–Ccp (non-opaque minerals: Qtz, Carb, Ser, With increasing Po content, Po and Sp can
Phsl, Amp), alternate in ground mass around Py grains and
Zn > 5 wt-%, Cu > 2 wt%. form banded textures.

MS – Type 5 Py–Ccp–Sp (non-opaque minerals: Qtz, Carb, Phsl, Locally close association of Apy–Po–Ccp (tight
Ser, Amp), intergrowth),
Zn < 5 wt-%, Cu > 2 wt%. high Ccp content tend to correlate with
relatively high Po content.

MS – Type 6 Py–Po–Mgt–Ccp–Sp (non-opaque minerals: Phsl, Similar as the Py–Ccp–Sp group (MS type 5)
Qtz, Carb, Amp). but with high Po or Mgt content.

The Sb minerals were studied in detail as a critical component of the minor and trace mineralogy of Rockliden. SEM
studies showed that there is a great variety of Sb minerals found at Rockliden, out of which tetrahedrite, bournonite,
meneghinite, boulangerite and gudmundite are commonly found in the Rockliden deposit (Minz 2013, Fig. 2.1). The
mineral chemistry of the main Sb minerals has been studied in detail with EPMA/WDS analysis (Table 2.2). Except
for tetrahedrite, the Sb minerals are close to their stoichiometric compositions (cf. Paper I). The Ag content of
tetrahedrite was found to be variable, being highest in the drill core samples from Po–Py–Sp massive sulphides
(Mattsson and Heeroma 1985) and mafic dykes. The variation in Ag content is probably related to the host rock, i.e.
tetrahedrite found in altered felsic volcanic rocks have an Ag content below 4 wt% and tetrahedrite found in mafic
dykes generally have an Ag content above 4 wt% (Minz 2013). The distinction between two types of tetrahedrite is
shown in Table 2.2.

10
Table 2.2

Mineral composition of the common Sb minerals at Rockliden according to EPMA/WDS analysis.

Mineral and No. of Ag / Zn / Cu / S/ Fe / Sb /


stoichiometric formula analysis Stat.c) wt% wt% wt% wt% wt% wt% Pb / wt%
(a
Tetrahedrite 6 AM 1.62 2.09 36.21 24.98 5.59 28.95 -
(Cu,Fe,Ag,Zn) 12 Sb 4 S 13 SD 1.31 0.25 0.90 0.37 0.33 0.60 -
Tetrahedrite(b 23 AM 7.60 1.73 32.28 24.49 5.85 27.60 -
(Cu,Fe,Ag,Zn) 12 Sb 4 S 13 SD 2.51 0.40 1.89 0.65 0.51 0.93 -
Bournonite 25 AM - - 13.34 19.68 0.46 23.34 42.67
PbCuSbS 3 SD - - 0.25 0.41 0.31 1.64 1.63
Meneghinite 20 AM - 0.44 1.56 17.36 0.47 18.72 60.86
Pb 13 CuSb 7 S 24 SD - 0.36 0.29 0.17 0.26 0.40 0.69
Boulangerite 8 AM - 0.50 - 18.43 0.39 24.19 55.84
Pb 5 Sb 4 S 11 SD - 0.52 - 0.11 0.28 0.77 0.46
Gudmundite 39 AM - - - 15.67 28.21 55.58 -
FeSbS SD - - - 0.64 0.62 1.33 -
- - LLD 0.05 0.21 0.16 0.04 0.1 0.06 0.19
a) tetrahedrite in altered felsic volcanic rock and part of the massive sulphides,

b) tetrahedrite in mafic dykes and part of the massive sulphides,

c) statistics (AM – arithmetic mean, SD – standard deviation, LLD – lower limit of detection (EPMA/WDS analysis)).

2.2 Process Mineralogy

2.2.1 Distribution of Sb minerals in the ore body

The distribution of Sb minerals in the deposit, as estimated by using both optical microscopy and SEM, shows that
tetrahedrite and gudmundite occur evenly distributed, with a slightly higher abundance in the upper part of the
deposit (Fig. 2.1). Meneghinite and boulangerite tend to show peak concentrations in rather narrow intervals, a
feature that might be related to their preferred occurrence in mafic dykes (see also section 4).

Systematic differences are recognized between the Sb mineralogy of Sb-rich rock types and of massive sulphides.
The Sb mineralogy of the mafic dykes is dominated by Cu- and Pb-bearing sulphosalts, while the acid volcanic
rocks contain mostly gudmundite with Pb–Sb sulphosalts, locally associated with quartz and calcite veinlets. No
systematic differences are noticed within the massive sulphides. A correlation between main mineralogy (or main
elements) and Sb as trace element (cf. Benzaazoua et al. 2002) could not be established for Rockliden massive
sulphide samples and this might be related to the complexity in the Sb mineralogy at Rockliden. However, the Sb
mineralogy seems to change towards depth in the Zn-dominated massive sulphides (Paper I). The SIMCA software
was used to conduct a principal component analysis (PCA) on the Sb mineralogy of the massive sulphides (32
samples). The Sb mineralogy of each sample was estimated from combined optical microscopy and SEM-based
analysis. It should be noted that the number of samples is too low for a proper PCA model. A provisional two
component PCA model using assays for the base metals and the fraction of Sb minerals (Sum to 100%) gives the
loading plot shown in Fig. 2.2a. Additionally, a two component PCA model is shown in figure 2.2b using Sb grade
differentiated by minerals instead of fractions of Sb minerals.

11
Fig. 2.1. Variation in the grade (wt%) of the main Sb minerals (Bour – bournonite, Gd – gudmundite, Mene – meneghinite
(including boulangerite), Ttr – tetrahedrite) based on the estimated proportion of Sb minerals in drill core samples collected
during 2012 and 2013 (length of sampled intervals: 0.3 to 1.0 m), in relation to rock types (MS-Cu/-Zn – Cu- and Zn-dominated
massive sulphides, MD – mafic dykes, volc – felsic volcanic rock). Viewing to the west with the Rockliden massive sulphide ore
body as translucent shapes.

12
Fig 2.2. Loading plot of a 2 component PCA model using SIMCA modelling software (Eriksson et al. 2006). Besides drill core
assays, graphic (a VKRZVPLQHUDOUDWLRVRIWKH6EPLQHUDOIUDFWLRQ  Ȉ 7WU%RXU0HQH*G , and graphic (b) shows
Sb grades hosted by tetrahedrite (Ttr), bournonite (Bour), meneghinite (Mene) and gudmundite (Gd). The Sb mineralogy was
estimated from combined optical microscopy and SEM.

It should be noted that the second component of the PCA models in figure 2.2a and figure 2.2b is not significant, the
R2X (degree of explanation) for both components is lower than 0.5 and the Q2 (predictablility) value is negative or
close to cerro. Further, the PCA model in figure 2.2b gives no distinct separation of mineral specific Sb grades, i.e.
grades based on the Sb distribution to the host minerals (tetrahedrite, bournonite, meneghinite and gudmundite),
along the first componenet and is thus discarded from further discussion. However, a clustering of the fraction of
bournonite and meneghinite with the Pb, Zn and Sb assays and the fraction of tetrahedrite and gudmundite with Cu,
As and S content is observed in figure 2.2a. Such clustering is supported by the mineral association as observed in
qualitative microscopic studies; the Cu- and Pb-bearing sulphosalts (tetrahedrite, bournonite, meneghinite and
boulangerite) are commonly associated with each other and with sulphide minerals, whereas the mineral association
of gudmundite is rather widespread (Paper I, Table 2.3). Only the very fine grained gudmundite is associated
systematically with chalcopyrite, pyrrhotite and galena (Table 2.3). A difference between the mineral association of
Cu- and Pb-bearing sulphosalts and gudmundite is supported by the results of automated mineralogy analysis (cf.
Table 4.4). However, it cannot be ruled out that the difference in mineral association could be partly due to
differences in the ratio of chalcopyrite and sphalerite caused by the selectivity when sampling the massive sulphides.

Table 2.3

Position and mode of occurrence of common Sb minerals, in ore and in particles found in grounded flotation products (Paper I
and II). Sb mineral: Boul – boulangerite, Bour – bournonite, Gd – gudmundite, Mene – meneghinite, Ttr – tetrahedrite. For
abbreviation of other minerals see Table 2.1. LR-Zn – Zn-dominated massive sulphides in the deeper part of the Rockliden
deposit (ca. 400 m below surface), MD – mafic dykes, MS – massive sulphides, Volc – felsic volcanic rock.

Mineral Position in the ore body Association in the mineralisation Association in the particles
• Tetrahedrite MS (only traces in LR-Zn ), Gn, Bour, Ccp, Sp, Po Gn, Sp, Ccp, Po, silicate
MD minerals
• Bournonite MD, MS Gn, Ccp, Ttr, Sp Gn, (Sp, Ccp, Boul, Py)

• Meneghinite & MD, in LR-Zn Apy, Gn, silicate minerals Gn, Bour, silicate minerals
Boulangerite
• Gudmundite MS, Volc, MD Po, Sp, Gn, Ttr, Apy (as small grains < 10 μm Gn, Ccp, silicate and carbonate
with Ccp, Po, Sp, Gn), silicate and carbonate minerals
minerals

13
2.2.2 Expected distribution of Sb minerals in flotation process

Beside the ore characterisation, studying flotation products with qualitative and semi-quantitative (microscopic)
methods is an important aspect in outlining rock- and ore-intrinsic factors, which potentially control the distribution
of minerals, in particular Sb minerals. According to Lager and Forssberg (1990), ferrous Sb minerals, such as
gudmundite reported largely into the tailings both in a soda and lime environment, Cu- and Pb-bearing Sb
sulphosalts floated in both environments, and for the Pb-bearing Sb sulphosalts a slightly better flotation was
documented in the soda environment. An initial flotation test on three Rockliden composites was conducted in 2010
and the Sb mineralogy of the final flotation products was documented by SEM/EDS-based INCA feature analysis
(Bolin 2010, Fig. 2.3). The results of the semi-quantitative study indicated that the behaviour of the Sb minerals is in
accordance with the findings of Lager and Forssberg (1990). Microscopic studies on the final flotation products
indicated that the distribution of Sb minerals is not only influenced by their own flotation behaviour related to the
mineral chemistry, but also by locking with other minerals in composite particles, which in turn could be related
back to the mineral association documented in the ore samples (Paper I, Table 2.3).

Fig. 2.3. Relative composition of Sb minerals in the final flotation products based on INCA analysis (summing to 100% Sb
minerals in each studied product). The recovery of Sb is given for all products from the initial flotation test (Bolin 2010). Mineral
abbreviations: Ttr – tetrahedrite, Bour – bournonite, M/B –meneghinite-boulangerite, Gd – gudmundite. Location of the samples
is indicated in figure 2.1.

14
2.2.3 Implications of ore characterisation for automated mineralogy

Optical microscopy is of limited use in the distinction of (minor and) trace minerals in multiple samples. This is
known for the complex Sb mineralogy of massive sulphide ores at the Rakkejaur deposit (Skellefte district) due to
the small grain size and similar optical properties (e.g. Lager 1989); and it is also true for quantitative ore
characterisation studies at the Rockliden deposit. Also, limitations are noticed in the distinction of minerals with
close mean atomic number (i.e. low BSE contrast) and similar chemistry using SEM-based automated mineralogy
such as MLA techniques. The setting up of MLA measurements would require a special modification (fine tuning)
of the BSE images in order to distinguish boulangerite and meneghinite (Paper III, and references therein). Given
the similar chemistry, the flotation behaviour of meneghinite and boulangerite is expected to be similar and thus
they were combined for SEM-based automated mineralogy. Generally, it is recommended to reduce the complexity
in the mineralogy for the purpose of quantifying the mineralogy of ore and processing products, i.e. before a particle
tracking technique can be applied on automated mineralogy data (Lamberg and Vianna 2007). The minerals
documented in the massive sulphides and the immediate host rocks at Rockliden are listed according to their mean
atomic number in figure 2.4. Figure 2.4 shows also a scheme which was used to group the minerals in order to
reduce the complexity in mineralogy. All non-sulphide minerals and non-Sb-bearing trace minerals are grouped
together as NSG (non-sulphide gangue). Furthermore, arsenopyrite and pyrite are combined. Arsenopyrite is a minor
mineral at Rockliden and is expected to show similar flotation behaviour as pyrite (Benzaazoua et al. 2002).
However, in a study considering the distribution of e.g. Sn or As during flotation, the mineral grouping would
obviously need to be changed.

Fig. 2.4. List of documented minerals sorted by their mean atomic number illustrating the possible distinction in BSE
(backscattered electron) images (Minz 2013). Outlined with arrows and boldly highlighted text is the mineral grouping used for
particle tracking of automated mineralogy measurements.

15
3 Materials and Methods

3.1 Sampling, assays and microscopy

A list of all samples used in this study is given in Table 3.1. Samples that have been specifically collected for this
study are marked and sampling for the second part of the Rockliden PhD project is described below in more detail.

Table 3.1

Summary of samples and analysis available to the Rockliden PhD project. *) Samples of first sampling campaign (part 1 of the
thesis), **) Samples of the second sampling campaign (part 2 of the thesis), remaining samples provided by Boliden AB.

No. sample type experimental, analytical methods

490 Drill core samples, RC drilling Surface mapping, geochemical analysis and drill core logging
(Exploration Department, Boliden AB)
59(* Drill core samples (mostly ca. 40 cm Geochemical analysis and mineralogical studies (optical microscopy,
quartered BQ drill core) SEM/BSE imaging, SEM/EDS, EPMA/WDS)
3 Composite drill core samples (each ca. 15 to Laboratory flotation tests (2009, Division of Process Technology,
30 kg) Boliden AB), geochemical analysis
9 Final products of laboratory flotation tests Geochemical analysis and mineralogical studies on polished epoxy
(2009) mounts (SEM/EDS, modal analysis with INCA feature (cf. Liipo et al.
2012), EPMA/WDS)
29(** Drill core samples (mostly ca. 1 m quartered Geochemical analysis, semi-quantitative mineralogical studies on
BQ drill core) polished epoxy mounts (SEM/EDS, modal analysis (INCA feature (cf.
Liipo et al. 2012))
60(** Sieved products of simplified laboratory Geochemical analysis on products of all composites, GXMAP-MLA on
flotation tests (2013) on 4 composites (each sieved products of 4 composites (= 60 samples), 10 samples selected for
ca. 2 kg), unsized products on 2 composites sequential leaching test

Based on the results of the first part of this study (see section 2), drill core material was selected for a laboratory
flotation test (cf. Paper III). A critical point when dealing with trace elements (such as Sb) or minerals is to measure
a sufficient number of particles to achieve sound statistics (e.g. Byrne et al. 1995; Chetty et al. 2009; Sutherland and
Gottlieb 1991). Thus, for this study drill core samples were selected such that the Sb grade was higher than ca
0.1 wt%. Twenty nine drill core intervals of ca. 1 m length were sampled, crushed and assayed (Fig. 3.1a and b).
The collected drill core intervals were grouped into six sets depending on their chemistry and ore characteristics,
representing Cu- and Zn-dominated massive sulphide ore types and Sb-rich mafic and altered felsic rock types (Fig.
3.1b). Polished epoxy grain mounts were prepared from all crushed drill core samples and the Sb mineralogy was
studied with a Zeiss Gemini FESEM at Luleå University of Technology, Sweden. From each set, two to three drill
core intervals were selected based on similarity in main and Sb mineralogy, blended and ground to a P 80 of ca.
50 μm in a laboratory rod mill (Fig. 3.1c).

16
Fig. 3.1. (a) Distribution of drill core samples in the Rockliden ore deposit. Viewing to the west on the Rockliden massive
sulphide ore body (translucent shapes). (b) Ternary plot showing the classification of the drill core samples. (c) Laboratory
flotation test using a Magotteaux bottom driven flotation machine. Flotation products: Cu–Pb cleaner concentrate (Cu CC), Cu–
Pb cleaner tailing (Cu CT), Zn cleaner concentrate (Zn CC), Zn cleaner tailing (Zn CT) and Zn rougher tailing (Zn RT).
Composites: UR-Cu – upper Rockliden Cu-dominated massive sulphide, UR-Zn – upper Rockliden Zn-dominated massive
sulphide (Sb/Ag ratio > 30), UR-Zn-Ag – upper Rockliden Zn-dominated Ag-ULFKPDVVLYHVXOSKLGH 6E$JUDWLR” /5-Zn –
lower Rockliden Zn-dominated massive sulphide, MD – amphibole-dominated mafic dyke, Volc – altered felsic volcanic host
rock.

17
The control on the Sb mineralogy made the samples of this study different from a snapshot approach which is based
only on grade and spatial location assuming constant mineralogy and lithology within the deposit (Lorenzen and
Barnard 2011). The sampling strategy did not aim for the collection of samples which represent the entire or parts of
the Rockliden ore, but attempted to collect samples clearly showing differences in the main and Sb mineralogy. This
can be described as sampling for end-member or ore variability testing (cf. Lamberg 2011; Lotter et al. 2003).
Generally, drill core samples will account for higher variability and tend to give more extreme grades compared to
samples which would represent mining blocks (Dunham and Vann 2007).

Directly after grinding, the six composites were subjected to a simplified laboratory flotation test in the technical
laboratory of the Boliden Mining Company (Fig. 3.1c, reagents are listed in Paper III). Five flotation products were
collected from each test: Cu–Pb cleaner concentrate (CuCC), Cu–Pb cleaner tailing (CuCT), Zn cleaner concentrate
(ZnCC), Zn cleaner tailing (ZnCT) and final tailing (ZnRT) (Fig. 3.1c). Their chemical composition was measured
by XRF analysis (X-ray fluorescence spectrometer SPECTRO X-LabPro, technical laboratory, Boliden Mining
Company, Sweden). Four composites and their flotation products were selected, and the products sieved into three
size fractions (-20 μm, -45 to +20 μm, +45 μm) and the chemical composition of each fraction was assayed by XRF
and for 21 major and trace elements by plasma atomic emission spectroscopy (ICP-AES) at ALS Minerals Division,
Piteå, Sweden.

3.1.1 Note on chemical assays as basis for element to mineral conversion

As mentioned above, different analytical methods were used to determine the elemental grades. Assays were
obtained for all sieved products by XRF (X-ray fluorescence spectrometer SPECTRO X-LabPro, Boliden Mining
Company) and ICP-AES analysis (ALS Minerals Division). Since several Sb minerals are documented in single thin
sections of uncrushed drill core material from the Rockliden deposit (Paper I and II), the usage of plain Sb grades
would not allow direct element to mineral conversion, e.g. in HSC Geo (cf. Roine 2009). Diagnostic leaching can
provide additional constrains in a system with element number lower than the mineral number by determining
additional Sb grades which would allow distinguishing Sb mineral group if they show selective leaching behaviour
(cf. Whiten 2008). Antimony (Sb) in gudmundite and the Cu- and Pb-bearing sulphosalts have different oxidation
state (Moëlo et al. 2008). A potential difference in solubility of the Cu- and Pb-bearing Sb sulphosalts and
gudmundite was expected from the differences in the oxidation status (e.g. Kuhar et al. 2011). The bromine-
methanol selective dissolution was tested as a possible diagnostic leaching technique (Labtium, Finland). The
method had been used successfully to distinguish pyrite and pyrrhotite (Penttinen et al. 1977). Bromine-methanol
selective dissolution and total dissolution was conducted on 10 samples selected from the sieved products of the
laboratory flotation test and assays were obtained with Flame AAS (Penttinen et al. 1977).

3.2 Automated Mineralogy, Particle Tracking and Simulation

Polished epoxy grain mounts were prepared at the “Erzlabor” (Technical University Freiberg, Germany) for all
sieved flotation products. A Mineral Liberation Analyser (MLA) was chosen for SEM-based automated
measurements (Fandrich et al. 2007) and the analysis was conducted with a FEI Quanta 650F. The MLA Suite 3.1.4
software was used for data acquisition and processing at the Geometallurgy Laboratory, Technical University of
Freiberg (cf. Paper III). The following analytical conditions were used: acceleration voltage of 25 keV and emission
current of 10 nA. The grain X-ray mapping mode (GXMAP) was chosen for measurement with the MLA (cf.
Fandrich et al. 2007). The MLA measurements were conducted at a pixel size of 0.75 μm/px, a minimum Quartz
EDX-count of 2000 and a total particle count of >45000 (for samples of sieving fraction >20 μm) and >190000 (for
samples of sieving fraction <20 μm) (Paper III).

18
Prior to particle tracking, the complexity in the bulk mineralogy found in the MLA measurements was reduced
according to the description in section 2.2 (Fig. 2.4). A particle tracking technique was used to mass balance the
MLA measurements and to calculate recoveries of particle types defined by binning algorithm used in HSC
Chemistry (Lamberg and Vianna 2007). Particles from MLA measurements were assigned to defined particle
classes (binary, ternary and complex particles) in a step referred to as basic binning. For all composites, the
mineralogical composition of the feed was back-calculated from the products. During advanced binning, mass
balancing was done on the level for narrow liberation classes defined by adjusting N min to 250 in the combined
products, i.e. in the corresponding bin in the feed.

For building a mineralogy-based flotation model (cf. Paper IV), a first-order kinetic process was assumed and the
mineral specific flotation rate constants (k value) were calculated according to the following equation (Eq. 1):

(Eq. 1) R = m i (1 – e(-kit))
with: R – Recovery, k i – kinetic constant or rate of flotation for the component i, m i – mass fraction of
component i, t – discrete time in a batch cell (Runge et al. 1997; Welsby et al. 2010)

Three levels of simulation were done based on the detail of mineralogical information used for each simulation
option: unsized (SL1) and sized bulk mineralogy (SL2) and particle information (SL3, Table 3.2). Depending on the
detailed level of information chosen for simulation, the k values were modelled from the observed recoveries of
minerals (liberated minerals in simulation option SL3). For each mineral (or mineral group, cf. Fig. 2.4), a k value
was chosen either derived from the recoveries measured for the Zn-dominated massive sulphide composite of the
upper part (UR-Zn) or from the Zn-dominated massive sulphide composite of the lower part of the Rockliden
deposit (LR-Zn). These selected k values were then used in HSC 7 SIM simulating the recovery of minerals based
on the feed information provided by particle-tracked MLA measurements of the four composites. The detailed level
of information for the feed was chosen corresponding to the level by which the k values were modelled (Table 3.2).
Beside the three mineralogical levels (SL3 to SL1), an assay-based option (SL0, Table 3.2) was included using
common drill core assays: Cu, Zn, Pb, As, Sb and S. It should be noted that this does not allow for distinction of Sb
bound by different minerals.

Table 3.2

Combinations for simulation. Abbreviations: SL –simulation level, UR-Zn – upper Rockliden Zn-dominated massive sulphide,
LR-Zn – lower Rockliden Zn-dominated massive sulphide.

Depth of information for feed to simulation and Assays-based Bulk Bulk Liberated
k values selected from UR-Zn and LR-Zn mineralogy, mineralogy, mineralogy, particles, sized
unsized unsized sized

simulation option SL0 SL1 SL2 SL3

The simulation was based on the flowsheet of the laboratory flotation test (Fig. 3.1c, see also Paper IV). A residence
time of 5.5 min for rougher flotation cells and of 3.5 min for the cleaner flotation cell was used. As the bubble sizes
were not measured and air flow rates were fixed in the tests, the bubble surface area flux (S b ) was set to the value of
“1” in the simulation. The other flotation cell parameters were kept as suggested by the flotation model in HSC SIM
7.1 modelling software.

19
3.3 Evaluation of Error Sources

A complete error analysis for the sampling, experimental and analytical work of the thesis was not included.
However, remarks on different types of errors and error sources are given in this section.

3.3.1 Notes on sampling errors

Errors related to sampling of particular material are described in Pitard (1993). Some aspects are listed below in
Table 3.3 and some notes are given in the following.

Table 3.3

List of errors regarding the sampling procedure in the Rockliden project. The table is not a complete list of all aspects of errors
occurring during sampling of particle material, and the reader should consider Pitard (1993) for this purpose.

Error Remarks

Overall estimation The OE cannot be empirically analysed (Pitard 1993);


error (OE) OE = TE + AE , with:
TE – WRWDOVDPSOLQJHUURU7( Ȉ n (SE n + PE n ) with n = number of preparation stage,
AE – analytical error (comments on AE are given below).

Preparation error The PE is an example for accidental errors. It can be introduced by contamination, loss, chemical or
(PE) physical alteration, human mistakes, fraud.
Comments on PE are given below.

Sampling (or SE = (a S – a L ) / a L , with:


selection) error a L – unknown true content of a lot L,
(SE) a S – estimator of the content of the lot L.
[also: SE = EE + DE + CE with DE – (increment) delimitation error, EE – (increment) extraction
error, CE – constitution selection error]

Fundamental The FE is related to constitution heterogeneity and is the remaining error when all sampling
sampling error (FE) operation is perfect.
Comments on FE are given below.

Care was taken to minimise the preparation errors (PE) during sampling of the Rockliden ore material. The
multitude of labs involved all analysis and related packing-shipping of samples to different labs is expected to have
a relatively large impact on the PE and hence on the total sampling error. However, the PE cannot be studied by
statistics (Pitard 1993) and remains an unknown error component of this project.

Depending on the purpose of the application, the following range is suggested for the fundamental error: s FE ”
for technical sampling and process control and s FE ”IRUH[SORUDWLRQDQGHQYLURQPHQWDOVDPSOLQJ(Pitard 1993).
It was pointed out that the Gy formula tends to overestimate the calculated variance (Bazin et al. 2013 and reference
therein). A method to describe the standard deviation of the fundamental error (FE) for a two species system (e.g.
valuable /target mineral and gangue) has been outlined for low grade ores by Bazin et al. (2013). The mineralogy of
Rockliden, however, should not be reduced to a binary system in this study. The HSC 7 Geo Sampler, based on Gy's
equation, has been used to estimate the impact of the fundamental error for the sampling of blended drill core
material before crushing, for sampling of MLA analysis from the Cu–Pb cleaner concentrate (CuCC) and the final
tailing (ZnRT) (Table 3.4).

20
Table 3.4

Estimates of fundamental error (%) during sampling before crushing and sampling of the Cu–Pb cleaner concentrate (CuCC) and
Zn rougher tailing (ZnRT) for chemical analysis (ALS), considering the trace mineralogy of Sb in the Zn-dominated massive
sulphide composite from the upper part of the Rockliden deposit (UR-Zn) and the mafic dyke composite (MD). For flowsheet see
figure 3.1c. Mineral abbreviations: Bour –bournonite, Gd – gudmundite, Mene – meneghinite, Ttr – tetrahedrite.

UR-Zn MD
s (%) Ttr Bour Mene Gd Ttr Bour Mene Gd
crushing Æ grinding 0.99 1.94 2.58 1.34 0.98 0.94 0.53 2.08
CuCC Æ ALS 0.30 0.85 1.16 0.88 0.33 0.19 0.09 2.11
ZnRT Æ ALS 10.92 4.81 6.34 1.42 1.70 2.13 1.23 4.22

3.3.2 Notes on analytical errors

The standard deviation of repeated measurements on the same sample is a measure of the precision of an analysis
(Menditto et al. 2007). The bias quantifies the closeness between a reference value and the average of large series of
test results; which is referred to as trueness of an analysis and has often been used indistinctively with accuracy
(Menditto et al. 2007).

For chemical analysis, ALS Minerals, the laboratory used, states the relative accuracy and precision with ±5% for
elements at high level concentration and ±10% for elements on trace level concentration (ALS Minerals Statement
on Accuracy and Precision, pers. comm. Tony Ökvist, 2014-03-07). External standards have not been supplied to
the samples for the products of the laboratory flotation test to confirm the bias assessing the accuracy. However,
earlier studies on massive sulphide drill core samples using similar analytical package and external standards
support the accuracy and precision given by ALS.

A relative error is assigned to assays of XRF analysis (X-ray fluorescence spectrometer SPECTRO X-LabPro)
which is stated to be less than ±5% for major and minor elements.

Figure 3.2a shows the reproducibility of XRF analysis, i.e. the combined error of repeating the sample preparation
and measurement (cf. Parian 2015). The average coefficient of variation CV AV (cf. Abzalov 2008) was estimated
with 11 % for the 18 samples (for repeated sampling and repeated analysis). Graphic b (Fig. 3.2) compares different
types of analysis. The reader should be aware that neither the criterion of repeatability (i.e. repeating only the
measurement) nor for reproducibility (repeated sample preparation and measurement) are fulfilled (cf. Parian 2015)
and the spread in the Sb values might be attributed to both analytical errors and errors in sampling (Figure 3.2b).
However, a pattern can be shown for the material of mafic dykes: for Sb grades < 2.5 wt% ICP-AES analysis gives
lower grades than XRF analysis, particularly in the coarse fraction (> 45 μm). Since highly oxidizing solutions
(HNO 3 , KClO 4 and HBr) have been used to dissolve the samples for ICP-AES analysis, the impact of grain size on
the error in analytical results might be more significant in XRF analysis. Large errors are known to occur in XRF
analysis of sample material with large variation in grain size, or with light elements, or having constituents with long
wave characteristic lines (e.g. Si, Al) such as silicates (Mzyk et al. 2002). Such reasons may partly explain the
deviation of ICP-AES analysis and XRF analysis on flotation products of the mafic dyke composite (Fig 3.2b).
Moreover, it should be noted that the mafic dyke samples, dominated by silicate minerals (cf. Table 4.2), were
analysed with the same program at the X-ray fluorescence spectrometer SPECTRO X-LabPro which was set-up and
used for massive sulphides, dominated by Fe sulphide minerals.

21
Fig. 3.2. (a) Reproducibility of Sb assays on ca. 5g sample, analysed with XRF; (b) parity chart for Sb comparing ICP-AES
analysis (ALS: ME-ICP41a, ca. 1g samples) with XRF analysis (ca. 5g sample). Abbreviations: MASS – massive sulphide
composites, MD – mafic dyke composite (coarse size fraction is marked with grey dots with a black circle, see legend graphic b).

3.3.3 Data reconciliation and error sources in MLA measurements

The relative standard deviation of MLA measurements was estimated from repeated analyses on one epoxy mount
and is below 5% for most minerals and elements.

The ALS chemical assays were used to evaluate the quality of the MLA data. Generally, chemical assays provide
more accurate assays than those derived from MLA measurements. The bulk composition derived from MLA
measurements is not effected by stereological error (Spencer and Sutherland 2000), and thus should be identical to
chemical analysis, under the assumption that the mineral chemistry used to derive assays from MLA measurements
is precisely known. Table 3.5 shows the average coefficient of variance (CV AV ) for assays derived from MLA data
and from ICP-AES analysis and the CV AV values indicate that they are generally in good agreement. The high
coefficient of variance for Cu assays might be related to the strong reporting of Cu to the Cu–Pb cleaner concentrate
(CuCC). Excluding the Zn cleaner tailing (ZnCT) and final tailing (ZnRT) gives a lower coefficient of variance.
Difference between the assays includes both sampling error and analytical errors of MLA and ICP-AES analysis.

Table 3.5

CV AV (coefficient of variance, cf. Abzalov, 2008) for assays (ICP-AES analysis) on sieved products and chemical composition
back-calculated from MLA measurements. CV AV    >  1  Ȉ i=1 N{(a i – b i )2 / (a i + b i )2}](1/2). For abbreviations of
composites see figure 3.1.

CV av Si S Fe Cu Cu* Zn Sb Pb
UR-Cu 76.0 1.3 1.2 24.5 5.6 30.4 19.7 24.1
UR-Zn 24.9 3.1 5.2 64.1 13.6 17.5 16.4 25.4
LR-Zn 31.3 4.2 5.8 62.3 9.7 10.4 21.1 26.0
MD 19.3 11.2 29.0 12.0 2.3 48.4 27.7 14.5
*) includes only assays from Cu–Pb cleaner concentrate.

22
After sieving no size reduction was conducted (neither before sampling for MLA analysis (about 3g per sample) nor
before sample submission for chemical assays (at least 1g for each sample)). Although about 3g of each flotation
product were used to prepare epoxy grain mounts, less than 5 wt% of the material is estimated to be present at the
polished surface recordable by MLA analysis. Thus, the total amount of material analysed with MLA is lower than
the material used for chemical assays and the FE (fundamental error) is supposed to be higher for MLA analysis (cf.
Table 3.4). Generally, it is assumed that the quality of chemical analysis is better than the assays derived from
liberation data (Lamberg and Vianna 2007; Spencer and Sutherland 2000). Since chemical assays (ICP-AES
analysis) are available for all products from the laboratory flotation test, element to mineral conversion could be
utilised to attain a mineralogical mass balance that is fully consistent with chemical assays. However, attempts to
reconcile the mineral liberation data with the assay-based bulk mineralogy in the HSC Chemistry software (Roine
2009) were not applied to the MLA data since it led to adjustment of mineral contents to zero in several products
(Paper III).

The following list summarises challenges with sample preparation for MLA and MLA measurements:

x Stereological corrections were not applied to the experimental data of this study. Spencer and Sutherland
(2000) found that the grades of binaries in feed and flotation products might be differently affected and it
should be expected that in the presented MLA data the recovery of as 3D particles would be different from
2D particles. Since stereological corrections were not applied to the mineralogical complex data for
Rockliden, the reader should be aware that there is an error of unknown magnitude in the experimental data
related to the measurement technique.

x During sample preparation, particle agglomeration is a challenge to automated mineralogy. De-


agglomeration can be partly achieved by careful sample preparation (e.g. Kwitko-Ribeiro 2012) and by
software solutions in measurement programs (Fandrich et al. 2007). Manual adjustment was necessary
during processing of MLA data, especially for the fine fraction (< 20 μm, Paper III).

x In addition to classification of particles due to differences in size, density segregation is a critical aspect in
preparing epoxy mounts (Kwitko-Ribeiro 2012). The sieved samples of the Rockliden case contain
minerals with a wide range of densities. Elements with high atomic number, such as Sb and Pb, tend to
have higher contents in MLA derived assays than in chemical assays compared to S and Fe in the massive
sulphide composites. This effect was tentatively attributed to density segregation (cf. Paper III).

x The minimum grain size for GXMAP-MLA analysis was set to 4 μm (Paper III). Generally, the minimum
mineral size is suggested to be around 5 μm, limited by the interaction volume of the electron beam in the
sample (e.g. Lund 2013). Misidentification of small minerals occurring at the rims of larger mineral grains
has been noticed in MLA analysis (Paper III). Such fine rims may be detected in BSE images but are too
small to give clean spectra in EDS analysis. This may be also a possible explanation for Cu deviation in the
Cu assays derived from MLA data compared to assays from ICP-AES analysis (cf. Table 3.5). Cu is carried
largely by chalcopyrite (more than 70% of Cu in all massive sulphide composites), which could be
misidentified as pyrrhotite in unclear spectra, e.g. collected along margins of pyrite. However, this effect
should be minimal since a Minimum Quartz EDS-count of 2000 was defined in the setup for MLA
measurement, limiting the collection and interpretation of spectra with high noise. Another challenge are
very fine intergrowths frequently found in BSE images in the Zn-dominated massive sulphide composite
from the lower part of the Rockliden deposit. The intergrowths comprise chalcopyrite, pyrrhotite, galena
and gudmundite of a mineral size <1μm, which is too small to be correctly distinguished in MLA data.

x The distinction of minerals is limited in SEM-based automated analysis, in particular if minerals have
similar chemical composition and mean atomic ratio. Meneghinite and boulangerite are an example of
minerals with similar chemical composition and mean atomic ratio. In some cases, where SEM-based
automated mineralogy reaches its limits, wavelength-dispersive X-ray spectroscopy (WDS) on an electron
probe microanalyser (EPMA) may be used as an alternative mineralogical analytical tool with option for
automatisation (Pownceby et al. 2007). However, MLA measurements could also be finely adjusted, i.e. the
back scattered electron image can be tuned and expanded to have a difference between grey levels.

23
Recommendations for measurement are found for the distinction of hematite and magnetite (Bachmann et
al. 2014; Figueroa et al. 2012).

x There are no international standards available for SEM-based automated mineralogy (Sylvester 2012) and
the quality of data is influenced on the set-up and conditions for measurement in laboratories. Keeping data
comparable for samples of an experimental series, the samples of this study were measured over a short
period of time in a single laboratory.

Monte Carlo simulation can be used to evaluate the standard deviation (STD) for the mass proportion and recoveries
of specific particle types (Lamberg and Vianna 2016, and references therein). The STD was based on the relative
standard deviation (RSD) calculated from the particle number (of liberated minerals) according to the following
equation (Eq. 2):

(Eq. 2) RSD = 100 / (NOP)(1/2)


(NOP – number of particles of a specific particle type in a flotation feed and product)

The RSD was then used to estimate the STD for recoveries of liberated minerals by using bootstrap Monte Carlo
simulation (cf. Paper III).

3.3.4 Notes on experimental and modelling errors

To embed the data of this project into a geometallurgical model several additional sources to more general errors
have to be considered. The following sources of error were listed by Bulled and McInnes (2005):

1. experimental errors in data derived from (laboratory) tests (e.g. precision in estimating kinetic factors),

2. errors in geostatistical models related to sample distribution of the ore body,

3. errors in model calibration.

A measured precision of the flotation test and errors on the estimation of the flotation rate factors for each composite
could not be calculated from the experimental data directly since no repetition measurement was conducted. This
was largely due to the project capacity limiting the selection of drill core material and number of composites to be
tested. However, at least one replicate sample would be necessary in order to evaluate the experimental error
(Chauhan et al. 2013; Lotter et al. 2003).

The second error source of the list above was not considered further since building a complete geometallurgical
model is beyond the scope of this study. However, it will become relevant when the data of this study is to be put in
a geometallurigcal context. Since only four composite samples were tested with drill core samples selected from a
limited number of drill cores, the error due to insufficient sample density would be unacceptably high. Currently,
there is no quantitative mineralogical analysis available on drill cores which would cover the ore body more
continuously. The only quantitative available data are chemical assays on intersections of the structurally complex
massive sulphide bodies (Paper II, and references therein). The available mineralogical information is not following
a regular (geometric) grid which would provide sufficient three-dimensional information for modelling (e.g. Farrel
et al. 2011; Keeney and Walters 2011, for more references see section 1.1 and for further discussion section 5.3).

Similar to the second point of the list above, the third error source was not considered further in this study. Results
from a full-scale operation (as outlined in Fig. 3.3) of the Rockliden prospect are not available and only laboratory
tests were part of the analytical work. This gives no basis for calibration of laboratory tests to full-scale operations
or pilot plant results (Ikumapayi 2013).

24
Fig 3.3. Schematic flotation circuit for ore material at Boliden Mineral AB (after Ikumapayi (2013)).

25
4 Results – Quantification and Modelling of the Flotation Distribution of Sb

4.1 Flotation Test – Distribution of Sb Minerals

4.1.1 Characteristics of the composites

Drill core samples representing different rock and massive sulphide types were collected. The drill core samples
were combined to six composite samples to be used as feeds in the laboratory flotation test: four Cu- and Zn-
dominated massive sulphides, further distinguished by their Sb mineralogy, and two Sb-rich host rocks, a felsic and
a mafic rock samples were collected (for description of sampling procedure see section 3.1, Fig 3.1). The flotation
products of four composites were selected for further analysis. The chemical composition of four selected
composites is given in Table 4.1, and the mineralogical composition is presented in Table 4.2.

Table 4.1

Chemical composition of the feed samples. The differentiated Sb grade based on the feed Sb mineralogy back calculated from
MLA measurement. Abbreviations: UR-Cu/-Zn – upper Rockliden Cu- and Zn-dominated massive sulphide, LR-Zn – lower
Rockliden Zn-dominated massive sulphide, MD – amphibole-dominated mafic dyke.

SiO 2 S Fe Cu Pb Zn Sb Ttr Sb Bour Sb Mene Sb Gd


composite (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) Total
UR-Cu 1.7 39.9 43.1 3.4 0.4 2.2 0.07 0.01 0.00 0.08 90.90
UR-Zn 12.4 37.8 31.2 1.5 2.2 8.6 0.21 0.03 0.01 0.21 94.10
LR-Zn 10.0 38.4 32.1 0.3 2.3 10.5 0.00 0.04 0.09 0.16 94.06
MD 30.5 7.3 7.1 2.4 10.6 0.5 0.24 0.57 1.65 0.10 60.89

In the Zn-dominated massive sulphide composites (UR-Zn and LR-Zn), the main sulphide minerals are pyrite (FeS 2 )
and sphalerite (ZnS), followed by pyrrhotite (Fe 1-x S), with minor amounts of chalcopyrite (CuFeS 2 ) and galena
(PbS) (Table 4.2). Chalcopyrite is the main mineral in the Cu-dominated massive sulphide composites (UR-Cu,
Table 4.2). The non-sulphide minerals (plus remaining trace minerals such as stannite and cassiterite) are grouped
together in the non-sulphide gangue (NSG).

In the feed of the Zn-dominated massive sulphide samples, quartz is dominating the gangue (NSG group, Table 4.3).
In the Cu-dominated massive composite, magnetite and carbonate minerals are the main gangue minerals. Other
gangue minerals include silicates, such as feldspar, muscovite, amphibole, chlorite and epidote group minerals
(Table 4.3). The mafic dykes contain minor amounts of sulphide minerals (Table 4.2) and the main gangue minerals
are amphiboles and phyllosilicates (Table 4.3, see also section 2.1, Table 2.1).

26
Table 4.2

Feed mineralogy by size, back-calculated from MLA data. Mineral abbreviations: Bour – bournonite, Ccp – chalcopyrite, Gn –
galena, NSG – non-sulphide gangue (mostly silicates), Po – pyrrhotite, Py – pyrite (with arsenopyrite), Sp – sphalerite, Gd –
gudmundite, Mene – meneghinite (with boulangerite), Ttr – tetrahedrite.

NSG Py Po Ccp Gn Sp Ttr* Bour Mene Gd


Composite Size fraction (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%)
UR-Cu bulk 9.0 50.1 14.3 20.4 0.6 4.9 0.39 0.04 0.02 0.24
UR-Cu 0-20 μm 8.9 47.0 15.3 22.3 1.0 4.5 0.51 0.08 0.04 0.38
UR-Cu 20-45 μm 8.1 45.6 14.3 24.6 0.5 6.3 0.42 0.02 0.01 0.18
UR-Cu >45 μm 10.0 57.5 13.2 14.6 0.3 4.0 0.21 0.01 0.01 0.14
UR-Zn bulk 13.4 55.9 7.9 4.1 2.5 14.7 0.85 0.13 0.05 0.45
UR-Zn 0-20 μm 12.8 47.2 7.7 5.9 4.4 19.8 1.15 0.25 0.12 0.69
UR-Zn 20-45 μm 14.0 49.9 9.8 4.8 2.2 17.6 1.12 0.11 0.02 0.41
UR-Zn >45 μm 13.6 65.8 7.1 2.3 1.2 9.2 0.48 0.04 0.01 0.27
LR-Zn bulk 13.6 53.6 9.0 1.3 2.3 19.0 0.01 0.19 0.60 0.35
LR-Zn 0-20 μm 10.6 46.5 9.6 1.9 3.9 25.7 0.01 0.32 0.97 0.52
LR-Zn 20-45 μm 9.3 56.5 10.3 1.3 1.7 19.9 0.01 0.17 0.51 0.33
LR-Zn >45 μm 17.7 58.3 8.1 0.8 1.3 13.1 0.00 0.08 0.33 0.21
MD bulk 81.7 5.0 0.4 2.2 3.7 0.4 0.45 1.30 4.67 0.09
MD 0-20 μm 78.1 5.0 0.4 2.4 5.1 0.5 0.57 1.79 6.09 0.10
MD 20-45 μm 80.3 6.8 0.6 2.8 3.0 0.6 0.44 1.19 4.22 0.09
MD >45 μm 86.6 4.3 0.4 1.8 2.3 0.3 0.32 0.76 3.13 0.07
*) Ttr composition for LR-Zn and MD is of type 1 tetrahedrite and for UR-Cu and UR-Zn composite of type 2 tetrahedrite (Table
2.2).

Table 4.3

Composition of the group of non-sulphide gangue (wt%) as estimated by MLA measurements. Mineral abbreviations: Amp –
amphibole minerals, Bt – biotite (partly altered), Cal – calcite, Chl – chlorite group, Dol – dolomite group, Epi – epidote group,
Fsp – alkali feldspar minerals, Mgt – magnetite, Plg – plagioclase, Px – pyroxene minerals, Qtz – quartz, Ser – sericite. Others
include oxides (chromite e.g. in MD), Ca sulphates, apatite and remaining silicate minerals such as undefined clay minerals,
titanite and zircon.

Mgt Dol Cal Epi Amp & Px Chl Bt Ser Qtz Fsp Plg Others
UR-Cu 3.4 2.5 0.5 0.0 0.6 0.5 0.5 0.0 0.3 0.0 0.0 0.2
UR-Zn 0.0 0.0 0.5 0.4 0.8 0.7 0.1 0.7 8.7 0.0 0.5 0.5
LR-Zn 0.2 1.5 0.9 0.1 0.9 0.2 0.3 0.3 8.6 0.0 0.2 0.1
MD 1.2 0.8 0.5 3.8 34.0 10.7 8.0 1.9 8.2 0.1 10.5 1.8

Tetrahedrite is the main Sb mineral in the massive sulphide composites from the upper part of the Rockliden deposit
(Table 4.2). Meneghinite is the main Sb mineral in the Zn-dominated massive sulphide composite from the lower
part of the Rockliden deposit and in the mafic dyke composite. Although gudmundite is the second most abundant
Sb mineral, it is the main Sb carrier in the massive sulphide composites (Table 4.1).

Generally, it can be stated that the Sb minerals have a smaller mineral grain size (P 80 ” —P  WKDQ WKH PDLQ
sulphide minerals (P 80 ranging from 40 to 50 μm) in the grounded feed. Pyrite shows the largest mineral grain size
with the highest P 80 , followed by sphalerite in the massive sulphide composites (Fig. 4.1).

27
Fig. 4.1. Comparison of mineral grain size distribution for selected minerals in the grounded feed of the four composites.
Indicated with a grey box is the range of the P 80 . The light grey area marks the part of the curve below a minimum grain size
which was set to be detected by GXMAP-MLA analysis. Abbreviations: Bour – bournonite, Ccp – chalcopyrite, Gd –
gudmundite, Mene – meneghinite, NSG – non-sulphide gangue (silicates with remaining trace minerals), Py – pyrite (with
arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite. For abbreviations of composites see Table 4.1.

Given the comparatively small grain size of Sb minerals, it is expected that they show a lower degree of liberation
than other sulphides. This is shown in figure 4.2: the Sb minerals are poorly liberated compared to the main sulphide
minerals, pyrite and sphalerite or chalcopyrite in the massive sulphide composites, and the NSG in the mafic dyke
composite. This is noteworthy since the locking of Sb minerals with the main minerals is relevant to the distribution
of Sb minerals in the flotation process. For example the locking of Cu- and Pb-bearing Sb sulphosalts with
sphalerite, pyrite or NSG could cause that these mineral are not reporting to the Cu concentrate, or locking of
gudmundite with chalcopyrite or galena might cause reporting of gudmundite to the Cu–Pb concentrate. Figure 4.2
shows only the liberation of minerals from bulk feed. It was noticed for the Sb minerals that the degree of liberation
is lowest for the coarse size fraction and increases to finer size fractions.

28
Fig. 4.2. Cumulative mineral liberation yield for the two main minerals and the dominant Sb minerals of the four composites.
The mineral grades are given to the left of each curve. Abbreviations: Bour – bournonite, Ccp – chalcopyrite, Gd – gudmundite,
Gn – galena, Mene – meneghinite, NSG – non-sulphide gangue (silicates with remaining trace minerals), Py – pyrite (with
arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite. For abbreviation of composite names see Table 4.1.

Figure 4.3 shows the association of Sb minerals in the grounded feed. For the liberated Sb minerals a standard
deviation is given which is based on the number of particles (see section 3.3.3, Eq. 2). Sb minerals show higher
relative standard deviation related to their lower particle number compared to the main sulphide minerals (Paper III).

The association of Sb minerals with other minerals in locked particles, appears to depend on the main mineralogy,
particularly for the Cu- and Pb-bearing sulphosalts, i.e. they show strong association with chalcopyrite and galena in
the massive sulphide composites (UR-Cu, UR-Zn and LR-Zn). Gudmundite shows a more variable association with
the main minerals. This is largely in agreement with observations made in the semi-quantitative process mineralogy
study (Table 2.3, Paper I).

The chalcopyrite content varies between the three composites, decreasing from UR-Cu (20.4 wt% Ccp) to UR-Zn
(4.1 wt% Ccp) to LR-Zn (1.3 wt% Ccp, Table 4.2). A lower proportion of gudmundite-chalcopyrite binaries was
expected in LR-Zn than in UR-Cu. However, the only distinct decrease in proportion of gudmundite-chalcopyrite
binaries is noticed between UR-Zn and LR-Zn (Fig. 4.3).

29
Fig. 4.3. Proportion of Sb minerals in liberated and binary particles, and the distribution in different size fractions of the feed.
Mineral abbreviations: Bour – bournonite, Gd – gudmundite, Mene – meneghinite (with boulangerite), Ttr – tetrahedrite , B Ccp
– binary with chalcopyrite, B Gn – binary with galena, B NSG – binary with non-sulphide gangue (mostly silicates), B Py –
binary with pyrite (with arsenopyrite), B Sp – binary with sphalerite, B* – other binaries (with pyrrhotite and remaining Sb
minerals), MC – ternaries and more complex particles. For abbreviation of composite names see Table 4.1.

An association index was calculated for the Sb minerals with the main minerals in the studied composite samples
(cf. Lund 2013). If the association index is above 1, the association is said to be more common than the bulk
composition would suggest and if it is below 1, the association is said to be less common than expected (Lund
2013). For example a sample which is rich in chalcopyrite and poor in sphalerite, the Sb minerals would be more
likely to associate with chalcopyrite. Thus, the association index refers to original textures and breakage structures.
The association indices for the main Sb minerals are given in Table 4.4. In particular the Cu- and Pb-bearing Sb
sulphosalts show strong association with galena and show low association with pyrite. The results underline the
observation that gudmundite shows a more variable association with the main minerals in the composites, in contrast
to the strong association of Cu- and Pb-bearing sulphosalts with base-metal sulphides. Also, it points out the higher
than expected association of gudmundite with chalcopyrite in Zn-dominated massive sulphide composites (UR-Zn
and LR-Zn). The association of gudmundite with chalcopyrite is lower than expected for the Cu-dominated massive
sulphide composite (UR-Cu, Table 4.4).

30
Table 4.4

Association index for the main Sb minerals with base-metal sulphides and gangue minerals. For calculation of the association
index see Lund (2013). Abbreviations: Bour – bournonite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene –
meneghinite, NSG – non-sulphide gangue (silicates), Py – pyrite (with arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite, MD –
mafic dyke composite, LR-Zn – Zn-dominated massive sulphides from the lower part of the Rockliden ore deposit, UR-Zn/-Cu –
Zn- and Cu-dominated massive sulphides from the upper part of the Rockliden ore deposit.

Composite UR-Cu UR-Zn LR-Zn MD MD UR-Cu UR-Zn LR-Zn


WDUJHWPLQHUDOĺ Ttr Ttr Mene Mene Bour Gd Gd Gd
associated with Ļ
Ccp 1.75 2.27 1.76 1.29 1.69 0.80 4.15 2.02
Gn 26.04 13.37 12.24 4.17 2.47 3.26 1.60 3.27
Sp 1.93 1.39 1.02 0.27 0.70 0.72 0.62 0.64
Py 0.07 0.05 0.13 0.10 0.14 0.29 0.16 0.19
Sil 0.70 0.41 0.30 0.68 0.57 3.02 1.96 1.31

Trace element mineralogy besides Sb minerals

In addition to Sb, other deleterious elements are also detected at Rockliden, and the average content of these
elements is given in Table 4.5. Although these elements are not the focus of this study, a brief mention of their
distribution in minerals is given. Table 4.5 also compares of chemical assays and assays derived from MLA
measurements. It should be noted that the contents of trace elements, such as Hg in sphalerite, are below the
detection limit of SEM/EDS analysis and a fixed mineral chemistry had to be assumed from previous mineralogical
studies (cf. Minz 2013) before assays were extracted from MLA measurements. The MLA-derived assays, and
particularly those concerning Hg and Ag, should be considered with caution.

Table 4.5

Content of critical elements in the composites, derived by mass-balance over analysis on flotation products. ALS – ICP-AES
chemical assays, MLA – assays back-calculated from MLA measurements. In the case of Bi only ICP-AES chemical assays are
available.

ALS MLA ALS MLA ALS MLA ALS ALS MLA ALS
composite As / wt% As / wt% Ag / ppm Ag / ppm Hg / ppm Hg / ppm Cd / ppm Sn / ppm Sn / ppm Bi / ppm
UR-Cu 2.1 2.2 88 76 140 26 49 665 1288 32
UR-Zn 0.5 0.5 127 146 634 19 208 1356 1629 33
LR-Zn 0.5 0.4 53 8 360 14 220 946 939 51
MD 1.8 1.5 441 199 33 30 18 215 333 496

Arsenopyrite is the main carrier of arsenic (As), hosting nearly 100 wt% of the As detected in the four feed samples
as recorded by MLA analysis (Table 4.6). It is a minor mineral in the feed of the massive sulphides. In the mafic
dyke sample the arsenopyrite content is higher than the pyrite content, but the combined arsenopyrite–pyrite content
is below 5 wt% and assigned to the minor minerals. Additionally, other As minerals such as native arsenic were
found in trace amounts in the mafic dykes by SEM/EDS analysis. The recovery of arsenopyrite based on non-
particle-tracked MLA measurements indicates that its distribution pattern in the flotation test is similar to pyrite (see
section 4.1.2 below).

31
Tetrahedrite is a main carrier mineral for silver (Ag). However, the Zn-dominated sample from the lower part of the
Rockliden deposit has a low tetrahedrite content (Table 4.2), and thus Ag cannot be accommodated by tetrahedrite,
even if high Ag values in tetrahedrite are assumed (e.g. as high as for the tetrahedrite from mafic dykes, Table 2.2).
The occurrence of solid solutions (Table 4.6) was observed which could host the Ag not accommodated by
tetrahedrite. These solid solutions appear as small sulphide droplets or inclusions mostly in association with
chalcopyrite, and comprise Ag, Hg and Sb. A fixed composition of the solid solution was used for the calculation
presented in Table 4.6: 55 wt% Ag, 20 wt% Sb, and 25 wt% Hg. Using this composition makes it possible to explain
about 90% of the Ag grade (chemical assay) by mineralogical MLA data in the case of the massive sulphides from
the upper part of the Rockliden deposit (UR-Cu and UR-Zn). However, less than 50% Ag (chemical assay) can be
explained by assays back-calculated from MLA data using the fixed composition of solid solutions and tetrahedrite
in the remaining composites (LR-Zn and MD, Table 4.5). Thus, the current approach using fixed Ag content in
tetrahedrite and solid solutions might be insufficient to derive reliable Ag assays by MLA measurement. Further, Ag
might be hosted in other minerals, e.g. as a trace element in sphalerite or chalcopyrite. However, more detailed
mineralogical studies are necessary to improve the knowledge of the Ag mineralogy, e.g. by using EPMA/WDS or
Lacer ICP-MS analysis.

The Hg assays for the feed samples derived from MLA measurement are lower than those from chemical assays
(Table 4.5), which is explained by the fact that the Hg-free sphalerite was used for back-calculation of the assays
from MLA measurements. However, mercury (Hg) is known to occur in sphalerite from VHMS deposits (e.g.
Benzaazoua et al. 2002; Cook et al. 2009; Grammatikopoulos and Roth 2002; Grammatikopoulos et al. 2006). In
sphalerite of Rockliden massive sulphides, Hg has been recorded to be below the lower limits of detection of
EPMA/WDS analysis, i.e. in concentrations below 0.2 wt% (Minz 2013). Assuming sphalerite and solid solution to
be the only Hg-bearing minerals, and the solid solution to contain on average 25 wt% Hg, an artificial Hg content
can be calculated for sphalerite using the Hg grades of the composites (chemical assays) and the grade of sphalerite
and solid solution in the composites as indicated by MLA measurement (Table 4.6). Mercury (Hg) contents above
0.3 wt% were not confirmed by EPMA/WDS analysis on sphalerite (Minz 2013). In addition to sphalerite, other
sulphide minerals such as tetrahedrite might contain Hg. An Hg content of 0.2 wt% is known for tetrahedrite from
Neves Corvo VHMS deposit (Benzaazoua et al. 2002). Such a low Hg content is below the lower limit of detection
for EPMA/WDS measurements conducted on Sb minerals in this study (Paper I).

No distinct Cd-bearing mineral has been documented so far at Rockliden. It is assumed that Cd is largely bound by
sphalerite. This is supported by the observation that Cd and Zn show almost identical recovery into flotation
products and a strong correlation between Cd and Zn assays in drill core samples was found (RSQ = 0.96, n = 94).
Traces of Cd were detected in galena; EPMA/WDS analysis indicated 0.13 wt% Cd in galena. Assuming that galena
in all composites had a Cd content of 0.13 wt%, and locating the remaining Cd to sphalerite, the Cd grade would
range between 0.2 wt% to 0.5 wt% in sphalerite (Table 4.6). Such high Cd values have not been documented by
EPMA/WDS analysis of Rockliden sphalerite and the mineral chemistry ought to be studied in more detail if the Cd
flotation distribution by mineral would need to be disclosed.

EPMA/WDS analysis indicated a Bi content of 0.5 wt% in galena and this allows to accommodate all Bi in the
galena of (Zn-dominated) massive sulphide composites. However, the content of galena in mafic dykes is not
sufficient to accommodate all Bi at a Bi grade of 0.5 wt% in galena. Other minerals might contain Bi and the
presence of Bi-bearing phases was indicated by SEM/BSE analysis (single grains, <5μm length, observed in
chalcopyrite with a composition of ca. 10 wt% Bi and ca. 90 wt% Sb).

Tin (Sn) is a trace element in the Rockliden composites (Table 4.5) and largely hosted by cassiterite and stannite
(Table 4.6). The Tin (Sn) grade is expected to range generally between 20 and 100 ppm and peak concentrations of
ca 3500 ppm Sn are documented only in clast-rich, isolated massive sulphide intersections (Paper II). The collected
composites have Sn grade above the expected general range and below the peak concentrations (Table 4.5). MLA
measurements indicate that the flotation behaviour of stannite and cassiterite is opposite to each other, i.e. stannite is
expected to report to the Cu–Pb cleaner concentrate whereas cassiterite is expected to report to the final tailings. If
the flotation distribution of Sn ought to be studied, a mineral grouping should be chosen for particle-tracking
different to the one used in this study (cf. Fig. 2.4).

32
Table 4.6

Distribution of trace elements in minerals (Ttr – tetrahedrite, SoSo – solid solution, Hrz – herzenbergite, Cst – cassiterite, Stnn –
stannite, Apy – arsenopyrite) and calculated Hg and Cd content of sphalerite (Sp**, for explanation see text).

Ag % Ag % Sn % Sn % Sn % As % As % Hg wt% Cd wt%
by Ttr by SoSo* by Hrz by Cst by Stnn by Ttr by Apy in Sp** in Sp**
UR-Cu 83.5 16.5 1.4 49.4 49.2 0.04 99.96 0.22 0.27
UR-Zn 89.5 10.5 0.8 40.8 58.4 0.47 99.53 0.32 0.41
LR-Zn 37.3 62.7 0.5 51.4 48.1 0.01 99.99 0.14 0.19
MD 93.3 6.7 1.7 57.5 40.9 0.05 98.41 0.58 -
*) assuming the following average composition of SoSo: 55 wt% Ag, 20 wt% Sb, 25 wt% Hg.

4.1.2 Characteristics of the flotation products

Figure 4.4 gives the flotation distribution of the bulk chalcopyrite and sphalerite, the main base-metal carriers, and
the two most abundant Sb minerals of each composite. The recovery of fully liberated chalcopyrite into the Cu–Pb
cleaner concentrate is high, above ca. 90%, indicating that flotation selectivity was very good for liberated
chalcopyrite from the massive sulphide composites (Table 4.7). Similarly high recovery into the Cu–Pb cleaner
concentrate is measured for (liberated) Cu- and Pb-bearing Sb sulphosalts and galena from the massive sulphides
(Table 4.7, see also Fig. 4.5 and Fig. 4.6). In contrast to this, the distribution of liberated sphalerite into the Zn
concentrate is lower ranging from about 55% to 75%. The recovery of bulk gudmundite into the final tailing (Zn
rougher tailing) is relatively low, ranging from about 30% to 40%, while 14% to 28% of bulk gudmundite are
recovered into the Cu–Pb cleaner concentrate (Fig. 4.4, Table 4.7). About 8% to 13% of the liberated gudmundite
reports to the Cu–Pb cleaner concentrate which is slightly above the recovery of liberated pyrite (6% to 13%) and
above the recovery of liberated silicates (2% to 3%) into the Cu–Pb cleaner concentrate (Table 4.7). The relatively
high recovery of liberated gudmundite into the Cu–Pb cleaner concentrate was attributed to the slight
hydrophobicity of this mineral in Cu flotation (Paper III).

33
Fig. 4.4. Mineralogical characterisation of the feed and distribution of the valuable minerals and the main Sb minerals in the
laboratory test flotation, based on the bulk mineralogy (wt%) derived from MLA data. The thickness of arrows indicates the Sb
distribution (Sankey diagram). Horizontal numbers in the enlarged columns of the Sb mineralogy give the Sb distribution (%) in
minerals in the feed (cf. Table 4.1). Abbreviations: Bour – bournonite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene
– meneghinite (with boulangerite), NSG – non-sulphide gangue (with remaining trace minerals), Po – pyrrhotite, Py – pyrite
(with arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite, MD – mafic dyke composite, LR-Zn – Zn-dominated massive sulphides
from the lower part of the Rockliden ore deposit, UR-Zn/-Cu – Zn- and Cu-dominated massive sulphides from the upper part of
the Rockliden ore deposit; Cu CC – Cu–Pb cleaner concentrate, Cu CT – Cu–Pb cleaner tailing, Zn CC – Zn cleaner concentrate,
Zn CT – Zn cleaner tailing, Zn RT – final tailing.

34
Table 4.7

Comparison of distribution of minerals based on bulk mineralogy (bulk) and grade of liberated particles (lib.), without distinction
of size fractions. Abbreviations: Bour – bournonite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene – meneghinite,
NSG – non-sulphide gangue (with remaining trace minerals), Py – pyrite (with arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite,
MD – mafic dyke composite, LR-Zn – Zn-dominated massive sulphides from the lower part of the Rockliden ore deposit, UR-
Zn/-Cu – Zn- and Cu-dominated massive sulphides from the upper part of the Rockliden ore deposit, CuCC – Cu–Pb cleaner
concentrate, ZnCC – Zn cleaner concentrate, ZnRT – final tailing.

bulk lib.
UR-Cu Ccp Gn Sp Py Po NSG Ttr Gd Ccp Gn Sp Py Po NSG Ttr Gd
CuCC 87.3 78.5 25.0 8.8 9.7 13.3 82.0 14.0 91.4 93.3 14.6 7.2 7.5 8.9 85.9 8.3
ZnCC 0.7 4.7 53.0 16.8 21.0 1.9 0.9 10.1 0.0 0.6 67.2 8.4 19.8 1.0 0.4 10.4
ZnRT 0.9 2.4 0.5 52.4 22.6 76.3 0.4 36.9 0.0 0.4 0.2 60.8 22.0 83.9 0.3 41.6
bulk lib.
UR-Zn Ccp Gn Sp Py Po NSG Ttr Gd Ccp Gn Sp Py Po NSG Ttr Gd
CuCC 80.7 85.3 27.8 10.9 6.2 5.1 96.2 28.5 96.0 93.7 17.5 13.3 3.4 2.2 96.2 13.2
ZnCC 1.4 2.6 46.9 6.1 8.1 2.8 0.4 8.5 0.0 0.4 58.1 3.4 6.1 1.3 0.2 9.2
ZnRT 8.5 1.9 1.0 66.6 48.3 83.6 0.2 32.7 0.0 0.1 0.7 60.1 48.4 90.4 0.0 46.3
bulk lib.
LR-Zn Ccp Gn Sp Py Po NSG Mene Gd Ccp Gn Sp Py Po NSG Mene Gd
CuCC 63.1 86.2 20.3 7.8 8.5 5.2 94.6 28.3 99.7 97.5 14.7 5.9 6.1 2.8 97.8 12.8
ZnCC 7.8 4.1 59.7 4.3 9.3 4.5 1.1 6.8 0.0 0.6 66.9 0.9 3.6 1.4 0.6 4.5
ZnRT 16.4 2.9 3.4 71.5 48.3 81.6 1.4 35.4 0.2 0.2 2.4 81.2 52.3 89.0 0.3 52.1
bulk lib.
MD Ccp Gn Sp Py Po NSG Bour Mene Ccp Gn Sp Py Po NSG Bour Mene
CuCC 67.9 41.4 21.2 6.2 8.4 4.4 80.6 77.3 73.9 50.5 11.6 2.5 3.7 2.2 91.1 87.2
ZnCC 13.4 19.4 52.4 57.2 50.1 7.0 7.1 5.9 8.9 14.7 61.6 62.5 53.7 3.6 2.2 2.2
ZnRT 4.4 7.9 6.2 8.8 12.6 71.3 3.4 3.3 1.1 0.7 6.7 5.5 10.6 78.4 0.5 0.3

Figure 4.5 gives a summary of mode of occurrence of the main Sb minerals in the different flotation products
compared to the feed. The proportion of Cu- and Pb-bearing sulphosalts locked with sphalerite in the Zn cleaner
concentrate is relatively high in the feed. The proportion of gudmundite locked with chalcopyrite or galena in the
Cu–Pb cleaner concentrate is also high. Since the bulk recovery of Cu- and Pb-bearing sulphosalts in the Zn cleaner
concentrate is low (ca. 1%, Fig. 4.4 and Fig. 4.5), the percentage of these binaries (Cu- and Pb-bearing sulphosalts
locked with sphalerite) reporting into the Zn cleaner concentrate is low. However, this is different for the locking of
gudmundite with chalcopyrite and galena since about 14% to 28% of the bulk gudmundite is recovered into the Cu–
Pb cleaner concentrate (Fig. 4.4 and Fig. 4.5).

35
Fig. 4.5. Mode of occurrence of Sb minerals in different products of the composites compared with the feed samples (cf. figure
4.3). Numbers inside the columns give the liberation distribution of the Sb minerals (* lib.) and number below the columns give
distribution as calculated from bulk mineralogy (** bulk, cf. Table 4.7). The standard deviation is given as vertical numbers and
is estimated using bootstrap Monte Carlo simulation (based on the relative standard deviation = 100 / (NOP)^(1/2) with NOP –
number of particles of a specific particle type in a flotation feed and product). Abbreviations: B Ccp – binary with chalcopyrite, B
Gn – binary with galena, B NSG – binary with non-sulphide gangue (silicates), B Py – binary with pyrite (with arsenopyrite), B
Sp – binary with sphalerite, Bour – bournonite, Gd – gudmundite, Mene – meneghinite (with boulangerite), Ttr – tetrahedrite. For
abbreviation of composite and product names see figure 4.4.

Figure 4.6 shows the distribution of Sb minerals in selected particles types. The distribution of liberated Sb minerals
and the distribution of selected binary types, hosting Sb mineral grains, are systematic and similar in different
samples (Fig. 4.6). As indicated above, locking with base-metal minerals is controlling the distribution of
gudmundite. Gudmundite–chalcopyrite and gudmundite–galena binaries account for 5 to 10% of the gudmundite in
the feed and report to 75 to 78% to the Cu–Pb cleaner concentrate of all massive sulphide composites (Fig. 4.6).
Binaries of tetrahedrite and meneghinite with sphalerite report largely to the Cu–Pb cleaner concentrate of the tested
massive sulphide composites.

36
Fig. 4.6. Percentage of selected particle types (binaries and liberated Sb minerals) to the total amount of the respective Sb
minerals shown in the feeds of the four composites, and the distribution of selected particle types in the flotation products. Sb
minerals in binaries with main minerals are shown as coloured numbers in boxes coloured according to the coding in the legend.
Number on white boxes give the liberation distribution. Graphic (B) to (E) show examples of Sb minerals in particles of selected
products, as indicated in graphic (A). Abbreviations: Bour – bournonite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena,
Mene – meneghinite, NSG – non-sulphide gangue (mostly silicates), Po – pyrrhotite, Py – pyrite (with arsenopyrite), Sp –
sphalerite, Ttr – tetrahedrite. For abbreviation of composite and product names see figure 4.4.

4.2 Sb Distribution Model for Flotation

4.2.1 Modelling of flotation rates and testing of different simulation options

Theoretical considerations of the modelling and simulation of the Sb flotation distribution are given in Paper IV.
The recovery of minerals was used to model flotation rates (k values) according to a first order kinetic model (cf.
Runge et al. 1997; Welsby et al. 2010, see section 3.2, Eq. 1). The flotation distribution of liberated minerals was
used to calculate k values for simulation option SL3. Modelled k values were selected from the two Zn-dominated
massive sulphide composites (UR-Zn and LR-Zn), one for each of the 10 minerals (groups as defined in figure 2.4).
The k values are given in Table 4.8. The selected mineral-specific k values were used in HSC Chemistry Sim 7.1
modelling and simulation software to calculate the distribution of minerals into the flotation products using
information of the feed providing the same depth of mineralogical details, i.e. unsized and sized, for simulation

37
option SL1 and SL2, and particle data for simulation option SL3 (cf. Table 3.2). The simulation option SL0 is based
on chemical assays converted into minerals (in HSC Geo, see section 3.1 above), which does not allow the
distinction of Sb minerals.

Chalcopyrite, galena and Cu- and Pb-bearing Sb sulphosalts show the highest k values in Cu–Pb flotation and
sphalerite in Zn flotation. The other minerals show distinctively lower values (Table 4.8). This observation is
emphasised for k values derived from recoveries of liberated particles (SL3) compared to k values derived from bulk
mineralogy (SL2 and SL1, Table 4.8).

Table 4.8

Flotation rates (k values) calculated from mineral recoveries (partly shown in Table 4.7). Model options are listed in Table 3.2.
Flotation rates of option SL3 are based on distribution of liberated minerals (data presented for size fractions). Abbreviations:
Bour – bournonite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene – meneghinite (with boulangerite), NSG – non-
sulphide gangue (mostly silicates), Po – pyrrhotite, Py – pyrite (with arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite; CuF – k
values for Cu–Pb flotation circuit, ZnF – k values for Zn flotation circuit.

NSG* Py Po Ccp Gn Sp Ttr Bour** Mene Gd


CuF 0-20μm SL3 0.01 0.02 0.04 0.86 0.83 0.09 0.85 0.63 1.09 0.06
CuF 20-45μm SL3 0.01 0.02 0.04 1.32 0.90 0.10 1.43 1.17 1.31 0.09
CuF >45μm SL3 0.01 0.02 0.01 1.38 0.81 0.09 1.44 0.87 1.02 0.08
ZnF 0-20μm SL3 0.01 0.02 0.07 0.11 0.30 0.67 0.39 0.08 0.29 0.08
ZnF 20-45μm SL3 0.01 0.02 0.13 0.07 0.33 1.30 0.51 0.05 0.20 0.08
ZnF >45μm SL3 0.01 0.01 0.06 0.11 0.28 1.43 0.22 0.08 0.18 0.10
CuF SL1 0.02 0.03 0.04 0.51 0.60 0.14 0.97 0.62 0.85 0.14
ZnF SL1 0.01 0.04 0.10 0.07 0.24 0.76 0.26 0.11 0.18 0.10
CuF SL0 0.03 0.03 - 0.56 0.72 0.14 - 0.34 - -
ZnF SL0 0.02 0.05 - 0.08 0.78 0.76 - 0.14 - -
*) for simulation option SL0 quartz has been used in element to mineral conversion to represent the NSG group.
**) for simulation option SL0 the total Sb grade was allocated to bournonite in element to mineral conversion.

Observed and simulated grade-recovery curves are given in figure 4.7. Simulation options based on bulk mineralogy
(SL2 and SL1) tend to overestimate Cu and Sb recoveries, whereas the simulation based on particles (SL3) tends to
underestimate recoveries. Curves based on simulation option SL0 show most variable shapes (Fig. 4.7).

The predicted and experimental recoveries were compared using average coefficient of variation (CV av , cf. Abzalov
2008) of which the results are tabulated in Paper IV. No model option gave CV av values consistently below an error
of 5% when experimental and simulated Cu, Pb, Zn and Sb grades and recoveries were compared (Paper IV). Thus,
simulation options (SL0 to SL3) are not recommended to be used beyond the exploration stage or scoping study, for
which the error is suggested to be below 5% (Paper IV and references therein). However, model option SL2 and
SL1 might be sufficient for estimation of element recoveries in an exploration project.

Since the Sb mineralogy is known to be variable in the Rockliden deposit and Sb minerals show different flotation
behaviour (Paper III), it cannot be expected that the k values based on single Sb grades is usable in simulating ore
samples with variable Sb mineralogy. Using a single k value for Sb, gave poor simulation results regarding Sb
grades and recoveries of one composite (UR-Cu, Paper IV). Thus, simulation option SL0 is not recommended to be
used for forecasting the flotation distribution of Sb.

38
Fig. 4.7. Recovery-grade curves for Cu and Sb comparing results from the simulation with the experimental data in Cu–Pb
flotation. Model options: EXP – experimental data; SL3, SL2, SL1 and SL0 simulation options are listed in Table 3.2.

Generally, it is assumed that with increasing detail in mineralogical information more precise modelling and
simulation could be achieved (e.g. Lishchuk et al 2015a). However, currently the simulation option SL3, using
particle information, does not significantly improve the simulation results in comparison to mineralogical options
SL2 or SL1. The relatively large deviation was partly related to the fact that not all binary particles show a linear
recovery pattern in relation to the composition of the particles (Paper IV). However, such linear correlation is used
by default in the flotation model of HSC Chemistry Sim 7.1. Examples of binaries, which show non-linear
behaviour, are given in figure 4.8. It can be seen from the bended shape to the recovery curves (Fig. 4.8), that even a
small amount of Cu- or Pb-bearing minerals (5 to 10 wt%) in binary particles will give a high recovery of such
binary particles into the Cu–Pb cleaner concentrate.

39
Fig. 4.8. Examples of bended shape of recovery curves for selected binaries in selected products. X-axes: weight (wt%) of one
mineral in binary particles (the weight % of other mineral is the difference to 100%), Y-axes: recovery of the binary particle as
defined by experimental data. Abbreviations: Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Py – pyrite (with arsenopyrite),
Ttr – tetrahedrite, Sp – sphalerite, LR-Zn – Zn-dominated massive sulphides from the lower part of the Rockliden ore deposit,
UR-Zn/-Cu – Zn- and Cu-dominated massive sulphides from the upper part of the Rockliden ore deposit; CuCC – Cu–Pb cleaner
concentrate, ZnCC – Zn cleaner concentrate, ZnRT – final tailing.

4.2.2 Distinction of Sb minerals in flotation feed

A comparison between different simulation options (SL0 to SL3) showed that the minimum requirement for
simulation of Sb grades and recoveries is the quantitative distinction of Sb minerals in the feed to flotation. The Sb
minerals show different flotation behaviour and could be distinguished by MLA measurements.

An alternative for MLA measurements, selective leaching was tried, which might allow fast collection of
information on the Sb mineralogy in the exploration stage of a project. The number of (Sb-bearing) minerals
calculated from mineral-specific Sb grades could be increased if Sb minerals show highly selective dissolution
((Lamberg et al. 1997; Whiten 2008), see also section 3.1.1). The solubility of Sb minerals was tested using the
bromine-methanol method (Penttinen et al. 1977). Figure 4.9 shows the Sb grade of the tested samples as bound by

40
gudmundite (Gd) against the difference in the Sb grade treated with total dissolution and bromine-methanol
selective leaching. No indication on selective behaviour of Sb minerals in bromine-methanol leaching could be
shown. The difference between the AAS analysis of the total dissolution and of the dissolution from the bromine-
methanol method is within the margin of sampling and analytical error, as indicated by the negative value -0.07 wt%
(Fig 4.9).

Fig. 4.9. Antimony (Sb) grade bound by gudmundite (Gd, x-axis) vs. difference in Sb grade determined in solution derived from
total leaching and sequential leaching (y-axis).

Alternative leaching methods which are documented in the literature on Sb minerals are for example ferric chloride-
sodium chloride-hydrochloric acid solutions (Correia et al. 2001) and alkaline sodium sulphide leaching (Anderson
et al. 1991; Awe 2013; Filippou et al. 2007). However, these methods are aimed on full dissolution of the contained
Sb minerals. Alkaline sulphide leaching is expected to effectively leach all Sb minerals in Cu–Pb concentrates from
Boliden operations $FKLPRYLþRYi DQG %DODå  $ZH HW DO  )LFHULRYi et al. 2005). Pérez-López et al.
(2008) used the modified BCR-sequential extraction procedure (European Community Bureau of Reference). The
BCR procedure did not result in sequential leaching of Sb from mining products at São Domingos Mine, Iberian
Pyrite Belt (Pérez-López et al. 2008). Bacterial leaching might be an unconventional option for sequential leaching.
Bacterial media described in the literature has been used to extract Cu from tetrahedrite with Acidithiobacillus
ferrooxidans (Filippou et al. 2007 and references therein) and from disordered tetrahedrite in grinding material with
Thiobacillus ferrooxidans (Baláž et al. 1994). Research concerning the selective leaching of Sb using such
unconventional options is not known to the author.

41
5 Summary and Discussion

5.1 Summary of Results

The complex Sb mineralogy of the Rockliden Zn–Cu VHMS deposit was studied and the mineralogy of the ore was
characterised in detail. The main Sb minerals are Cu- and Pb-bearing sulphosalts (tetrahedrite, bournonite and
meneghinite) and gudmundite. The Sb mineralogy differs between massive sulphides and the Sb-rich host rocks. A
general classification was proposed for the massive sulphides, which distinguishes them by the relative abundance
of the main minerals, i.e. pyrite, sphalerite, chalcopyrite and pyrrhotite (Table 2.1). It was documented that the Sb
mineralogy varies between the massive sulphides from the upper and lower part of the Rockliden deposit. However,
no systematic difference could be found in the Sb mineralogy in relation to the classification of the massive
sulphides. Antimony-rich mafic dykes are dominated by Cu- and Pb-bearing Sb sulphosalts and Sb-rich altered
felsic volcanic rocks by gudmundite and Pb–Sb sulphosalts.

Four composite samples were collected from drill core material and combined in order to have a distinctive Sb
mineralogy in relation to the main mineralogy (the two main Sb minerals are given in brackets): one mafic dyke
composite (MD, with meneghinite and bournonite), one Cu-dominated massive sulphide (UR-Cu, tetrahedrite and
gudmundite), one Zn-dominated massive sulphide from the upper part of the deposit (UR-Zn, tetrahedrite and
gudmundite), and one Zn-dominated massive sulphide composite from the lower part of the deposit (LR-Zn,
meneghinite and gudmundite). A laboratory flotation test was carried out on the four composite samples.

The general behaviour of liberated Sb minerals was found to be in close agreement with studies by Lager and
Forssberg (1990) on the Rakkejaur deposit. Similar to liberated chalcopyrite and galena, the liberated Cu- and Pb-
bearing Sb sulphosalts showed a high recovery into the Cu–Pb cleaner concentrate from all massive sulphide
composites. The distribution of bulk Cu- and Pb-bearing Sb sulphosalts into the Cu–Pb cleaner concentrate was
only a few per cent lower than for liberated Cu- and Pb-bearing Sb sulphosalts, although their liberation degree is
relatively low compared to that of the main minerals. This was partly related to their strong association with galena
(and chalcopyrite) (Table 4.4); and both (liberated) galena and chalcopyrite are strongly reporting to the Cu–Pb
cleaner concentrate; but also other particle types (e.g. Cu- and Pb-bearing Sb sulphosalts locked with gangue
minerals or sphalerite) show relatively high recovery into the Cu–Pb cleaner concentrate. Given the high recovery of
all bulk Cu- and Pb-bearing Sb sulphosalts into the Cu–Pb cleaner concentrate, their contribution to the Sb grade of
this product could be estimated by knowing only the grade of these minerals.

Similar to the Sb sulphosalts, gudmundite showed a relatively poor liberation compared to the main minerals in all
ground composites. The distribution of liberated gudmundite particles was found to be most similar to the
distribution of liberated pyrite, i.e. they reported mostly into the final flotation tail. However, the recovery of bulk
gudmundite into the Cu–Pb cleaner concentrate was higher than for bulk pyrite. This is partly due to the occurrence
of gudmundite locked with chalcopyrite and galena, which report largely to the Cu–Pb cleaner concentrate.

Lund (2013) has developed an association index evaluating the association of two minerals in a ground (feed)
sample. Calculating the association index for chalcopyrite and gudmundite gives indices above 1 for the Zn-
dominated massive sulphide composites (UR-Zn and LR-Zn) and an index below 1 for the Cu-dominated massive
sulphide composite (UR-Cu, Table 4.4). Indices below 1 mean that the mineral association is lower than the bulk
composition would suggest and vice versa (Lund 2013). Calculated association indexes showed that a direct
correlation between the chalcopyrite content of the massive sulphide composites and the amount of gudmundite-
chalcopyrite binaries cannot be drawn. Textural features such as grain size, preferred mineral orientation (e.g.
alignment due to deformation), and association, are factors possibly controlling the amount of gudmundite-
chalcopyrite particles derived from the massive sulphide composites. For example, very fine complex particles of


comprising all types of particles which contain the targeted minerals (here Cu- and Pb-bearing Sb sulphosalts)

42
gudmundite intergrown with chalcopyrite, pyrrhotite and galena was documented only in the Zn-dominated massive
sulphide composite from the lower part of the Rockliden deposit (LR-Zn). The occurrence of carbonate cement and
veinlets was locally observed in the massive sulphide samples during drill core logging. In BSE images of polished
thin sections the carbonate minerals were documented to be associated with gudmundite (Paper I). Similar textural
features might be an explanation for the relatively low association index of gudmundite and chalcopyrite but
relatively high proportion of gudmundite-NSG binaries in the Cu-dominated massive sulphide composite of the
upper part of the Rockliden deposit (UR-Cu) compared to the Zn-dominated massive sulphide composites.
However, more detailed and systematic investigations would need to be conducted including quantitative
mineralogical studies on the mineral association in the incoming ore before crushing.

A flotation model was built in HSC Chemistry Sim 7.1 modelling and simulation software (Paper IV, section 4.2),
using the experimental data derived from the laboratory flotation test (Paper III, section 4.1). Three levels of
simulation were defined based on the detail of mineralogical information: unsized bulk mineralogy (SL1), bulk
mineralogy by size (SL2) and particle information (SL3, liberated minerals used for modelling of flotation rates (k
values) and full particle information for the feed to the simulation). Additionally, an assay-based simulation option
was run (SL0). The flotation rates were derived from first-order kinetic process using the mineral recovery as
defined by particle-tracked MLA results on the flotation products. For each mineral, as defined by the particle
tracking technique, a k value was chosen from the Zn-dominated massive sulphide composite (either UR-Zn or LR-
Zn).

Using the mineral-specific k values from two composite samples gave comparatively similar simulation results for
all four tested composite samples in terms of Cu, Pb and Sb grades and recoveries. This fact indicates that the
mineral-specific k values can be used globally, i.e. they are independent of the feed material. However, the four
composite samples which were tested in this study had Sb grades above the average of the Rockliden resource. Also,
the main sulphide mineralogy of the composites was chosen to represent end-member samples. Despite this the
usage of flotation reagents was kept similar in all samples, resulting in slightly lower recoveries of liberated
chalcopyrite in the UR-Cu composite compared to the Zn-dominated massive sulphide composites (UR-Zn and LR-
Zn). It is highly recommended to study additional samples to verify the global use of k values, which was suggested
in this study (cf. Paper IV). Otherwise the final geometallurgical model might be erroneous due to the lack of
representative sampling which is also crucial to the particle-based modelling (Lamberg 2011). Previous flotation
tests were done on three composite samples derived from entire massive sulphide lenses with similar grades as of
the Rockliden inferred mineral resource (Bolin 2010, Fig. 2.1 and Fig. 2.3), and semi-quantitative SEM-based
mineralogical analysis indicates that the distribution of Sb minerals from large-scale samples is similar and hence
their flotation rates are expected to be similar too. However, semi-quantitative SEM-based mineralogical data does
not provide sufficient quantitative input for simulation, in order to verify the global applicability of mineral-specific
k values as suggested in Paper IV.

Relative errors between the simulated and measured grades and recoveries were mostly in the range of 5% to 20%
and in this case the simulation would be usable within the stage of an exploration project according to the error
margins suggested by Pitard (1993). Large errors in the simulated and measured Sb grade and recovery were
documented for the UR-Cu composite using the feed simulation option SL0. This option was based on assays only
and did not allow distinguishing the Sb mineralogy. Thus, it was not recommended to use only assays for simulation
of Sb grade and recovery in the Rockliden case (cf. Paper IV).

Generally, it is assumed that with an increasing detail in mineralogical information the precision in the simulation
and geometallurgical models would increase (e.g. Lishchuk et al 2015a). However, in most cases the simulation
option SL3 did not give much improvement in the simulation of grades and recoveries compared to options SL2 and
SL1. One reason for the relatively poor prediction is that behavoiur of binary and more complex particles was not
correctly implemented in SL3. It is expected that implementing the non-linear behaviour of binary particles will
result in an improvement of the simulation of grades and recoveries based on full particle information. The
implementation of full particle information could eventually give results acceptable to build a simulation usable for
process control.

43
The up-scaling of the laboratory flotation test results might be achieved by determining unit scale up constants for
mineral-specific flotation rates (cf. Runge et al. 2003). However, data from larger scale flotation experiments like
pilot plant test is currently not available which would provide calibration data for scaling-up on a mineralogical
level (e.g. Dobby and Savassi 2005).

The flotation tests and simulations (Paper III and IV) were done with Sb-rich samples. To estimate the Sb grade of
the Cu–Pb cleaner concentrated of an average feed, a simplified flotation simulation was built using a feed
composition derived from the average grade of the inferred mineral resource of the Rockliden deposit (5.2%
chalcopyrite, 6.0% sphalerite, 0.5% galena, 50% pyrite and 37.5% NSG; for metal grades see: New Boliden 2015)
and using split values based on the recovery of minerals (cf. Table 4.7). From all drill core assays, an average Sb
grade of 0.1 wt% was calculated. In the simplified simulation the Sb grade was fixed to 0.1 wt%, with the option of
alternating Sb mineral ratios. Applying the proportion of Sb minerals as given for the composite sample in Fig 4.4,
the Sb grade of the Cu–Pb concentrate was in the range of 0.8 to 1.6 wt% (i.e. recoveries of Sb above 50%).
Assuming hypothetically all Sb to be hosted by gudmundite, which is not documented for the studied drill core
samples, the simplified simulate gave a Sb grade of 0.34 wt% (i.e. recovery of Sb of 20%) in the Cu–Pb concentrate
at a feed grade of 0.1 wt% Sb. Such a Sb grade is above the typical penalty threshold for Cu concentrates at smelters
(e.g. Larouche 2001).

The impact of certain particle types on the distribution of Sb reporting from the feed to the Cu–Pb cleaner
concentrate can be evaluated from the particle-tracking data (Fig. 5.1). The proportion of Sb bound in gudmundite-
chalcopyrite binaries reporting to the Cu–Pb cleaner concentrate ranges from 2% to 4% of the total Sb grade of the
feed, i.e. in all three massive sulphide composites (Fig. 5.1; for example 3.64 % (= 0.52 * 7%) of Sb is bound by
gudmundite-chalcopyrite particles in the UR-Cu composite reporting to the Cu–Pb concentrate and 7.28 % (= 0.52 *
14%) of Sb is bound by bulk gudmundite in the UR-Cu composite reporting to the Cu–Pb concentrate). Assuming
that approximately the same proportion of Sb bound in gudmundite-chalcopyrite and gudmundite-galena particles
would report from an average ore sample to the Cu–Pb cleaner concentrate as in the tested composite and that the
mineral ratios and textures of the average ore would be similar to the massive sulphide composites, this would give
about 20 to 40 ppm of Sb bound in such form reporting to the Cu–Pb concentrate from an average ore with an Sb
grade of 1000 ppm. Further, assuming a penalty threshold of 2000 ppm (Larouche 2001), then the proportion of Sb
wrongly reporting to the Cu–Pb concentrate due to locking of gudmundite with chalcopyrite would be 1% to 2% of
the Sb grade which could cause penalties at the smelter. The low percentage of Sb bound by gudmundite-
chalcopyrite particles in relation to a penalty threshold of 0.2 wt% Sb is currently within the error margins of
simulation options and such effects could not be taken into consideration (Paper IV).

44
Fig. 5.1. Amount of Sb reporting to the Cu–Pb cleaner concentrate (CuCC) via selected particle types. Percentage below the
columns are in relation to Sb head grade and numbers inside the columns are in relation to percentages below the columns. The
proportion of the Sb minerals of each composite reporting to the Cu–Pb cleaner concentrate (CuCC) is shown in colour in the
columns labelled with the composite names (UR-Cu, UR-Zn, LR-Zn, MD). The parallel columns (lined in one row) show the
mode of occurrence (MoO) for the two main Sb minerals in the CuCC of the corresponding composite. Mineral abbreviations:
Bour – bournonite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene – meneghinite, NSG – non-sulphide gangue
(silicates), Py – pyrite (with arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite. Abbreviations of composites: MD – mafic dyke
composite, LR-Zn – Zn-dominated massive sulphides from the lower part of the Rockliden ore deposit, UR-Zn/-Cu – Zn- and
Cu-dominated massive sulphides from the upper part of the Rockliden ore deposit.

The possibilities in decreasing the Sb content in the Cu–Pb concentrate by flotation would require selective flotation
of Cu- and Pb-bearing Sb sulphosalts from chalcopyrite during Cu–Pb separation (cf. Fig. 1.1 and Fig. 3.3). The
common Cu- and Pb-bearing Sb sulphosalts contain both Cu and Pb in their formula (cf. Table 2.2) or are strongly
associated with galena (and chalcopyrite) (cf. Table 4.4). A separation of Cu- and Pb-bearing Sb sulphosalts from
chalcopyrite and galena is not expected to be easily achieved in the case of Rockliden ore, even if regrinding would
be applied to the ore. An increase in the liberation of gudmundite is expected to slightly decrease the amount of Sb
reporting to the Cu–Pb cleaner concentrate. Regrinding, however, could easily cause overgrinding of the main base
metal sulphide minerals and thus has a negative effect on their recovery (Trahar 1981; Trahar and Warren 1976).
Besides this, it would cause extra grinding costs. Further, regrinding might cause sliming effects, particularly for the
relatively soft base metal sulphides and Cu- and Pb-bearing Sb sulphosalts (e.g. Lager 1989). Moreover, the
simplified flotation simulation (see above), has indicated that even if all Sb would be hosted in (liberated)
gudmundite in a feed with an average Sb grade of 0.1 wt%, the Sb grade of the Cu–Pb concentrate could be above a
penalty threshold of 0.2 wt% Sb for Cu concentrates at smelters (cf. Larouche 2001). Thus, successful options in
adjusting particle size distribution are limited.

An effective method to reduce the Sb grade of the Cu–Pb concentrate would be blending the ore so that the head
grade would be low enough for the production of saleable concentrate. Further, there is a possibility to remove Sb
from the Cu–Pb concentrate by leaching. Alkaline sulphide leaching has been shown to be highly effective in
removing Sb from copper concentrates (e.g. Awe 2013). However, as an unconventional process solution, alkaline

45
sulphide leaching is an additional process step and should only be applied to Cu–Pb cleaner concentrate when the Sb
grade is above the penalty threshold at the smelter. Thus, the precise simulation of the grade and recovery of Sb in
the flotation products based on the incoming ore is seen as a necessity for designing the flowsheet for processing of
the Rockliden ore.

5.2 Components for a Particle-Based Geometallurgical Model

The results of this study are to be put into the larger framework of a geometallurgical model. The study was
intended to provide basic information for building a model which is capable to simulate results of mineral
processing related back to ore characteristics, considering the geological and mineralogical variability of the ore.
This study covers in parts a full geometallurgical program as outlined by Lamberg (2011) and listed in Table 5.1. It
should also be noted that the Rockliden deposit is in an exploration stage and that data from laboratory tests is
relatively scarce. In particular quantitative mineralogical information is only available at some points irregularly
scattered over the deposit.

Table 5.1

Steps of a geometallurgical program (Lamberg 2011and references therein) and results available from the Rockliden project.

Step of geometallurgical program Availability

• Collection of geological data (logging, petro-physical measurements, Yes, for intersection of massive sulphide lenses
chemical analyses etc.). (cf. Paper I and II)

• Ore sampling program for metallurgical testing. Yes*, sample selection for a simplified laboratory
flotation test (Paper III).

• Laboratory testing to extract process model parameters. Yes*, simplified laboratory flotation with
GXMAP-MLA analysis and particle analysis
(Paper III and IV).

• Checking the metallurgical validity of the geological ore type Not yet completed, part of the ore types were not
definition. analysed in detail (cf. Table 2.1).

• Mathematical model for the estimation of metallurgical parameters Currently not possible due to the limited number
across the geological database. of tested composites, for discussion see below.

• Metallurgical model consisting of unit operations using the defined Suggestions are given for flotation as unit
metallurgical parameters. operation (applicability of particle data was tested
in simulating of laboratory flotation based on
first-order kinetic process, Paper IV).

• Plant simulation based on the metallurgical process model and No, currently no complete metallurgical process
“calibration of the models via bench-marking for existing operations” model or particle-based geometallurgical model
(Lamberg 2011). available.

*) with the limit to the number of four composites being tested in this study.

The number of samples tested in the laboratory flotation cell in this study is currently not sufficient to extract or
estimate relevant metallurgical parameters from a geological database by means of statistics (cf. Bulled and McInnes
2005). Further, it might be noted that an adjustment of ore type definition towards the definition of potential

46
geometallurigcal units is not possible (e.g. Lotter et al. 2011), which is in part related to the currently incomplete
testing of all ore types outlined in section 2.1.

Despite the limited number of composites studied in the laboratory flotation test, a theoretical concept for the
particle-based geometallurgical model will be outlined in this section for the Rockliden deposit, following the
guidelines given by Lamberg (2011) and Lund (2013). Such model will be composed of three sub-models (Fig. 5.2).
A process-adapted geological model and a particle-breakage model for the Rockliden deposit have not been studied
in depth so far. However, in the following, a short account will be given on them. Mineralogical parameters, which
are necessary when building a particle-based geological model, will be discussed. Finally, the section 5.3 gives
comments on a statistical modelling alternative in comparison to the particle-based approach.

Fig. 5.2. Components of a particle-based model on the forecast of product quality (e.g. Sb grade). The scheme is constructed
after Lamberg (2007) and Lund (2013).

5.2.1 Process-related geological sub-model

Lund (2013) highlighted the importance of quantitative mineralogical information arguing that samples with same or
similar chemical composition do not need to have the same mineralogy. Indeed, the mineralogical information is
critical to mineral processing and the overall performance of a plant (for examples see section 1.3). Another aspect
that should be extracted from the geological model is textural information. This includes information such as grain
size and mineral association. The grain size is important as to estimate the degree of liberation which is aimed to
range 90 – 95% in the example of iron oxide ores (Lund 2013). Similarly, higher liberation of main sulphides is not
the target in Rockliden, since this can cause overgrinding and an overall lower efficiency in the recovery.

Considerations for the forecast of Sb mineralogy in a block model

Generally, the mineralogical data needs to be collected systematically throughout the full 3D extension of a deposit.
SEM-based automated semi-quantitative and quantitative studies are only available for four composite samples in
this study (Fig. 3.1c). The quantitative information available most consistently from the Rockliden ore bodies, is
chemical assays (Cu, Zn, S, Pb, As, Sb), which in turn can be converted to mineral grades if a simplified mineralogy

47
is assumed. However, there are more minerals recorded than chemical assays collected from drill cores, in particular
concerning the complex Sb mineralogy. It is known that selective dissolution can provide additional information to
the chemical assays derived on a standard basis in the mining industry (Lamberg et al. 1997). The bromine-methanol
leaching method has been tested (Penttinen et al. 1977), but the results did not show selective leaching of Sb
minerals in samples selected from the flotation products of the four composites (Fig. 4.9).

Since no further testing has currently been conducted on selective leaching methods regarding the Sb mineralogy, a
preliminary scheme for element to mineral conversion in HSC Geo (cf. Roine 2009) was suggested (Fig. 5.3). It is
based on the mineralogical information summarised in section 2 and on the MLA measurements of four tested
composites (Table 2.1, Table 2.3, and Table 4.2). The recommended element to mineral conversion in figure 5.3 is
using the assay data which comprises S, Cu, Zn, As, Pb and Sb for most drill core intervals at Rockliden. Since very
little quantitative mineralogical data is available, in particular for the Sb mineralogy when Sb assays are below 0.05
wt%, it is recommended to study the Sb mineralogy quantitatively for a larger number of samples. Furthermore, the
non-sulphide mineralogy is also simplified in figure 5.3, i.e. the occurrence of minerals potentially harmful to the
flotation process, such as chlorite, talc or clay minerals, is not taken into consideration.

Fig. 5.3. Scheme for element to mineral conversion proposed for the Rockliden deposit. Abbreviations: Amp – amphibole, Bour
– bournonite, Ccp – chalcopyrite, Fsp – alkali-feldspar, Gd – gudmundite, Gn – galena, Mene – meneghinite, NSG – non-
sulphide gangue (silicates with remaining trace minerals), Plg – plagioclase, Po – pyrrhotite, Py – pyrite (with minor amounts of
arsenopyrite), Qtz – quartz, Sp – sphalerite, Ttr – tetrahedrite, TS – total sulphides (= 100% - Qtz). For abbreviation of
composites see figure 5.1.

48
Considerations for the forecast of textural information

Textural features, relevant to processing of the ore material, comprise: mineral assemblage and association, grain
size distribution, orientation, shape and grain boundary relationships (Butcher 2010; Walters and Kojovic 2006).
The quantification of textural information is challenging and textural parameters are not necessarily additive
(Lamberg 2011and references therein). However, it is known that liberation characteristics are closely connected to
mineral textures (Lund 2013). Thus, the measurement of textural information ought to be considered in process-
adapted geological models as a component of particle-based geometallurgical models (Fig. 5.2).

Most methods to record textural information (in unbroken ores) give qualitative information based on observations
in thin sections (e.g. McClung and Viljoen 2012; Petruk and Hughson 1977; Petruk 2000), and in the form of
descriptions using classification or coding systems (e.g. Bojcevski et al. 1998; Ghorbani et al. 2013; McArthur
1996). A classification of the Rockliden massive sulphides based on the main mineralogy (Paper II) was proposed,
which is similar to the meso-type classification developed for the George Fisher Ag-Pb-Zn deposit (cf. Bojcevski et
al. 1998).

The distinction of the main minerals in mineral maps is possible by optical microscopy (e.g. Hunt et al. 2011).
However, to quantify textural parameters of a mineralogically complex system such of the Rockliden deposit would
require acquisition of mineral maps using SEM-based automated mineralogy tools. Further, testing of the etching
techniques might be necessary to enhance textural features such as grain size in the uncrushed ore. These procedures
are time consuming and have not been included in this study. It should also be considered that the textural
parameters vary on a centimeter to decimeter scale at the Rockliden deposit and that several standard-size thin
sections would need to be analysed to gain representative textural information. However, it might be sufficient to
use sections of crushed (or ground) drill core material for textural analysis.

When crushed or grinded ore material is studied by SEM-based automated mineralogy tools in order to evaluate the
impact of textural parameters on ore processing, there are two options available to derive textural information from
particles: (1) Particles are described by bulk composition and classified into liberated, binary, ternary, etc. using a
particle tracking technique (e.g. Lamberg and Vianna 2007). This is the approach chosen to describe the ground feed
composites in this study. It should be noted that the original grain size of the main minerals could not be studied in
the ground material since the particle size is reduced to a P 80 of ca. 50 μm. For example, the grain size of pyrite
generally extends above 50 μm (Table 2.1). Furthermore, the particle-tracking approach will not give textural
information on complex intergrowth of particles such as stockworks, coating and emulsion (cf. Pérez-Barnuevo et
al. 2013). A second option (2) to describe textural information was suggested by Pérez-Barnuevo et al. (2013).
Particles are analysed with a linear intercepts method, from which perimeters, specific surface area and linear
grades, are computed, and then used to calculate a set of indices classifying particles into different textural types.
The quantification of different textural types allows estimating the impact of a certain particle type on the
concentration processes (Pérez-Barnuevo et al. 2013). The amount of ternary or more complex particles has been
found low in the Rockliden composites, generally below 5%, and complex textural types are found rarely in
particles documented by MLA analysis. Assessing complex textural types would be interesting only in the case of
complex intergrowths of gudmundite with chalcopyrite, pyrrhotite and galena in the Zn-dominated massive sulphide
composite of the lower part of the Rockliden deposit. However, these textures are so fine grained that it is
challenging to correctly record them with MLA measurements.

So far, textural information is suggested to be implemented in a geometallurgical model using host rock
classification and massive sulphide types as outlined by qualitative ore characterisation (section 2.1, Paper II). The
simplified classification scheme for massive sulphides, as proposed in figure 5.3, might be sufficient to calculate the
bulk mineralogy in the exploration stage of the Rockliden deposit.

49
5.2.2 Flotation-related comminution model

In order to complete a particle-based geometallurgical model, a particle-breakage model should be derived from
comminution tests (Fig. 5.2). The particle-breakage model should define the particle size distribution and liberation
distribution based on feed characteristics (e.g. bulk mineralogy and mineral textures) obtained from a process-
adapted geological model and the outcome of the particle-breakage model is to be used in a unit process model
(Lamberg 2011; Lund 2013). Parameters such as rock hardness, crushing and grinding indices, energy consumption
and throughput will be critical in a geometallurgical model, and the throughput influencing the residence time in the
flotation circuit (Bulled and McInnes 2005; Mwanga et al. 2015b). Measuring and simulating of throughput is
particularly critical to large-scale operations e.g. for porphyry Cu deposits (Cropp et al. 2013; Keeney et al. 2011).

Parameters describing the crushability and grindability do not provide information on the composition of the feed in
terms of particles. Since the main focus of grinding is to achieve liberation, required for effective mineral separation
and further utilisation of the resource, a new term liberability was suggested by Mwanga et al. (2015b) to describe
the “ease with which an ore mineral can be liberated by grinding”. The liberability was shown in graphs plotting the
liberation degree vs. specific grinding energy (Mwanga et al 2015b). Further, Mwanga et al. (2015b) suggested not
to fix the particle size for the final comminution product when a big variation in micro-texture and liberation size is
reported. This proposal might also be relevant to massive sulphide deposits. For example at the George Fisher Zn–
Pb sulphide ore deposits, coarse grained massive galena and sphalerite samples showed simple processing
characteristics, but regrinding was recommended when finer grained micro-textures occurred in order to improve the
recovery of base-metal sulphides from e.g. pyrite (Bojcevski et al. 1998; Bojcevski 2004). Similarly, variation of
pyrite grade and grain size is documented for massive sulphide classes outlined in this study (Table 2.1, Paper II).

The degree of preferential breakage in grinding of the valuable minerals compared to gangue minerals will be
critical to the shape of the particle size distribution curve (Runge et al. 2013). In this study, the largest difference in
the shape of the particle size distribution was noticed between mafic dyke and massive sulphide composites, with
the mafic dyke composite having a flatter curve than the massive sulphide composites. However, this is probably
related to mineral size distribution of the main minerals of the composites, i.e. mafic dyke composite dominated by
silicates and massive sulphide composites dominated by pyrite (Table 4.2, Fig. 4.1), rather than preferential grinding
of the target minerals (chalcopyrite and sphalerite). However, the bulk mineralogy of the Rockliden composites
showed that the grade of hard minerals such as pyrite and NSG is highest in the coarsest size fraction of the feed
while relatively soft minerals such as sphalerite, chalcopyrite and Sb minerals have the highest content in the finest
size fraction. This indicates that the breakage is selective on minerals (cf. Lund 2013).

Hunt et al. (2011) used mineral maps (from multilayer images recorded by optical microscopy), which were
processed by a chessboard segmentation algorithm at different pre-defined particle sizes, to estimate the liberation
degree and the related flotation potential. By testing material from an IOCG deposit it was found that liberation
potential for valuable minerals is most variable for low grade ores (Hunt et al. 2011). As pointed out above, the
acquisition of mineral maps would require SEM-based automated tools in the Rockliden case. Besides this,
segmentation algorithms, mathematical more advanced than the chessboard model, are known in the literature and
e.g. a Boolean model with Poisson polyhedra grains are suggested to be used in case of complex textures (Barbery
and Leroux 1988). However, random breakage models should not be used when selective breakage between
minerals is observed (Lund 2013 and references therein), as it is in Rockliden (cf. Table 4.2). Choi et al. (1988)
developed a model in order to estimate parameters for breakage distribution and breakage rate functions from image
analysis on sized samples for a binary ore system. It was possible to simulate the size distribution and particle
composition in the binary system. However, the Rockliden mineralogy is much more complex and the direct
application of the model built by Choi et al. (1988) would require strong simplification, e.g. rejecting the distinction
of Sb minerals, and thus has not been used.

Lamberg and Lund (2012) developed a calculation algorithm to produce a liberation distribution for a feed sample
for which only (unsized) bulk mineralogy is given. The bulk composition by size and the liberation distribution are
calculated based on the particle information as it is given by archetype samples. The liberation distribution is
adjusted iteratively until the bulk mineralogy reaches the required target. To apply this algorithm to an ore sample,

50
the texture (type) of the incoming ore material has to be recorded in order to choose the most appropriate archetype
for simulating the particle composition after comminution. The approach chosen by Lamberg and Lund (2012)
assumes that the liberation distribution is conserved in narrow size fractions regardless of the overall particle size
distribution and breaking mechanism (cf. Vizcarra et al. 2010). Within this study only four composite samples with
end-member mineralogy were tested. These four composites might serve as archetypes themselves. However, the
approach suggested by Lamberg and Lund (2012) cannot be validated for the Rockliden project due to limited
sample material.

Within this study no data was collected specifically to estimate comminution parameters. However, general trends
have been determined for other base-metal sulphide deposits. For example at the George Fisher Zn–Pb sulphide ore
deposit, ore hardness was attributed to textural end members which were sorted from softest to hardest: massive
galena, massive galena/sphalerite, massive sphalerite, banded sphalerite and banded pyrite (Bojcevski et al. 1998;
Bojcevski 2004). Comminution tests on bulk samples representing Rockliden ore material indicate that an increasing
content of Zn (and hence sphalerite) is one factor causing easier grinding (Bolin, pers. comm., 2014-04-08). Two
massive sulphide samples, with different Zn and sphalerite content, were collected from the Rockliden outcrop and a
laboratory test was conducted on these samples (Mwanga, pers. comm. 2016-01-28). Similar Bond index values
were obtained for the grindability of the two samples. This does not support observations made on the other
Rockliden samples. However, the limited number of samples in the Rockliden study does not allow drawing a
conclusion for the Rockliden deposit so far.

Currently, no comminution model is available which would provide particle size distribution and liberation
distribution in the Rockliden case study. However, it was observed that with a higher grade of a mineral in the feed,
a higher degree of liberation can be obtained for the main and trace minerals shown in figure 5.4a (cf. Table 4.2 and
Fig. 4.2). In order to validate this observation, it is recommended to study the impact of micro-textures on the
liberation distribution, by testing more ore samples. Figure 5.4 might merely point to the possibility to develop a
breakage model, e.g. based on archetypes as proposed by Lamberg and Lund (2012). However, it might also serve
as a qualitative assessment whether a mineral is expected to show a high or low degree of liberation based on the
grade of the mineral in the sample (Fig. 5.4).

Besides the degree of liberation, the particle composition of binary, ternary and more complex particles ought to be
provided by a particle-breaking model (Fig. 5.2). As pointed out previously (see section 4.1 and 5.1), the association
indexes of a mineral pair are different for the composites (e.g. chalcopyrite and gudmundite, Table 4.4). Thus, it is
suggested that the composites might serve as archetypes and that the abundance of certain particle types might be
similar if an untested composite has similar mineralogical composition and textural characteristics as the studied
composites. However, the limit of four composites with mineralogical end-member characteristics does not allow
proofing this idea. Further, not all textural classes outlined for massive sulphides in the ore characterisation study
(Table 2.1) have been studied in laboratory tests. Particle information should not be derived from the association
index for samples with a combination of main and trace Sb mineralogy different to the four composites.

51
Fig. 5.4. (a) Comparison of degree of liberation by size fraction in feed of the four composites (UR-Cu, UR-Zn, LR-Zn and
MD). The grade (wt%) of each mineral in the corresponding composite is given besides the degree of liberation in the bulk. (b)
Bulk degree of liberation (Deg-lib) vs grade of mineral (in wt%) of the four composites. Mineral abbreviations: Bour –
bournonite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene – meneghinite, NSG – non-sulphide gangue (silicates), Po
– pyrrhotite, Py – pyrite (with arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite. For other abbreviations see Fig. 5.1.

52
5.3 Model Alternatives

5.3.1 Outlining alternative approaches

As pointed out in section 5.2, neither a complete process-adapted geological model nor a particle-breakage model is
currently available for the Rockliden deposit. Both components are critical to build a geometallurgical model (Fig.
5.2). Regarding the unit process model, a limited number of composites were tested and more flotation tests on
mixtures of the end-member ore types as well as repetition tests ought to be done to simulate the production mill
feed accurately enough to be used for mine planning (Baum et al. 2004).

Model alternatives mentioned in previous chapters will be summarised here. Generally, there are two groups of
alternatives to build a geometallurgical model for an ore deposit. A first approach is through relating rock intrinsic
parameters with processing proxies using statistical tools, i.e. developing empirical equations connecting parameters
along the processing chain, e.g. using laboratory flotation tests calibrated by pilot plant studies, benchmark
surveying etc. (e.g. Chauhan et al. 2013; Keeney et al. 2011; Lotter et al. 2011). The particle-based approach is
counted to the second group of alternatives, and has already been discussed in the previous section.

In order to formulate a model using the statistical approach, a database sufficiently large to apply statistical tools
needs to be build which means that a reasonably big number of laboratory tests need to be conducted. Kenney et al.
(2011) showed that comminution models based on assay only, give poorer accuracy than using additional
petrophysical data and results from comminution tests. Similarly, it is expected that only using drill core assays in a
flotation model would give poorer precision of the simulated Sb grade then using additional mineralogical or
petrographic information. However, quantitative data additionally to assay data is not (consistently) available in the
Rockliden case. This means that mineralogical and petrographic data and test parameters would need to be collected
on a large number of samples to build a 3D model forecasting process parameters. The limited number of samples
tested in the Rockliden case allows not to build a model forecasting the recovery based on sound statistics (e.g.
Walters 2011). Similarly to the approach described by Kenney et al. (2011), Hunt et al. (2013) demonstrated the use
of mineralogical parameters (including lithology and alteration) derived from drill core logging for building a
comminution model using empirical equations. To find a statistical sound relation between geological data and
comminution indices, a sufficient number of comminution tests need to be conducted (Hunt et al. 2013).

Considering separation processes, Bulled and McInnes (2005) suggested to run a full flotation test per ore type and
to couple the full test with a simplified (“mapping”) flotation test for each ore block. The results of the mapping
flotation test were calibrated by the results of the full flotation test. They suggested to have 15% full and 85%
“mapping” flotation test in the case of an ore deposit with a simple mineralogy. However, it requires that a
“mapping” flotation test is conducted for every mining block. Thus, the amount of samples which would need to be
tested might be similarly large as for the statistical approach.

For the statistical approach, it should be noted that processing parameters and proxies, such as inherent floatability
of the ore, are often non-additive and it requires special considerations when using such parameters to build a
geometallurgical model (e.g. Deutsch 2013; Dunham and Vann 2007). On the other hand, a particle-based
geometallurgical model would require the modelling of textural information as this is part of the output of a process-
adapted geological model (Fig. 5.2). Generally, the modelling of complex geologic relationships of an ore deposit
using geostatistical tools is found very challenging (Barnett and Deutsch 2012). Newton and Graham (2011)
suggested to test different methods including the calculation of variograms for modelling of classes defined by
Principal Component Analysis. The spatial treatment of categorical variables e.g. by indicator kriging (Newton and
Graham 2011), might be extended to modelling of textural types in massive sulphides. However, in the Rockliden
case, variations in textures occur on a centimeter to decimeter scale (Paper II). So far, the determination of micro-
textural types of the Rockliden massive sulphides is based on optical microscopy, which allows for example
studying variation in grain size of pyrite. The fine scale of the textures is much below the scale of theoretical mining
blocks. However, in order to use textual types in a mining model the proportion of each meso- or micro-textural type
need to be determined which are found within theoretical mining blocks. Knowing the amount of textural types and

53
their mineralogy, and given a particle breakage model is available, this would allow generating particle information
for each mining block. Knowing the proportion of particle types per mining block, processing parameters (e.g.
recovery) could be estimated using unit process models. Thus, a non-additive character of processing parameters
would not be of concern in a complete particle-based geometallurgical model.

5.3.2 Evaluation of increased precision in modelling using mineralogical data

It has been generally stated that with increasing depth in mineralogical information more precise models can be
derived (Gu et al. 2014; Lishchuk et al. 2015). More precise models will reduce uncertainty about decisions making
and give a better basis for mine planning and management. Gu et al. (2014) proposed to calculate the value of
mineralogical information as the difference between expected opportunity loss due to wrong decision before precise
information was implemented, and the expected opportunity loss after precise information was available or
implemented in a geometallurgical model.

In order to evaluate the value of mineralogical information in a particle-based geometallurgical model at Rockliden,
one would need a process-adapted geological model, a particle-breakage model and a unit process model (Fig. 5.2),
e.g. to model the Sb content of final flotation products. So far, only a comparison of different simulation options
(SL0 to SL3, cf. Table 3.2) for a flotation model (unit process model) in the Rockliden case has been made (Paper
IV). SL1 and SL2 were based on bulk mineralogy and SL3 was based on full particle information (with the
limitation that the mineral-specific k values were modelled only for distribution of liberated particles). As discussed
above, currently no distinct improvement can be documented when simulation option SL3 is compared to SL1 and
SL2. The lack of improvement in precision of the simulation of option SL3 was partly attributed to challenges in
correctly modelling of the behaviour of binary and more complex particles. This is part of on-going research. The
final value of mineralogical information cannot be determined currently since a complete (particle-based)
geometallurgical model could not be developed for the Rockliden deposit solely by this study.

However, the usage of bulk mineralogy (SL1 and SL2) is expected to be sufficient for building a process model
precise enough to be used as part of a geometallurgical model in the exploration and scoping stage of the Rockliden
resource. For the estimation of the Sb grade and recovery into the Cu–Pb concentrates, it is recommended to
distinguish the Sb minerals in the incoming ore. So far, SEM-based measurements are seen as the most efficient
method in recording the Sb mineralogy in this study. Disadvantages of SEM-based automated measurements are that
sample preparation and analysis require advanced analytical techniques and specially trained personal (cf. section
3.3.3). Further, to measure the same amount of material as with chemical analysis will require measurement of
several standard-size epoxy blocks (cf. section 3.2 and section 3.3.3). Analytical and data processing time are
expected above analysis by chemical assays. However, selecting for example simple Point X-ray analysis for MLA
measurements (cf. Fandrich et al. 2007) might slightly shorten the analytical time. Simple Point X-ray analysis
might be sufficient for the purpose of distinguishing the main Sb minerals since these minerals show a rather wide
spread in mean atomic number in BSE images (Fig. 2.4). The only exceptions are meneghinite and boulangerite,
which have similar mineral chemistry and similar expected flotation behaviour, and thus were combined in this
study.

Different plant modifications need to be considered in order to evaluate the best processing option (Gu et al. 2014).
Thus, the actual values of mineralogical information and a geometallurgical model may also be considered in the
context of the chosen plant modification. In order not to face too high Sb grades in the Cu–Pb concentrates derived
from the Rockliden deposit, different process solutions are thinkable: e.g., blending of ore material before flotation
or of Cu–Pb concentrates after flotation, production of a Pb concentrate extracting Pb–Sb sulphosalts from the Cu
concentrate by selective flotation, or roasting and alkaline sulphide leaching (cf. Fig. 1.1). Each solution has
different investment factors depending on the amount of material which is expected to be treated by the chosen
processing option and the investment in appropriately sized facilities. Despite the fact that the economic impact
cannot be currently evaluated, the relevance of detailed knowledge of Sb mineralogy in order to estimate Sb grades
and recoveries into products is obvious under all suggested treatment options.

54
An economic evaluation of a statistical approach is not given in this study. Since the number of composite samples
that are tested in a laboratory flotation test is limited, statistical tools cannot be applied in order to extract
mineralogical or petrographic parameters which will have most control on the processing results. If the geological
(rock-intrinsic) parameters controlling processing parameters cannot be outlined, it is also not possible to judge on
the economic impact of systematic collection of relevant rock-intrinsic and corresponding processing parameters.

Thus, an economical comparison of the statistical and the particle-based approach cannot be given. It might
however, be noted that a particle-based approach considers the behaviour of physical entities in mineral processing
i.e. the particles (also termed bridge between geology and metallurgy (Lamberg 2011)), whereas the statistical
approaches would give for example a numerical equation connecting rock- and ore-intrinsic parameters to
processing proxies, and thus it might give poorer forecasts when the process or plant operation is changed compared
to the time when processing proxies were recorded.

55
6 Conclusions and Recommendations for Future Work

Detailed ore characterisation and automated SEM-based measurements were used to study the complex trace
mineralogy of antimony (Sb) of different rock and ore types present in the Rockliden Zn–Cu volcanic-hosted
massive sulphide deposit (Paper I and II). Based on the results of the ore characterisation (Paper I and II), four
composites were collected from drill cores: three Sb-rich end-member massive sulphide types and one mafic dyke
composite. Systematic distribution of Sb minerals as liberated and locked particles was documented in a laboratory
flotation test of four composite samples (Paper III). This confirms the results from earlier semi-quantitative work
done on composite samples representing average ore material from three sulphide lenses of the Rockliden deposit
(Bolin 2010) and a quantitative study considering the Sb mineralogy of flotation products at the Rakkejaur deposit
(Lager and Forssberg 1990). These previous studies do not consider mineralogical factors such as the description of
particles by the degree of liberation and type of locking. The relevance of such mineralogical factors was pointed
out in Paper III and they are suggested to be implemented in a unit process model as a component of a possible
particle-based geometallurgical model for the Rockliden deposit (Paper IV). However, a complete particle-based
geometallurgical model would require systematic collection of bulk mineralogy and textural data as minimum
information from a process-adapted geological model of the deposit, as well as the development of a precise
particle-breakage model and a unit process model (cf. Lamberg 2011; Lund 2013).

Several simulation options have been tested to develop a unit process model (Paper IV). The options were
distinguished based on the depth of mineralogical information used in simulation. It was found that simulation based
on bulk mineralogy gives estimations on the Sb grade and recovery of the Cu–Pb concentrate precise enough to be
applicable in the exploration and scoping stage of a mining project (Paper IV). It is however, required to distinguish
at least semi-quantitatively the Sb minerals throughout the deposit. SEM-based automated mineralogy is considered
the best analytical option to determine the trace mineralogy of Sb. The bromine-methanol sequential leaching
method could not distinguish Sb minerals, and thus is not an alternative to SEM-based automated mineralogy for the
purpose of this study. The Sb mineralogy has not yet been determined systematically throughout the Rockliden
deposit and only Sb assays are recorded on regular basis. Naturally, this does not allow distinguishing and
quantifying Sb minerals. Based on qualitative and quantitative ore characterisation (Paper I to III), an element to
mineral conversion scheme was developed, which refines the Sb mineralogy using single Sb assays. It takes into
account that the Sb mineralogy of the massive sulphides is expected to change with depth and that Sb mineralogy is
distinct for Sb-rich altered felsic and mafic rocks. A fixed ratio of Sb minerals for the Sb-rich rock types and
massive sulphide classes was proposed for the refined element to mineral conversion, which allows estimating the
bulk Sb mineralogy.

A statistical modelling approach is discussed as an alternative to the particle-based geometallurgical model. In the
former approach, statistical tools are applied to establish a link between detailed information on process-relevant
rock-intrinsic parameters and corresponding processing parameters. In this study, processing parameters were
obtained for four composite samples and this limited number does not allow building a geometallurgical model
using a statistical approach. More laboratory tests would need to be conducted on several small-scale (drill core)
samples in order to record proxies, e.g. estimating the recovery, with a sample density sufficient to cover the ore
variability over the spatial extent of the deposit (e.g. Chauhan et al. 2013; Mwanga et al. 2015b). Moreover, in
contrast to element or mineral grades, processing proxies are largely non-additive during blending or modelling (e.g.
Deutsch 2013; Dunham and Vann 2007; Newton and Graham 2011). A particle-based model might have the
advantage that the recovery of e.g. Cu, Pb and Sb of each mining block could be simulated directly from the grades
of particle types in a mining block. The grades of particle types might be forecasted by a particle-breakage model
applied to (weighted proportion of) rock and ore types found within a mining block. This approach would avoid
challenges in modelling of non-additive processing parameters in a deposit model. However, the recording and
modelling of textural parameters is challenging too. Textural parameters form a critical input for a particle-based
geometallurgical model. Different options of analysing textural features are discussed. The approach described by
Bojcevski et al. (1998) classifying ore based on meso- and micro-texture types is recommended to be adapted for the
Rockliden deposit and a suggestion for the classification of the massive sulphides is given in Paper II.

56
Beyond all these considerations, it is to state that the content and requirements of the geometallurgical model of a
specific deposit are dependent on process selection and on the question how accurately Sb distribution and its grades
in products needs to be forecasted.

The following list highlights some aspects for which further research is recommended, if a complete particle-based
geometallurgical model should be built for the Rockliden deposit:

x A classical geological model will provide the outline of different rock types and the massive sulphides.
Knowledge on the shape and size of Sb-rich mafic dykes will be helpful since they can increase
concentrations of Sb in Cu–Pb concentrate if they are not sufficiently blended with other ore material
(Paper II). A final process-adapted geological model has not yet been completed for the Rockliden deposit.

x A mineralogical and textural classification of massive sulphides at Rockliden has been proposed (Paper II).
Semi-quantitative recording of textural types might be possible on a meso-scale (cf. Bojcevski 2004).
However, systematic usage of textural ore types has not been implemented in drill core logging at
Rockliden. Ore textures are expected to affect comminution parameters, such as liberability, and hence
flotation results. Thus they are important information which should be provided by a process-adapted
geological model.

x Also, the bulk mineralogy is seen as a critical input parameter to a geometallurgical model, in particular
concerning the Sb mineralogy (Paper I and III). In this study a simplified element to mineral conversion
scheme is proposed to derive the Sb mineralogy of the massive sulphide ore. Further sampling and SEM-
based quantitative ore characterisation studies should be conducted to support or modify the scheme.

x A particle breakage model is not yet ready for the Rockliden deposits. It is suggested that an expected
degree of liberation and mineral association in particles can be estimated based on results from this study.
However, laboratory tests are highly recommended to be conducted, and the recording of parameters such
as comminution energy and throughput.

x To attain higher confidence in the proposed unit flotation model (Paper IV), testing of additional
composites as well as repetition of the laboratory test on composites is recommended for error analysis in
experimental studies.

x Data from pilot scale or plant service should be collected in order to scale-up the unit process model which
is currently based on laboratory test data only. Studying the correlation of flotation cell parameters
influencing the floatability of a particle in the different environments is required for the purpose of up-
scaling or expanding of results from laboratory test to flotation plant conditions (e.g. Niemi 1995).

If finally a complete particle-based model can be presented, it is expected to be a helpful tool in mine planning and
management. Such model might also allow simulating the Sb grade and mineralogy of tailings. Similar to arsenic,
Sb is a potential environmental hazard concerning the discharge to rivers (Wilson et al. 2010). The inclusion of the
consideration of environmental and sustainability issues would broaden the perspective of the geometallurgical
model for Rockliden (cf. Jackson et al. 2011).

Besides antimony (Sb) and arsenic (As), the Rockliden polymetallic Zn–Cu VHMS deposit contains further
deleterious trace elements, which could form additional environmental hazards. Process mineralogical assessment of
other trace elements such as Cd, Hg, Bi, and Sn would require additional mineralogical research. For building a
model accounting for these elements and their mineral hosts, it is recommended to use a mineral grouping different
from the one used in this study for Sb. For example, arsenopyrite (the main As carrier) and pyrite ought not to be
combined if the processing behaviour of As should be studied in detail. So far, it is assumed that the arsenopyrite
and pyrite show similar flotation distribution based on studies of non-particle tracked MLA data (see also
Benzaazoua et al. 2002). The main tin (Sn) minerals are expected to show opposed behaviour during flotation, i.e.
stannite reporting to Cu–Pb concentrate and cassiterite to tailings. Tin might be seen as a valuable element;
however, the concentration is estimated to be low at Rockliden (generally ranging between 20 and 100 ppm, Paper

57
II). Such low grades of Sn are found in e.g. placer deposits, from which Sn can be recovered easily compared to
Rockliden.

Finally, this study also discloses the analytical effort which has to be taken to describe the complex trace mineralogy
such as Sb in a polymetallic ore deposit. It also shows the limitations in applying knowledge about complex
mineralogy in modelling and simulation of the process behaviour of respective trace elements. Including additional
trace elements was beyond the limit of this study. However, this study can be used as a guideline for future research
on remaining trace elements and minerals in the processing of complex sulphide ore at Rockliden and elsewhere.

58
References

Abzalov M. 2008. Quality control of assay data: A review of procedures for measuring and monitoring precision
and accuracy. Exploration and Mining Geology, 17(3-4), 131–144.

$FKLPRYLþRYi0DQG%DODå3.LQHWLFVRIWKHOHDFKLQJRIPHFKDQLFDOO\DFWLYDWHGEHUWKLHULWHERXODQJHULWHDQG
franckeite. Physics and Chemistry of Minerals, 35(2), 95–101.

Alruiz O, Morrell S, Suazo C, Naranjo A. 2009. A novel approach to the geometallurgical modelling of the
Collahuasi grinding circuit. Minerals Engineering, 22(12), 1060–1067.

Anderson CG, Nordwick SM, Krys LE. 1991. Processing of antimony at the Sunshine Mine. In: Residues and
effluents: Processing and environmental considerations. Reddy RG, Imrie WP, Queneau PB, editors. The
Minerals, Metals and Materials Society, 349–366.

Ashley K and Callow M. 2000. Ore variability: Exercises in geometallurgy. Engineering and Mining Journal,
201(2), 24–28.

Awe SA, Khoshkhoo M, Krüger P, Sandström Å. 2012. Modelling and process optimisation of antimony removal
from a complex copper concentrate. Transactions of Nonferrous Metals Society of China, 22(2012), 675–685.

Awe SA. 2013. Antimony recovery from complex copper concentrates through hydro- and electrometallurgical
processes. PhD Thesis, Luleå University of Technology.

Bachmann K., Bartzsch A, Gutzmer J. 2014. Discrimination of hematite and magnetite in finely intergrown natural
iron ores by automated mineralogy. EMAS 2014 11th regional workshop on electron probe microanalysis of
materials today, Leoben, Austria.

%DOiå 3 DQG $FKLPRYLþRYi 0  6HOHFWLYH OHDFKLQJ RI DQWLPRQ\ DQG DUVHQLF IURP PHFKDQLFDOO\ Dctivated
tetrahedrite, jamesonite and enargite. International Journal of Mineral Processing, 81(1), 44–50.

%DOiå3.DPPHO5.XãQLHURYi0$FKLPRYLþRYi00HFKDQR-chemical treatment of tetrahedrite as a new


non-polluting method of metals recovery. In: Hydrometallurgy’94. Springer Netherlands, 209–218.

Barbery G and Leroux D. 1988. Prediction of particle composition distribution after fragmentation of heterogeneous
materials. International Journal of Mineral Processing, 22(1), 9–24.

Barnett RM and Deutsch CV. 2012. Practical implementation of non-linear transforms for modeling
geometallurgical variables. In: Geostatistics Oslo 2012, Springer, 409–422.

Baum W. 2014. Ore characterization, process mineralogy and lab automation a roadmap for future mining. Minerals
Engineering, 60, 69–73.

Baum W, Lotter NO, Whittaker PJ. 2004. Process mineralogy-A new generation for ore characterisation and plant
optimization. SME Annual Meeting 04-12, 1–5.

Bazin C, Hodouin D, Blondin RM. 2013. Estimation of the variance of the fundamental error associated to the
sampling of low grade ores. International Journal of Mineral Processing, 124, 117–123.

Benzaazoua M, Marion P, Liouville-Bourgeois L, Joussemet R, Houot R, Franco A, Pinto A. 2002. Mineralogical


distribution of some minor and trace elements during a laboratory flotation processing of Neves-Corvo ore
(Portugal). International Journal of Mineral Processing, 66(1), 163–181.

59
Björnberg A, Holmström SA, Lindkvist G, inventors. 1986. Method for processing copper smelting materials and
the like containing high percentages of arsenic and/or antimony. Patent US4626279 A, US 06/609989.

Bojcevski D. 2004. Metallurgical characterisation of George Fisher mesotextures and microtextures. Master Thesis,
University of Queensland, Australia.

Bojcevski D, Vink L, Johnson NW, Landmark V, Johnston M, Mackenzie J, Young MF. 1998. Metallurgical
characterisation of Georg Fischer ore texture and implications for ore processing. Mine to mill 1998, 11 -14
October 1998, Brisbane, Queensland, Australia, 29–41.

Bolin N-J. 2010. Rockliden - mineralogy investigation of products produced from three different ore samples.
Internal Report Boliden Mines, 7 pg.

Bolin N-J, Brodin P, Lampinen P. 2003. Garpenberg–an old concentrator at peak performance. Minerals
Engineering, 16(11), 1225–1229.

Bonnici NK. 2012. The mineralogical and textural characteristics of copper-gold deposits related to mineral
processing attributes. PhD Thesis, School of Earth Sciences, University of Tasmania.

Bulled D and McInnes C. 2005. Flotation plant design and production planning through geometallurgical modelling.
SGS MINERALS SERVICES, TECHNICAL BULLETIN 2005-3, 1–8.

Butcher AR. 2010. A practical guide to some aspects of mineralogy that affect flotation. In: Flotation plant
optimisation. Greet CJ, editor. Spectrum Series 16 ed. Australasian Institute of Mining & Metallurgy, 83–93.

Byrne M, Grano S, Ralston J, Franco A. 1995. Process development for the separation of tetrahedrite from
chalcopyrite in the Neves-Corvo ore of Somincor SA, Portugal. Minerals Engineering, 8(12), 1571–1581.

Chauhan M, Napier-Munn T, Keeney L, Bradshaw DJ. 2013. Progress in developing a geometallurgy flotation
indicator. The second AusIMM International Geometallurgy conference, 30 September - 2 October 2013,
Brisbane, Australia, 201–206.

Chayes FA. 1944. Petrographic analysis by fragment counting; part 1, the counting error. Economic Geology, 39(7),
484–505.

Chetty D, Gryffenberg L, Lekgetho T, Molebale I. 2009. Automated SEM study of PGM distribution across a UG2
flotation concentrate bank: Implications for understanding PGM floatability. The Journal of The Southern
African Institute of Mining and Metallurgy, 587–593.

Choi WZ, Adel GT, Yoon RH. 1988. Liberation modeling using automated image analysis. International Journal of
Mineral Processing, 22(1), 59–73.

Cook NJ, Ciobanu CL, Pring A, Skinner W, Shimizu M, Danyushevsky L, Saini-Eidukat B, Melcher F. 2009. Trace
and minor elements in sphalerite: A LA-ICPMS study. Geochimica et Cosmochimica Acta, 73(16), 4761–
4791.

Correia MJ, Carvalho J, Monhemius J. 2001. The effect of tetrahedrite composition on its leaching behaviour in
FeCl 3 -NaCl-HCl solutions. Minerals Engineering, 14(2), 185–195.

Cropp AF, Goodall WR. Bradshaw DJ. 2013. The influence of textural variation and gangue mineralogy on
recovery of copper by flotation from porphyry ore – A review. The second AusIMM International
Geometallurgy conference, 30 September - 2 October 2013, Brisbane, Australia, 279–291.

Deutsch CV. 2013. Geostatistical modelling of geometallurgical variables - problems and solution. The second
AusIMM International Geometallurgy conference, 30 September - 2 October 2013, Brisbane, Australia, 7–15.

60
Dobby G and Savassi ON. 2005. An advanced modelling technique for scale-up of batch flotation results to plant
metallurgical performance. SGS MINERALS SERVICES, TECHNICAL PAPER 2005-09, 1–6.

Dunham S and Vann J. 2007. Geometallurgy, geostatistics and project value—does your block model tell you what
you need to know? Project evaluation conference, Melbourne, Victoria, 189–196.

Eriksson L, Kettaneh-Wold N, Trygg J, Wikström C, Wold S. 2006. Multi-and megavariate data analysis: Part I:
Basic principles and applications. Umetrics Inc.

Fandrich R, Gu Y, Burrows D, Moeller K. 2007. Modern SEM-based mineral liberation analysis. International
Journal of Mineral Processing, 84(1), 310–320.

Farrel JN, Miller AD, Gaze RL. 2011. Geometallurgical sampling and resource estimation for magnetite deposits.
The first AusIMM international geometallurgy conference, 5-7 September 2011, Brisbane, Australia, 311–319.

Ficeriová J, Baláž P, Boldižárová E. 2005. Combined mechanochemical and thiosulphate leaching of silver from a
complex sulphide concentrate. International Journal of Mineral Processing, 76(4), 260–265.

Figueroa G, Moeller K, Buhot M, Gloy G, Haberla D. 2012. Advanced discrimination of hematite and magnetite by
automated mineralogy. In: Proceedings of the 10th International Congress for Applied Mineralogy (ICAM),
197–204.

Filippou D, St-Germain P, Grammatikopoulos T. 2007. Recovery of metal values from copper—arsenic minerals
and other related resources. Mineral Processing and Extractive Metallurgy Review: An International Journal,
28(4), 247–298.

Ghorbani Y, Becker M, Petersen J, Mainza AN, Franzidis J-P. 2013. Investigation of the effect of mineralogy as
rate-limiting factors in large particle leaching. Minerals Engineering, 52, 38–51.

Gorain B, Franzidis J, Ward K, Johnson N, Manlapig E. 2000. Modelling of the Mt Isa rougher-scavenger copper
flotation circuit using size-by-liberation data. Minerals and Metallurgical Processing, 17(3), 173–180

Gottlieb P, Wilkie G, Sutherland D, Ho-Tun E, Suthers S, Perera K, Jenkins B, Spencer S, Butcher A, Rayner J.
2000. Using quantitative electron microscopy for process mineralogy applications. JOM, 52(4), 24–25.

Grammatikopoulos TA and Roth T. 2002. Mineralogical characterization and Hg deportment in field samples from
the polymetallic Eskay Creek deposit, British Columbia, Canada. International Journal of Surface Mining,
Reclamation and Environment, 16(3), 180–195.

Grammatikopoulos TA, Valeyev O, Roth T. 2006. Compositional variation in Hg-bearing sphalerite from the
Polymetallic Eskay creek deposit, British Columbia, Canada. Chemie Der Erde - Geochemistry, 66(4), 307–
314.

Gu Y, Schouwstra RP, Rule C. 2014. The value of automated mineralogy. Minerals Engineering, 58, 100–103.

Helle S, Kelm U, Barrientos A, Rivas P, Reghezza A. 2005. Improvement of mineralogical and chemical
characterization to predict the acid leaching of geometallurgical units from Mina Sur, Chuquicamata, Chile.
Minerals Engineering, 18(13), 1334–1336.

Hunt J, Lottermoser BG, Parbhakar-Fox A, Van Veen E, Goemann K. 2016. Precious metals in gossanous waste
rocks from the Iberian Pyrite Belt. Minerals Engineering, 87, 45–53.

Hunt J, Kojovic T, Berry R. 2013. Estimating comminution indices from ore mineralogy, chemistry and drill core
logging. The second AusIMM International Geometallurgy conference, 30 September - 2 October 2013,
Brisbane, Australia, 173–176.

61
Hunt J, Berry R, Bradshaw D. 2011. Characterising chalcopyrite liberation and flotation potential: Examples from
an IOCG deposit. Minerals Engineering, 24(12), 1271–1276.

Ikumapayi FK. 2013. Recycling process water in complex sulphide ore flotation. PhD Thesis, Luleå University of
Technology.

Jackson J, McFarlane AJ, Olson Hoal, K. 2011. Geometallurgy - back to the future: Scoping and communicating
Geomet programs. The first AusIMM International Geometallurgy conference, 5-7 September 2011, Brisbane,
Australia, 125–131.

Keeney L and Walters SG. 2011. A methodology for geometallurgical mapping and orebody modelling. The first
AusIMM International Geometallurgy conference, 5-7 September 2011, Brisbane, Australia, 217–225.

Keeney L, Walters SG, Kojovic T. 2011. Geometallurgical mapping and modelling of comminution performance at
the Cadia East porphyry deposit. The first AusIMM International Geometallurgy conference, 5-7 September
2011, Brisbane, Australia, 73–83.

Knight R, Olson Hoal K, Abraham APG. 2011. Three-dimensional geometallurgical data integration for predicting
concentrate quality and tailings composition in a massive sulfide deposit. The first AusIMM International
Geometallurgy conference, 5-7 September 2011, Brisbane, Australia, 227–232.

Knorr DB and Breedis JF, inventors. 1986. Copper alloys having an improved combination of strength and
conductivity. Patent CA 1255124 A1, CA 489814.

Kuhar LL, Jeffrey MI, McFarlane AJ, Benvie B, Botsis NM, Turner N, Robinson DJ. 2011. The development of
small-scale tests to determine hydrometallurgical indices for orebody mapping and domaining. The first
AusIMM International Geometallurgy conference, 5-7 September 2011, Brisbane, Australia, 335–345.

Kwitko-Ribeiro R. 2012. New sample preparation developments to minimize mineral segregation in process
mineralogy. In: Proceedings of the 10th international congress for applied mineralogy (ICAM), Springer, 411–
417.

Lager T. 1989. The technology of processing antimony-bearing ores. PhD thesis, Luleå University of Technology.

Lager T. 1985. The Rakkejaur deposit, a challenge to flotation process technology. in: Flotation of Sulphide
Minerals, Forssberg KSE (editor), Stockholm, Sweden, 255–275.

Lager T and Forssberg K. 1990. Separation of antimony mineral impurities from complex sulphide ores.
Transactions of the Institution of Mining and Metallurgy (Section C: Miner. Process. Extr. Metall.), 99, 54–61.

Lamberg P, Hautala P, Sotka S, Saavalainen S. 1997. Mineralogical balances by dissolution methodology. IMA
COM Short course, Portugal, 29 pg.

Lamberg P. 2011. Particles – the bridge between geology and metallurgy. In: Preprints Conference in Minerals
Engineering 2011, 16 pg.

Lamberg P and Lund C. 2012. Taking liberation information into a geometallurgical model – case study
Malmberget, Northern Sweden. Process mineralogy ‘12 conference, Cape Town, South Africa, 1–13.

Lamberg P and Vianna S. M. 2016. Mass balancing mineralogical data. Chapter, 17 in: Process mineralogy,
Monograph by University of Queensland, (in print).

Lamberg P and Vianna S. M. S. 2007. A technique for tracking multiphase mineral particles in flotation circuits. In:
VII meeting of the southern hemisphere on mineral technology Ouro Preto, Minas Gerais, Brazil, 195–203.

Larouche P. 2001. Minor elements in copper smelting and electrorefining. Master Thesis, Mining and Metallurgical
Engineering, McGill University, Montreal, Canada.

62
Lastra R. 2007. Seven practical application cases of liberation analysis. International Journal of Mineral Processing,
84, 337–347.

Liipo J, Lang C, Burgess S, Otterstrom H, Person H, Lamberg P. 2012. Automated mineral liberation analysis using
INCAMineral. Process mineralogy ‘12 conference, Cape Town, South Africa, 1–7.

Lishchuk V, Lamberg P, Lund C. 2015a. Classification of geometallurgical programs based on approach and
purpose. Proceedings of the 13th biennial SGA meeting (Mineral resources in a sustainable world), Nancy,
France, 1431–1435.

Lishchuk V, Koch P-H, Lund C, Lamberg P. 2015b. The geometallurgical framework: Malmberget and
Mikheevskoye case studies. Conference of doctoral students and young scientists Szklarska Poreba, Poland, 12
pg.

Lorenzen L and Barnard MJ. 2011. Why is mineralogical data essential for designing a metallurgical test work
program for process selection and design? The first AusIMM International Geometallurgy conference, 5-7
September 2011, Brisbane, Australia, 163–172.

Lotter N, Whiteman E, Bradshaw D. 2014. Modern practice of laboratory flotation testing for flowsheet
development–A review. Minerals Engineering, 66, 2–12.

Lotter N, Kormos L, Oliveira J, Fragomeni D, Whiteman E. 2011. Modern process mineralogy: Two case studies.
Minerals Engineering, 24(7), 638–650.

Lotter N, Kowal D, Tuzun M, Whittaker P, Kormos L. 2003. Sampling and flotation testing of Sudbury Basin drill
core for process mineralogy modelling. Minerals Engineering, 16(9), 857–864.

Lund C. 2013. Mineralogical, chemical and textural characterisation of the Malmberget iron ore deposit for a
geometallurgical model. PhD thesis, Luleå University of Technology.

Lund C and Lamberg P. 2014. Geometallurgy – A tool for better resource efficiency. European Geologist 37, 03
ANNIVERSARY, 39–43.

Mattsson B and Heeroma P. 1985. Rockliden prospekteringsresultat 1981-1985. Internal Report Boliden Mineral
AB, Report No GP 85 005 2, 35pg.

McArthur GJ. 1996. Textural evolution of the Hellyer massive sulphide deposit. PhD thesis, University of
Tasmania.

McArthur GJ. 1988. Using geology to control geostatistics in the Hellyer deposit. Mathematical Geology, 20(4),
343–366.

McClung CR and Viljoen F. 2012. Mineralogical assessment of the metamorphosed Broken Hill sulfide deposit,
South Africa: Implications for processing complex orebodies. Proceedings of the 10th international congress
for applied mineralogy (ICAM), Springer, 427–434.

Menditto A, Patriarca M, Magnusson B. 2007. Understanding the meaning of accuracy, trueness and precision.
Accreditation and Quality Assurance, 12(1), 45–47.

Minz FE. 2013. Mineralogical characterisation of the Rockliden antimony-bearing volcanic-hosted massive sulphide
deposit, Sweden. Licentiate thesis, Luleå University of Technology.

Moëlo Y, Makovicky E, Mozgova NN, Jambor JL, Cook N, Pring A, Paar W, Nickel EH, Graeser S, Karup-Møller
S. 2008. Sulfosalt systematics: A review. Report of the sulfosalt sub-committee of the IMA commission on ore
mineralogy. European Journal of Mineralogy, 20(1), 7–46.

63
Mwanga A. 2014. Test methods for characterising ore comminution behavior in geometallurgy. Licentiate thesis,
Luleå University of Technology.

Mwanga A, Rosenkranz J, Lamberg P. 2015a. Testing of ore comminution behavior in the geometallurgical
context—A review. Minerals, 5, 276–297.

Mwanga A, Lamberg P, Rosenkranz J. 2015b. Comminution test method using small drill core samples. Minerals
Engineering, 72, 129–139.

Mzyk Z, Baranowska I, Mzyk J. 2002. Research on grain size effect in XRF analysis of pelletized samples. X-Ray
Spectrometry 31, 39–46.

Navarro P and Alguacil FJ. 2002. Adsorption of antimony and arsenic from a copper electrorefining solution onto
activated carbon. Hydrometallurgy 66, 101–105.

New Boliden. 2015. New Boliden Annual Report 2014. Stockholm, 124 pg.

Newton MJ and Graham JM. 2011. Spatial modelling and optimisation of geometallurgical indices. The first
AusIMM International Geometallurgy conference, 5-7 September 2011, Brisbane, Australia, 247–261.

Niemi AJ. 1995. Role of kinetics in modelling and control of flotation plants. Powder Technology, 82(1), 69–77.

Parian M. 2015. Development of the mineralogical path for geometallurgical modeling of iron ores. Licentiate
thesis, Luleå University of Technology.

Parian M, Lamberg P, Möckel R, Rosenkranz J. 2015. Analysis of mineral grades for geometallurgy: Combined
element-to-mineral conversion and quantitative X-ray diffraction. Minerals Engineering, 82, 25–35.

Penttinen U, Palosaari V, Siura T. 1977. Selective dissolution and determination of sulfides in nickel ores by the
bromine–methanol method. Bulletin of the Geological Society of Finland, 49, 79–84.

Perez-Barnuevo L, Pirard E, Castroviejo R. 2013. Automated characterisation of intergrowth textures in mineral


particles. A case study. Minerals Engineering, 52, 136–142.

Pérez-López R, Álvarez-Valero AM, Nieto JM, Sáez R, Matos JX. 2008. Use of sequential extraction procedure for
assessing the environmental impact at regional scale of the São Domingos Mine (Iberian Pyrite Belt). Applied
Geochemistry 23(12), 3452–3463.

Petruk W. 2000. Applied mineralogy in the mining industry. Amsterdam: Elsevier Science BV, chapter 4
(Volcanogenic base metal deposits in the Flin Flon-Snow Lake Areas, Manitoba, Canada), 73–101.

Petruk W and Schnarr JR. 1981. An evaluation of the recovery of free and unliberated mineral grains, metals and
trace elements in the concentrator of Brunswick mining and smelting Corp. Ltd. Canadian Mining and
Metallurgical Bulletin, 75(833), 132–159.

Petruk W and Hughson MR. 1977. Image analysis evaluation of the effect of grinding media on selective flotation of
two Zn-Pb-Cu ores. Canadian Mining and Metallurgical Bulletin 70(782), 128–135.

Pitard FF. 1993. Pierre Gy's sampling theory and sampling practice: Heterogeneity, sampling correctness, and
statistical process control. CRC Press.

Pownceby MI, MacRae CM, Wilson NC. 2007. Mineral characterisation by EPMA mapping. Minerals Engineering,
20(5), 444–451.

Raat H and Årebäck H. 2009. Geology, grade and tonnage estimate of the Rockliden VMS deposit, north central
Sweden. Internal Report Boliden Mines, 56 pg.

64
Raat H, Lasskogen J, Ronkainen K. 2011. Results Rockliden diamond drilling 2010. Internal Report Boliden Mines,
17 pg.

Roine A. 2009. HSC Chemistry 7.0 User’s Guide – Chemical Reaction and Equilibrium Software with Extensive
Thermochemical Database and Flowsheet Simulation. Volume 1/2.

Runge KC, Tabosa E, Jankovic A. 2013. Particle size distribution effects that should be considered when performing
flotation geometallurgical testing. The second AusIMM International Geometallurgy conference, 30
September - 2 October 2013, Brisbane, Australia, 335–344.

Runge KC, Franzidis JP, Manlapig EV, Harris MC. 2003. Structuring a flotation model for robust prediction of
flotation circuit performance. Proceedings XXII international mineral processing congress, Cape Town, South
Africa, 973–984.

Runge KC, Harris MC, Frew JA, Manlapig EV. 1997. Floatability of streams around the Cominco Red Dog lead
cleaning circuit. Proceedings of the 6th annual mill operators conference, Madang, Papua New Guinea, 157–
163.

Schouwstra R, De Vaux D, Hey P, Malysiak V, Shackleton N, Bramdeo S. 2010. Understanding Gamsberg–A


geometallurgical study of a large stratiform zinc deposit. Minerals Engineering, 23(11), 960–967.

Spencer S and Sutherland D. 2000. Stereological correction of mineral liberation grade distributions estimated by
single sectioning of particles. Image Analysis and Stereology, 19, 175–182.

Suazo C, Kracht W, Alruiz O. 2010. Geometallurgical modelling of the Collahuasi flotation circuit. Minerals
Engineering, 23(2), 137–142.

Sutherland DN and Gottlieb P. 1991. Application of automated quantitative mineralogy in mineral processing.
Minerals Engineering, 4(7–11), 753–762.

Sylvester PJ. 2012. Use of the mineral liberation analyzer (MLA) for mineralogical studies of sediments and
sedimentary rocks. Mineralogical association of Canada Short Course 42, 1–16.

Trahar W. 1981. A rational interpretation of the role of particle size in flotation. International Journal of Mineral
Processing, 8(4), 289–327.

Trahar WJ and Warren LJ. 1976. The flotability of very fine particles — A review. International Journal of Mineral
Processing, 3(2), 103–131.

Tungpalan K, Manlapig E, Andrusiewicz M, Keeney L, Wightman E, Edraki M. 2015. An integrated approach of


predicting metallurgical performance relating to variability in deposit characteristics. Minerals Engineering,
71, 49–54.

Van den Boogaart KG, Weissflog C, Gutzmer J. 2011. The value of adaptive mineral processing based on spatially
varying ore fabric parameters. Proceedings of IAMG, 1–14.

Vizcarra TG, Wightman EM, Johnson NW, Manlapig EV. 2010. The effect of breakage mechanism on the mineral
liberation properties of sulphide ores. Minerals Engineering, 23(5), 374–382.

Walters S. and Kojovic T. 2006. Geometallurgical mapping and mine modeling (GEM3) - the way of the future.
International autogenous and semi-autogenous grinding technology 2006, Vancouver, Canada, IV411–IV425.

Walters SG. 2011. Integrated industry relevant research initiatives to support geometallurgical mapping and
modelling. The first AusIMM International Geometallurgy conference, 5-7 September 2011, Brisbane,
Australia, 273–278.

65
Wang S. 2004. Impurity control and removal in copper tankhouse operations. JOM Journal of the Minerals, Metals
and Materials Society, 56(7), 34–37.

Welsby S, Vianna S, Franzidis J. 2010. Assigning physical significance to floatability components. International
Journal of Mineral Processing, 97(1), 59–67.

Whiten B. 2008. Calculation of mineral composition from chemical assays. Mineral Processing and Extractive
Metallurgy Review, 29(2), 83–97.

Wilson SC, Lockwood PV, Ashley PM, Tighe M. 2010. The chemistry and behaviour of antimony in the soil
environment with comparisons to arsenic: A critical review. Environmental Pollution 158(5), 1169–1181.

66
PAPER I

Detailed Characterisation of Antimony Mineralogy in a Geometallurgical Context at the


Rockliden Ore Deposit, North-Central Sweden

Friederike Minz, Nils-Johan Bolin, Pertti Lamberg, Christina Wanhainen

Minerals Engineering (2013), Vol. 52, pp. 95–103


Minerals Engineering 52 (2013) 95–103

Contents lists available at SciVerse ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Detailed characterisation of antimony mineralogy in a geometallurgical


context at the Rockliden ore deposit, North-Central Sweden
Friederike Minz a,⇑, Nils-Johan Bolin b, Pertti Lamberg a, Christina Wanhainen a
a
Luleå University of Technology, Department of Civil, Environmental and Natural Resources Engineering, Sweden
b
Boliden, Division of Process Technology, Sweden

a r t i c l e i n f o a b s t r a c t

Article history: The antimony (Sb) content of the Rockliden complex Zn–Cu massive sulphide ore lowers the quality of
Available online 9 May 2013 the Cu–Pb concentrate. The purpose of this study is to characterise the Sb mineralogy of the deposit.
The Sb-bearing minerals include tetrahedrite (Cu,Fe,Ag,Zn)12Sb4S13, bournonite PbCuSbS3, gudmundite
Keywords: FeSbS and other sulphosalts. On a microscopic scale these minerals are complexly intergrown with
Sulphide ores base-metal sulphides in the ore. Based on these observations mineralogical controls on the distribution
Ore mineralogy of Sb-bearing minerals in a standard flotation test are illustrated. Deposit-scale and rock-related variation
Antimony
in the Sb-content and distribution of Sb-bearing minerals were found. This underlines the importance in
Mineral processing
North-Central Sweden
understanding the geological background as a basis of a 3D geometallurgical model for Rockliden. Such a
model is expected to predict the Sb content of the Cu–Pb concentrate, among other process-relevant fac-
tors, and helps to forecast when the Cu–Pb concentrate has to be treated by alternative processes, such as
alkaline sulphide leaching, before it is sold to the smelter.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction (Anderson, 2012; Butterman and Carlin, 2004). An example is the


Sunshine Mine, Coeur d’Alene mining district, Idaho (Anderson
Antimony (Sb) has no commercial use as metal itself, but it et al., 1991) with Pb–Ag–Zn–Cu–Sb sulphides bearing tetrahe-
finds application in form of metallo-organic and inorganic com- drite-rich quartz–siderite veins (Wavra et al., 1994).
pounds as flame retardants, pigments, heat and radiation stabiliz- Mineral processing techniques traditionally applied to the pri-
ers for plastics and clarification of specialty glasses, as well as Sb- mary, stibnite dominated, Sb ore material are hand-sorting, heavy
based catalysts and Sb alloys, to increase strength and hardness of medium separation, and flotation (Anderson, 2012; Tian-Cong,
Pb (Butterman and Carlin, 2004; Masters, 2005; Anderson, 2012). 1988). The processing of the more complex and polymetallic ore
Given its usage, it was defined as critical raw material for the mod- material requires, in addition to mineral processing, the involve-
ern society (Anderson, 2012; European Commission, 2010). The ment of alternative pyrometallurgical and hydrometallurgical pro-
world market of Sb is dominated by China although the closure cess solutions to extract Sb from the ore material (Anderson, 2012).
of mines in the Hunan Province has caused a decrease in surplus The zinc–copper (Zn–Cu) volcanic-hosted massive sulphide
(Anderson, 2012; Masters, 2005). (VHMS) deposits of the Skellefte district (northern Sweden,
The world’s largest natural resources of Sb, such as those in Chi- Fig. 1) contain Sb as impurity element with grades commonly
na, primarily contain stibnite associated with pyrite, arsenopyrite, above 200 ppm (e.g. Rakkejaur), but generally below 2.5 wt% (Allen
jamesonite, and various oxides if the deposits are affected by oxi- et al., 1996). The Sb grades at the Rockliden massive sulphide min-
dation (Butterman and Carlin, 2004; Tian-Cong, 1988; Table 1). eralisation also fall into this range (Raat and Årebäck, 2009). For
Other Sb deposits are more complex in terms of their metal con- this range in the Sb content pyrometallurgical methods, such as
tent, i.e. they are polymetallic and are primarily mined for com- oxide volatilisation or roasting, which are normally used for pro-
modities such as Cu, Pb, and Ag. Sulphosalts form the dominant duction of Sb from low grade Sb-ores (5–25 wt%) (Anderson,
Sb-bearing minerals (e.g. tetrahedrite, Table 1) in these deposits, 2012 and references therein), are not considered further. However,
accompanied by sulphide minerals such as galena and pyrite attempts have been made to remove Sb from As–Sb–Bi-rich sul-
phide concentrates by partial roasting before the material was to
⇑ Corresponding author. Address: Division of Geosciences and Environmental be sent to the smelter in Rönnskär, Skellefteå, Sweden (Björnberg
Engineering, Department of Civil, Environmental and Natural Resources Engineer- et al., 1986).
ing, Luleå University of Technology, SE-971 87 Luleå, Sweden. Tel.: +46 920 491068, Studies by Lager (1989) on massive sulphides from the
mobile: +46 72 5430958. Rakkejaur deposit (Skellefte district, Sweden) showed that certain
E-mail address: friederike.minz@ltu.se (F. Minz).

0892-6875/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.mineng.2013.04.017
96 F. Minz et al. / Minerals Engineering 52 (2013) 95–103

Table 1 and 71 g/t Ag (New Boliden, 2013). The Cu and Zn contents are
Common Sb bearing minerals (cf. Anderson, 2012; Butterman and Carlin, 2004). fairly constant throughout the steeply dipping mineralisation.
Name Composition However, the Sb content varies. For example, the average Sb-con-
Stibnite Sb2S3 tent of all analysed intersections drilled below ca. 400 m depth is
Jamesonite Pb4FeSb6S14 about half of that from the upper part of the deposit (Raat and
Senarmontite, Valentinite Sb2O3 Årebäck, 2009), ranging on average from ca. 750 to 1500 ppm.
Kermesite Sb2S2O Most of the mineralisation at Rockliden is found at the altered
Tetrahedrite Cu12Sb4S13
Bournonite PbCuSbS3
stratigraphic top of rhyolitic–dacitic volcanic rocks (Depauw,
2009; Mattsson and Heeroma, 1985). The volcanic rocks are
thought to form an inlier, isolated within the metamorphosed sed-
imentary rocks of the Bothnian Basin (Depauw, 2009; Kousa and
Lundqvist, 2000). Silicification and sericitation, and locally carbon-
ate- and chlorite-alteration, have affected the volcanic host rocks
(Depauw, 2009; Raat and Årebäck, 2009). The metamorphic grade
of the volcanic and sedimentary host rocks reaches greenschist fa-
cies (Depauw, 2009), which is low compared to the regional,
amphibolite facies, metamorphic grade (Kousa and Lundqvist,
2000). Folding and faulting resulted in steeply dipping mineralised
zones (Fig. 2) with complex geometry (Evins, 2011; Raat and
Årebäck, 2009). Additionally, NE–SW to ENE–WSW trending faults
and parallel mafic dykes cross-cut the Rockliden mineralisation
(Depauw, 2009). Remobilisation of Sb during the intrusion of the
dykes has been suggested and some of the dykes contain about
1 wt% Sb (Depauw, 2009).
The work presented here is part of the Rockliden geometallurgi-
cal case study. Generally, geometallurgy combines all steps in-
volved in finding and extracting metals from the earth under
current economic conditions (e.g. Dunham et al., 2011; Jackson
et al., 2011; Lamberg, 2011). This general description, however, im-
plies that a broad range of aspects are considered from all involved
disciplines. The Rockliden geometallurgical project focuses on
understanding and gaining knowledge of the response to processes
involved in extracting minerals from this deposit (cf. Jackson et al.,
2011). Process-relevant, i.e. metallurgical-driver, rock-intrinsic
parameters need to be identified (Walters and Kojovic, 2006;
Williams, 2011). Further, the variability of these parameters should
Fig. 1. Location of the Rockliden project south of the Skellefte district.
be captured over the spatial extent of the Rockliden deposit to
outline domains representing homogeneous populations in terms
of the parameters under consideration (cf. Jackson et al., 2011).
(liberated) Sb-bearing minerals (e.g. gudmundite) can be de- These domains form a geometallurgical model to be built for
pressed during Cu–Pb and Zn flotation, whereas other minerals Rockliden and to be used in forecasting the metallurgical response
(e.g. bournonite and tetrahedrite) show flotation properties similar and furthermore in production planning and management.
to chalcopyrite and will report to the Cu–Pb concentrate. Due to Critical process-relevant parameters at Rockliden include the Sb
complex mineralogy, poor liberation and selectivity in flotation, content of the Cu–Pb concentrate, and the purpose of this work is
Sb-bearing minerals end up in Cu–Pb and Zn concentrates. When to gain detailed knowledge of this component of response (cf. Jack-
Sb is transferred to the Cu–Pb concentrate, it lowers the quality son et al., 2011). This study compiles the Sb mineralogy at Rockl-
of this product and causes penalties at the smelter. Thus, solutions iden and outlines controls on the Sb distribution in laboratory
to remove Sb from the Rockliden Cu–Pb concentrate by hydromet- flotation tests. In earlier studies on the Sb mineralogy of Rockliden,
allurgical processes have been considered (Bolin, 2010). Generally, tetrahedrite, bournonite and gudmundite were identified (Karup-
hydrometallurgical methods can be employed to extract Sb from Møller and Makovicky, 1980). The list of Sb-bearing minerals was
polymetallic ore, for example by alkaline sulphide leaching as used extended in this study, and additional information was collected,
at the Sunshine Mine plant, Coeur d’Alene mining district, Idaho such as the grain size of the Sb-bearing minerals and their mineral
(Anderson et al., 1991). Sulphide leaching was tested successfully association, i.e. their mode of occurrence in the uncrushed drill
for Cu–Pb concentrates from the Rakkejaur massive sulphide core material. This data was then used to qualitatively evaluate
deposit and Rockliden (Awe et al., 2012; Lager, 1985). During the liberation distribution of the Sb-bearing minerals as found in
further processing steps Sb could be recovered via electrowinning the flotation products from laboratory flotation tests.
(Anderson, 2012; Anderson et al., 1991; Awe, 2010; Awe and
Sandström, 2012; Filippou et al., 2007).
The Rockliden Zn–Cu massive sulphide deposit is located in 2. Methods
North-Central Sweden, approximately 150 km south of the Skel-
lefte ore district (Fig. 1). The deposit was discovered by Boliden Two types of samples were studied: uncrushed drill cores and
in the 1980s but development was put on hold when flotation tests flotation products from laboratory tests from three composite sam-
showed that the quality of the Cu–Pb concentrate is lowered by the ples. Locations of drill core intersections are shown in Fig. 2.
Sb content (Raat and Årebäck, 2009). However, currently the de- Prior to sample collection for mineralogical studies on uncru-
posit is again under consideration. The inferred mineral resource shed drill core samples, all Rockliden drill core assays available
of Rockliden is 3.53 Mt with 4.2 wt% Zn, 1.9 wt% Cu, 0.7 wt% Pb from the Boliden archive were evaluated to capture the variation
F. Minz et al. / Minerals Engineering 52 (2013) 95–103 97

in Sb grades in relation to the base-metal content. Cu, Zn, Pb, As, S probe studies (Fig. 2), emphasising those with elevated Sb or Ag
and Sb assays were converted into mineral grades of pyrite, chalco- content in the chemical assays.
pyrite, sphalerite, galena, arsenopyrite and bournonite using the Laboratory batch flotation tests were run in 2009 by Boliden
Geo module of the HSC Chemistry software (Roine, 2009). Since Mineral AB to determine the metallurgical response of composite
chalcopyrite and sphalerite are the main base metal carriers at samples (each composed of 15–30 kg mineralised drill core) col-
Rockliden which need to be separated by flotation into Cu–Pb lected from three different parts of the Rockliden mineralisation
and Zn concentrates, the drill core assays were classified based (CA, CB and CC, Fig. 2B). The three composite samples were not
on the chalcopyrite and sphalerite content of the plain sulphide characterised petrographically and mineralogically in any detail
fraction, i.e. sulphides recalculated to 100 wt%. These preliminary prior to processing for this initial study. They were crushed to
classes were further subdivided by their Ag and Sb content, since minus 3.15 mm and divided by a rotary splitter into 1 kg lots.
these metals represent important bonus and penalty elements at Grinding was done in a laboratory rod mill. A laboratory flotation,
Rockliden. Using the preliminary classification as a guide, sixty using the flow sheet shown in Figs. 2 and 5, was done in a 2.7 l cell
samples were collected from six drill cores distributed throughout with a Wemco lab flotation machine. For each composite sample a
the Rockliden deposit (Fig. 2). In this paper these samples are re- final Cu–Pb concentrate, a Zn concentrate and final tailings product
ferred to as mineralogical samples, i.e. samples which were used was collected, assayed and epoxy samples were prepared for
to study primary features of Sb-bearing minerals as found in the microscopic studies (Fig. 2).
uncrushed ore. For each mineralogical sample, a polished thin sec- Scanning electron microscope (SEM) studies were conducted at
tion was prepared and chemical assays were obtained from the the following instruments and institutions: Hitachi SU6600 (FES-
residual drill core material (i.e. ca. 30–50 cm length of quartered EM, NTNU Trondheim, Norway), Ultra Plus – Zeiss Gemini (FESEM,
BQ and NQ drill core per sample). The polished thin sections were Oulu University, Finland), Merlin – Zeiss Gemini (FESEM, Luleå
studied by optical microscopy and 15 thin sections were selected University of Technology, Sweden), and SEM (Research Laboratory,
for scanning electron microscope studies and 6 for electron micro- Geological Survey of Finland (GTK, Helsinki, Finland)). Generally,

Fig. 2. Diagram showing the research approach. Sections (A) and (B) are viewing to the west on the Rockliden massive sulphide ore body (translucent shapes) with the
location of sampled and assayed drill core intervals (white dashed lines). (A) Samples for the mineralogical study (from six drill cores A–F). (B) Composite samples (CA, CB and
CC). Flotation sheet for the laboratory tests below graphic (B) (for more details see Fig. 5). The final products and samples used for microscopic studies are shown bold and
underlined. Abbreviation for microscopic and micro-analytical studies: BSE Back scattered electron imaging, EDS energy-dispersive X-ray spectroscopy, SEM scanning
electron microscope, EPMA electron probe microanalyser, WDS wavelength-dispersive X-ray spectroscopy.
98 F. Minz et al. / Minerals Engineering 52 (2013) 95–103

the acceleration voltage was set to 20 keV and the emission current and concentric zoning in BSE images, which is related to variations
ranged from 0.9 to 1.1 nA for EDS (energy-dispersive X-ray spec- in the Sb content (0–0.5 wt%).
troscopy) analysis. The following analytical conditions were cho- Additionally, other Sb-bearing minerals and sulphosalts such
sen for wavelength-dispersive X-ray spectroscopy (WDS) analysis as berthierite (FeSb2S4), jamesonite (Pb4FeSb6S14), geocronite
at the electron probe microanalyser (EPMA): acceleration voltage (Pb14(Sb,As)6S23), native antimony, pyrargyrite (Ag3SbS3) and
of 15–20 keV, emission current of 15–30 nA and a beam diameter miargyrite (AgSbS2) were found in SEM/EDS study.
of 1–5 lm. EPMA/WDS studies were conducted at the following
facilities: JEOL JXA-8200 microanalyser (Oulu University) and elec-
tron microprobe Cameca SX100 (GTK Helsinki). 3.2. Characterisation of Sb-bearing minerals in Mineralogical Samples

The mineralogical samples are subdivided into (1) disseminated


to semi-massive sulphides and (2) massive sulphide samples based
3. Results on their total sulphide content calculated from chemical assays by
element to mineral conversion (Roine, 2009). Generally, the total
3.1. Chemical composition of Sb-bearing minerals at Rockliden sulphide content of the disseminated to semi-massive sulphides
is on average ca. 15 wt% and for the massive sulphide samples
The average composition of common Sb-bearing minerals at above 50 wt%. The average chemical compositions of the mineral-
Rockliden is listed in Table 2. Analyses of bournonite, meneghinite, ogical samples are listed in Table 3. The main types are further di-
boulangerite and gudmundite, are close to the stoichiometric com- vided into subtypes based on their dominating base metal sulphide
position (Moëlo et al., 2008). Meneghinite and boulangerite have a (chalcopyrite/sphalerite) and their trace element (Ag/Sb) content,
similar chemical composition (Table 2) and show similar grey their host rock composition (mafic/felsic), and their spatial location
scales in BSE images (i.e., similar mean atomic number). Thus, their (upper/deeper part).
distinction is challenging by automated SEM-based modal analysis Massive sulphides consist mostly of pyrite followed by sphaler-
and they consequently are grouped in one mineral class. ite and chalcopyrite. Often sphalerite is the main base metal sul-
The Sb content of the tetrahedrite analysed with EPMA/WDS is phide mineral, i.e. the Zn content is relative high compared to
relatively constant, but the Ag-content varies between 0.5 and the Cu content (MASS, Table 3). In Sb- and Ag-poor samples impu-
12.5 wt% (i.e., relative high standard deviation, cf. Table 2). An in- rity minerals were very rarely observed by optical microscopy. In
crease in Ag in tetrahedrite corresponds to a concomitant decrease the Ag-rich massive sulphide samples, tetrahedrite is the most
in Cu and vice versa, a well-known relationship between Cu and Ag common Sb-bearing mineral (Table 4). It is also the main Sb-bear-
in tetrahedrite (Moëlo et al., 2008). The highest Ag content (above ing mineral in sphalerite-dominated, relative Sb-rich samples from
5 wt%) is measured in tetrahedrite from mineralogical samples re- the uppermost part of the massive sulphide mineralisation (Ta-
ferred to as mafic rocks (for explanation see next section) whereas ble 4). In these samples, tetrahedrite is often penetrated by chalco-
the Ag content in tetrahedrite from massive sulphide samples tend pyrite veinlets which contain small inclusions of native antimony
to be lower. Arsenopyrite grains show fine-scale (1–5 lm), blotchy (mostly <10 lm, Fig. 3D). However, relatively Sb-rich massive sulp-

Table 2
Common Sb-bearing minerals at Rockliden and the average chemical composition in wt% recalculated from EMPA/WDS analysis (not shown are values for Bi, Se, Ni, Hg, and As).

LLDb Tetrahedrite Bournonite Meneghinite Boulangerite Gudmundite


Formulaa – (Cu,Fe,Ag,Zn)12.1Sb3.89S13 Pb1.01Cu1.03Sb0.94S3 Pb13.02Cu1.10Sb6.81S24 Pb5.16Sb3.80S11 Fe1.04Sb0.94S
No. of analysis – 38 25 20 8 39
AMc and SDd – AMc SDd AMc SDd AMc SDd AMc SDd AMc SDd
Ag (wt%) 0.05 5.84 3.17
Zn (wt%) 0.21 1.80 0.36 0.44 0.36 0.50 0.52
Cu (wt%) 0.16 33.56 2.23 13.34 0.25 1.56 0.29
S (wt%) 0.04 24.63 0.56 19.68 0.41 17.36 0.17 18.43 0.11 15.67 0.64
Fe (wt%) 0.10 5.85 0.48 0.46 0.31 0.47 0.26 0.39 0.28 28.21 0.62
Sb (wt%) 0.06 27.96 0.93 23.34 1.64 18.72 0.40 24.19 0.77 55.58 1.33
Pb (wt%) 0.19 42.67 1.63 60.86 0.69 55.84 0.46
a
Formula recalculated from EMPA/WDS analysis.
b
Lower limit of detection (EPMA/WDS analysis).
c
Arithmetic mean.
d
Standard deviation.

Table 3
Average chemical composition of different types (main and sub types, see text) of mineralogical samples.

Descriptiona Chemical subclassb No. of samples S (wt%) Cu (wt%) Zn (wt%) Sb (wt%) Ag (wt%) Pb (wt%) As (wt%)
DISS (felsic) Sp–Sb 1 2.21 0.04 0.60 1.84 59 0.25 0.12
DISS (mafic) Ccp–Ag(/Sb) 4 3.53 1.32 0.50 1.00 253 5.58 0.80
MASS (UP) Sp–Sb 2 38.35 0.30 7.47 0.13 62 0.62 0.10
MASS (DP) Sp–Sb 3 30.73 0.19 10.08 0.23 34 1.50 0.39
MASS Sp–Ag 2 26.50 1.09 25.99 0.25 236 5.23 0.34
MASS Ccp(-Ag) 3 36.10 3.97 3.83 0.12 96 0.51 1.23
a
DISS: disseminated to semi-massive sulphides (acid or mafic host rock, for explanation see text), MASS: massive sulphides (from upper (UP, i.e. drill core A, cf. Fig. 2) or
deeper parts (DP) of the Rockliden deposit).
b
Classes as defined by the preliminary classification (see Section 2 for explanation, Sp–Sb (or -Ag): sphalerite-dominated Sb-rich (or Ag-rich) subclass, Ccp–Ag: chalco-
pyrite-dominated Ag-rich subclass).
F. Minz et al. / Minerals Engineering 52 (2013) 95–103 99

Table 4 ca. 50 lm for individual grains up to 200 lm in star-shaped


Relative abundance of Sb-bearing minerals within mineralogical samples as esti- agglomerates (Fig. 3E and F). Gudmundite is also found as small
mated by optical microscopy and SEM/EDS spot analysis (xxx – major, xx – common,
grains (<10 lm) in complex intergrowths with chalcopyrite or
x – minor, (x) – rare, Ttr – tetrahedrite, Bour – bournonite, Gd – gudmundite, Boul –
boulangerite, Mene – meneghinite).
Fe-sulphides (Fig. 3E).
Most of the studied disseminated to semi-massive sulphide
Descriptiona Chemical subclassb No.d Ttr Bour Gd Boul Mene
samples are dominated by Fe–Mg-bearing silicate minerals such
DISS (felsic) Sp–Sb 1c x as amphiboles and biotite. They are referred to as mafic rocks. Four
DISS (mafic) Ccp–Ag(/Sb) 4 xx xxx x x xx out of five studied disseminated sulphide samples are of this type
MASS (UP) Sp–Sb 2 xxx x xx
MASS (LP) Sp–Sb 3 (x) xxx x xx
(Table 3). Chalcopyrite is dominating over sphalerite in the mafic
MASS Sp–Ag 2 xxx xx x rock samples, and bournonite is the main Sb-bearing phase in
MASS Ccp(-Ag) 3 xxx x xx these samples (DISS mafic, Table 4, Fig. 3A and B). Chalcopyrite is
a
DISS: disseminated to semi-massive sulphides (acid or mafic host rock, for
commonly found as inclusions or intergrowths with bournonite
explanation see text), MASS: massive sulphides (from upper (UP, i.e. drill core A, cf. (Fig. 3B). The bournonite crystals are up to 200 lm in length. Ex-
Fig. 2) or deeper parts (DP) of the Rockliden deposit). cept for bournonite, the abundance of Sb-bearing minerals in the
b
Classes as defined by the preliminary classification (see Section 2 for explana- mafic rock samples is rather variable, comprising tetrahedrite,
tion, Sp–Sb (or -Ag): sphalerite-dominated Sb-rich (or Ag-rich) subclass, Ccp–Ag:
meneghinite and gudmundite. Meneghinite reaches similar sizes
chalcopyrite-dominated Ag-rich subclass).
c
This sample contains mainly berthierite and minor jamesonite. as bournonite, whereas tetrahedrite has a diameter of ca. 50–
d
Number of mineralogical samples included. 100 lm and is commonly associated with bournonite, galena and
chalcopyrite (Fig. 3C). The remaining, non-mafic, studied dissemi-
nated sample (DISS felsic, Tables 3 and 4) consists of quartz and
hides from the deeper part of the Rockliden deposit (below ca. clay minerals and has sphalerite as the main base metal carrier.
300 m depth) have gudmundite as a common Sb-bearing phase This sample derives from a Sb-rich interval within the altered vol-
(Table 4). The grain size of gudmundite is variable and ranges from canic rocks and is referred to as felsic rock. Berthierite, jamesonite

Fig. 3. BSE images showing the mode of occurrence of Sb-bearing minerals in polished thin sections of mineralogical samples. The black background represents silicate
minerals (graphics A, B and C). (A) meneghinite (Mene) intergrown with galena (Gn) and bournonite (Bour); (B) bournonite (Bour) with inclusions of chalcopyrite (Ccp)
adjacent to sphalerite (Sp); (C) arsenopyrite (Apy) with a fine rim which is chemically similar to gudmundite () and tetrahedrite (Ttr) in contact with sphalerite (Sp), galena
(Gn) and chalcopyrite (Ccp); (D) tetrahedrite (Ttr) with fractures filled by chalcopyrite (Ccp) and fine native antimony (Ant); (E) gudmundite (Gd) together with galena (Gn) as
porous crystals and in fine complex intergrowths with pyrrhotite (Po) and chalcopyrite (Ccp) in contact with magnetite (Mgt) and pyrite (Py); and (F) gudmundite (Gd)
intergrown with quartz (Qtz) and calcite (Cal), meneghinite (Mene), native antimony (Ant), sphalerite (Sp), galena (Gn) and pyrrhotite (Po).
100 F. Minz et al. / Minerals Engineering 52 (2013) 95–103

Table 5
Chemical assays for flotation products (for location of the products in the flow sheet see Fig. 5).

CSa Product Ag (ppm) Cu (wt%) Zn (wt%) Pb (wt%) As (wt%) S (wt%) Sb (wt%) Sb distribution (%)
CA Cu–Pb RC 211 8.17 3.43 1.33 0.55 40.30 0.09 30.4
CAb Cu–Pb conc. 278 13.28 2.12 1.50 0.39 37.90 0.11 16.9
CA Zn RC 64 1.49 5.58 0.41 0.98 38.50 0.05 56.7
CAb Zn conc. 63 1.67 8.07 0.38 0.38 47.10 0.05 18.3
CAb Final tailing 32 0.63 1.28 0.25 1.13 16.25 0.03 12.9
CB Cu–Pb RC 246 4.98 6.68 2.87 0.23 39.33 0.43 65.2
CBb Cu–Pb conc. 321 7.11 4.86 3.62 0.15 39.20 0.57 51.3
CB Zn RC 57 0.73 8.36 0.64 0.59 41.70 0.08 28.9
CBb Zn conc. 58 0.75 13.17 0.68 0.35 46.90 0.08 15.1
CBb Final tailing 14 0.14 0.36 0.22 0.13 2.87 0.03 5.9
CC Cu–Pb RC 264 18.91 4.30 3.78 0.35 28.80 0.48 48.9
CCb Cu–Pb conc. 291 25.35 2.38 4.09 0.20 29.50 0.54 36.0
CC Zn RC 33 0.82 7.48 0.26 0.55 38.40 0.06 35.8
CCb Zn conc. 34 1.02 17.51 0.32 0.18 45.90 0.05 12.0
CCb Final tailing 10 0.10 0.63 0.14 0.62 9.27 0.05 15.3
a
Composite sample (for location in the Rockliden deposit see Fig. 2, for location of the products in the flow sheet see Fig. 5).
b
Products sampled for microscopic studies (cf. Fig. 2).

Table 6 occurrence of Sb-bearing minerals in particles of the flotation


Relative abundance of minerals in Sb-bearing sulphide fraction within flotation products was studied more closely in BSE images.
samples based on values obtained with SEM-based quantification (xxx – major, xx – Beside liberated mineral grains, tetrahedrite (Fig. 4A and B) and
common, x – minor, (x) – rare, y – particles were documented by EDS analysis on
single grains; Ttr – tetrahedrite, Bour – bournonite, Gd – gudmundite, Boul –
bournonite occur together with galena in binary particles from the
boulangerite, Mene – meneghinite, Jms - Jamesonite). Cu–Pb concentrates, but also with sphalerite, chalcopyrite and pyr-
ite. Gudmundite is a minor Sb-bearing mineral in Cu–Pb concen-
CSa Product Ttr Bour Boul Mene Gd Jms
trate (Table 6), but it occurs locked together with galena in
CA Cu–Pb conc. xxx x y (x) y binary particles and also in ternary particles complexly intergrown
CA Zn conc. xxx x xxx
with chalcopyrite (Fig. 4B and C). In the Zn concentrates, tetrahe-
CA Tailing y xxx
CB Cu–Pb conc. xx xxx x y x drite is found locked with sphalerite, galena and chalcopyrite
CB Zn conc. x xxx xx (Fig. 4E), as well as with arsenopyrite and bournonite. In the Zn
CB Tailing xxx xxx concentrate bournonite (Fig. 4D) is locked with galena, pyrite, pyr-
CC Cu–Pb conc. x xxx xx y
rhotite, chalcopyrite, sphalerite, boulangerite and arsenopyrite. In
CC Zn conc. y xx xxx y x
CC Tailing x y xxx
the tailing products, gudmundite is mostly found as liberated
grains. But gudmundite as well as meneghinite occur also locked
a
Composite sample (for location see Fig. 2). mainly with gangue minerals like chlorite, quartz, calcite and pla-
gioclase (Fig. 4F).

and gudmundite comprise the Sb-bearing minerals documented in


this disseminated sulphide sample.
4. Discussion

Microscopic studies of mineralogical samples are summarised


3.3. Characterisation of Sb-bearing minerals in the flotation test and compared with the Sb deportment in the flotation products
products in Table 7. Similarities in the associations between the mineralog-
ical samples and flotation products can be seen in this table, i.e. the
The distribution of Sb-bearing minerals in the final flotation mode of occurrence of Sb-bearing minerals as found in the uncru-
products of the three Rockliden composite samples is summarised shed material apparently is reflected in particles in the flotation
in Fig. 5. Chemical analysis for the flotation products (as outlined in products. Thus reasons for the behaviour of particles during flota-
Fig. 5) are shown in Table 5 and the distribution of Sb-bearing min- tion can be linked to the mineral grain sizes and associations found
erals is further disclosed in Table 6. in the primary rock and mineralised material.
In the Cu–Pb concentrates, tetrahedrite and bournonite are the Based on observations done in a limited number of thin sections
major Sb-bearing phases (Table 6). The Cu–Pb concentrates of the on massive sulphide samples, there is an apparent change in the Sb
composite samples from the upper part of the Rockliden deposit mineralogy of the massive sulphide samples towards greater depth
(CA and CB (cf. Fig. 2), Table 6) have a higher abundance of tetra- (>300 m below surface at the Rockliden deposit). The Sb-rich mas-
hedrite compared to the Cu–Pb concentrate from the tests done sive sulphide samples from the upper part of the deposit contain
for composite sample CC (cf. Fig. 2) taken at greater depth in the abundant tetrahedrite often with complex intergrowths of pyrrho-
deposit (ca. 300–900 m). The Sb mineralogy of the Zn concentrates tite-chalcopyrite-native antimony, whereas gudmundite is the
is similarly complex to the Cu–Pb concentrate, comprising bourno- main Sb-bearing mineral the Sb-rich massive sulphide samples
nite, gudmundite, tetrahedrite and boulangerite (Table 6). In the from greater depth. In the disseminated sulphide samples, bourno-
tailing samples from the Rockliden flotation tests, gudmundite is nite is the common Sb-bearing phase of the mafic rocks. Mafic
the dominant Sb-bearing mineral (Table 6). rocks are documented within or close to the massive sulphides
In the laboratory flotation tests the P80 of the flotation feed was throughout the 3D extent of the deposit, potentially complicating
ca. 50 lm. Although most of the sulphide minerals are liberated at the Sb-mineralogy of composite samples and flotation products
this size, the Sb-bearing minerals remain partly locked with other irrespectively of the depth from which these products are derived.
minerals. This is not unexpected and reflects the size of Sb-bearing Moreover, one sample taken from the felsic rocks (i.e. volcanic
minerals observed in the mineralogical samples. Thus the mode of host rock of Rockliden) contains additional Sb-bearing phases,
F. Minz et al. / Minerals Engineering 52 (2013) 95–103 101

Fig. 4. BSE images of Sb-bearing minerals in the flotation products. (A) binary particle with tetrahedrite (Ttr) and chalcopyrite (Ccp); (B) gudmundite (Gd) together with
chalcopyrite (Ccp) and Sb-dominated sulphide mixture (white dots, <5 lm); (C) gudmundite (Gd) in complex intergrowth with chalcopyrite (Ccp) and galena (Gn); (D)
ternary particle with sphalerite (Sp), galena (Gn) and bournonite (Bour); (E) ternary particle with sphalerite (Sp), galena (Gn) and tetrahedrite (Ttr); and (F) meneghinite
(Mene) intergrown with plagioclase (Plg).

Table 7
Comparison of the mode of occurrence of common Sb-bearing minerals in the mineralogical samples and flotation products.

Mineral and ideal chemical Position in the ore body Associationa in the Distribution in the Associationa in particles
formula mineralisation flotation productsb
Tetrahedrite (Ttr), Mafic rocks, MASS Gn, Bour, Ccp, Sp, Po Cu–Pb conc., Zn conc. Gn, Sp, Ccp, Po, silicate minerals
(Cu,Fe,Ag,Zn)12Sb4S13 (not Sp–Sb in DP)
Bournonite (Bour), PbCuSbS3 Mafic rocks, MASS (only traces in Gn, Ccp, Ttr, Sp Cu–Pb conc., Zn conc. Gn, (Sp, Ccp, Boul, Py)
Sp–Sb in DP)
Meneghinite (Mene), Mafic rocks, MASS Sp–Sb in DP Apy, Gn, silicate minerals Cu–Pb conc., Zn conc., Gn, silicate minerals
Pb13CuSb7S24 (tailing)
Boulangerite (Boul), Pb5Sb4S11 Mafic rocks, MASS Sp–Sb in DP Apy, Gn, silicate minerals Cu–Pb conc., Zn conc., Bour
(tailing)
Gudmundite (Gd), FeSbS All Po, Sp, Gn, Ttr, Apy, (as small Zn conc., (Cu–Pb conc.) Gn, (Ccp, Ttr), silicate &
grains < 10 lm with Ccp, Po, Sp) carbonate minerals
silicate & carbonate minerals
a
Apy – arsenopyrite, Ccp – chalcopyrite, Gn – galena, Po – pyrrhotite, Py – pyrite, Sp – sphalerite.
b
Underlined are products where the Sb-bearing minerals are unexpected (cf. Lager and Forssberg, 1990).

jamesonite and gudmundite, indicating that the Sb-mineralogy of the Rockliden deposit that the Ag-content can be as high as
the host rocks is locally more complex. 15 wt% in tetrahedrite in the north-western parts of the Rockliden
Further, the highest Ag-values of tetrahedrite were documented deposit (Karup-Møller and Makovicky 1980; Mattsson and
in the mineralogical samples of mafic rock, indicating that the var- Heeroma, 1985) and variations might be related to even larger
iation of Ag in tetrahedrite is partly related to different rock and scale geological features. These relationships need to be under-
mineralisation types. Also, it is known from previous studies of stood in order to be able to predict the distribution of Sb combined
102 F. Minz et al. / Minerals Engineering 52 (2013) 95–103

Fig. 5. Flow sheet of the laboratory flotation test (chemical assays are given in Table 5 for the streams with bold arrows). Tables for the final products show the abundance of
tetrahedrite (Ttr), bournonite (Bour), boulangerite (Boul, including meneghinite (Mene)) and gudmundite (Gd) within the Sb-bearing mineral fraction (estimations are based
on quantifications done with INCA Mineral software (Oxford Instruments NanoAnalysis, 2012)). Sb-bearing minerals which are highly expected to be found in the flotation
products (cf. Lager and Forssberg, 1990) are bold and underlined in the figure.

with the economic valuable Ag in tetrahedrite and in other sul- centrate, before sending this product to the smelter (Awe and
phosalts throughout the entire deposit. Sandström, 2010; Bolin, 2010). Also, the Sb-recovery was above
The examples given above suggest that the Sb-mineralogy in 90 wt% in alkaline sulphide leaching tests from Cu–Pb and Zn con-
the Rockliden deposit can be linked to the geology, to rock and centrates which contain additionally bournonite and gudmundite
mineralisation types. Thus it is clear that a geological model, as common Sb-bearing phases (Sandström, personal communica-
accounting for the complex geometry of the Rockliden massive sul- tion 2013). Thus, process options are available when these miner-
phide body should form the basis of a detailed 3D mineralogical als report to the Cu–Pb concentrate.
map of the Sb mineralogy.
The Sb mineralogy of flotation products from Rockliden (Fig. 5)
is partly consistent with the expected distribution based on the 5. Conclusions
study of Lager and Forssberg (1990) on the separation of Sb miner-
als in flotation tests of Rakkejaur complex sulphide ore. The Sb- Since the Sb-content controls the quality of the Cu–Pb concen-
bearing minerals containing Cu, Ag or Pb are mainly found in the trates, it is important to know the distribution of Sb-bearing min-
Cu–Pb concentrate and those which combine Sb with Fe or As in erals and their mode of occurrence in the ore. This study outlines
the tailings. variations in the distribution of Sb-bearing minerals, which appear
Based on studies by Lager and Forssberg (1990), it is expected to be related to rock and mineralisation types found in the Rockl-
that liberated tetrahedrite, bournonite, boulangerite, meneghinite iden Zn–Cu massive sulphide deposit. During flotation, both the
and jamesonite float readily under the conditions of the Rockliden chemical composition and mineral association of the Sb-bearing
laboratory flotation tests. These minerals should largely report to minerals, inherited from the primary ore material, control the dis-
the Cu–Pb concentrate. The remaining Sb phase, gudmundite, is tribution of these minerals. Examples are illustrated in this study.
not expected to float in Cu–Pb and Zn flotation and therefore it In order to implement the data on the Sb mineralogy into a geo-
should be found in the tailings. However, the mineralogical com- metallurgical model of the Rockliden deposit, the results of this
position of the flotation products is generally more complex than study need to be quantified. The value of qualitative and quantita-
expected from the liberation distribution of the Sb-bearing miner- tive mineralogical and petrographic studies has already been
als (Fig. 5). When locked the associating minerals (cf. Table 7, pointed out in case studies on process optimisation for several
Figs. 3 and 4) will influence the behaviour of complex particle ore deposits (e.g. Baum et al., 2004; Lotter et al., 2011). Thus, a
depending on relative exposure proportions. Chalcopyrite- and main target for future work on Rockliden is to quantify the distri-
galena-rich particles are expected to be recovered into the Cu– bution and the mode of occurrence of Sb-bearing minerals in the
Pb concentrates and the locked Sb-bearing minerals will therefore ore and flotation products using automated SEM-based methods.
follow (e.g. Fig. 4B and C). Similarly sphalerite-rich particles with This will also show how rigorously the Sb mineralogy can be de-
Sb-bearing minerals most likely will be recovered into Zn concen- rived, e.g. from chemical assays, or whether other methods such
trate. Observations support these generalisations. For example the as diagnostic leaching must be applied to quantify the Sb mineral-
Sb-bearing minerals found in Zn concentrates occur locked in ogy at Rockliden. Quantitative data should also be used in order to
sphalerite-rich particles (e.g. Fig. 4D and E). In tailing products test proposed relationships between the Sb mineralogy and rock,
the Sb-bearing minerals, other than gudmundite, were found alteration, and mineralisation types.
locked with gangue minerals (e.g. meneghinite with plagioclase, If the influence of all factors governing the variations in the Sb
Fig. 4F). mineralogy, as pointed out in this study, is understood in detail,
Small grain size and complex mineral textures show that the quantified and used in a 3D model, it is expected that the Sb char-
liberation of Sb-bearing minerals in Rockliden is challenging and acteristics of flotation products can be predicted for different parts,
would require fine grinding. Furthermore their rejection from i.e. geometallurgical units, at the Rockliden massive sulphide de-
Cu–Pb concentrate would require special flotation conditions since posit. Textural information, as outlined by the mode of occurrence
most of the Sb minerals are Cu- or Pb-bearing and therefore having of Sb-bearing minerals in this study, plays a significant role in cre-
natural tendency to float in the chemical regime used (Benzaazoua ating the model. The task of transferring texture information from
et al., 2002; Byrne et al., 1995). It is, however, possible to remove the ore to quantitative mineral association data of corresponding
tetrahedrite via alkaline leaching processes from the Cu–Pb con- flotation feed is challenging but concepts have been proposed
F. Minz et al. / Minerals Engineering 52 (2013) 95–103 103

(e.g. Lamberg, 2011). The prediction needs to have good accuracy Butterman, W.C., Carlin, J.F., 2004. Mineral commodity profiles, antimony. Open-File
Report 03-019US. Department of the Interior, US Geological Survey.
to support production management, for example to make decisions
Byrne, M., Grano, S., Ralston, J., Franco, A., 1995. Process development for the
on the treatment of the flotation products, i.e. decisions on when separation of tetrahedrite from chalcopyrite in the Neves-Corvo ore of Somincor
the Cu–Pb concentrate has to be treated by unconventional process SA. Portugal Miner. Eng. 8 (12), 1571–1581.
solutions such as alkaline leaching. Depauw, G., 2009. Geology of the Rockliden volcanogenic massive sulphide
deposit, north central Sweden. Master Thesis, Luleå University of Technology,
pp. 1–66.
Acknowledgements Dunham, S., Vann, J., Coward, S., 2011. Beyond Geometallurgy – gaining
competitive advantage by exploiting the broad view of geometallurgy.
Paper presented at The First AusIMM International Geometallurgy
New Boliden is acknowledged for financing the analytical work Conference 2011, pp. 115–123.
in this study, and for permission to publish this paper. European Commission, 2010. Critical raw materials for the EU – Report of the Ad-
H. Raat and J. Lasskogen (Boliden) are thanked for introducing Hoc Working Group on Defining Critical Raw Materials. Enterprise and Industry,
Brussels, pp. 1–85.
us to the geology of Rockliden and to the geological interpretation Evins, P., 2011. New map and structural analysis of the Rockliden exposure. Int. Rep.
of drill cores. The help and assistance with SEM/EDS and EPMA/ Boliden Mines 46346000, 1–23.
WDS by following persons is acknowledged: S. Varpenius, L. Palmu, Filippou, D., St-Germain, P., Grammatikopoulos, T., 2007. Recovery of metal values
from copper–arsenic minerals and other related resources. Min. Process. Extr.
K. Aasly, K. Eriksen, G. Bark, I. Bhuiyan and J. Mouzon. Metall. Rev. 28 (4), 247–298.
This study is part of a PhD project financed by CAMM (Centre of Jackson, J., McFarlane, E.J., Olson Hoal, K., 2011. Geometallurgy - back to the future:
Advanced Mining and Metallurgy). ALS Minerals Division (Piteå, Scoping and communicating geomet programs. Paper presented at The First
AusIMM International Geometallurgy Conference 2011, pp. 125–131.
Sweden) is thanked for financial support for preparation of miner-
Karup-Møller, S., Makovicky, E., 1980. A mineralogical investigation of some
alogical samples for chemical assays and analytical work. Travel- mineral concentrates, feed samples and trench samples from central Sweden.
ling costs for SEM/EDS studies were partly covered by Int. Rep. Boliden Mines 35 (8), 1–44.
ProMinNET (Nordic network of Process Mineralogy and Geometal- Kousa, J., Lundqvist, T., 2000. Svecofennian domain. In Description of the bedrock
map of central Fennoscandia (mid- norden). In: Lundqvist, T., Autio, S (Eds.),
lurgy 380 funded by Nordforsk). Geological Survey of Finland, Espoo, Special Paper 28, pp. 47–74.
The authors also acknowledge comments of two anonymous Lager, T., 1985. The Rakkejaur deposit, a challenge to flotation process technology.
reviewers. In: Forssberg, K.S.E. (Ed.), Flotation of Sulphide Minerals. Elsevier, Amsterdam,
pp. 255–275.
Lager, T., 1989. The technology of processing antimony-bearing ores. PhD Thesis,
References Luleå Tekniska Universitet.
Lager, T., Forssberg, K.S.E., 1990. Separation of antimony mineral impurities from
Allen, R.L., Weihed, P., Svenson, S.A., 1996. Setting of Zn–Cu–Au–Ag massive sulfide complex sulphide ores. Trans. Instn. Min. Metall (Sect. C: Mineral Process. Extr.
deposits in the evolution and facies architecture of a 1.9 ga marine volcanic arc, Metall.) 99 (C54), 54–61.
Skellefte district, Sweden. Econ. Geol. 91 (6), 1022–1053. Lamberg, P., 2011. Particles – the bridge between geology and metallurgy. Preprints
Anderson, C.G., 2012. The metallurgy of antimony. Chem. Erde. 72, 3–8. Conference in Minerals Engineering 2011, 978–91-7439-220-3, pp. 1–16.
Anderson, C.G., Nordwick, S.M., Krys, L.E., 1991. Processing of antimony at the Lotter, N.O., Kormos, L.J., Oliveira, J., Fragomeni, D., Whiteman, E., 2011. Modern
Sunshine Mine. In: Reddy, R.G., Imrie, W.P., Queneau, P.B. (Eds.), In Residues and process mineralogy: two case studies. Miner. Eng. 24 (7), 638–650.
effluents: Processing and Environmental Considerations. The Minerals, Metals Masters, H. Antimony. Financial Times Energy, 2005. <http://www.mmta.co.uk/
and Materials Society, pp. 349–366. uploads/2012/09/19/153733_antimony.pdf> (cited 01.29.2012).
Awe, S.A., 2010. Hydrometallurgical upgrading of a tetrahedrite-rich copper Mattsson, B., Heeroma, P., 1985. Rockliden prospekteringsresultat 1981–1985.
concentrate. Licentiate thesis, Luleå University of Technology. <http:// Internal report Boliden Mines, GP 85 005 2, pp. 1–35.
pure.ltu.se/portal/files/5127574/Samuel_Ayowole_Awe_Lic2010.pdf>. Moëlo, Y., Makovicky, E., Mozgova, N.N., Jambor, J.L., Cook, N., Pring, A., Paar, W.,
Awe, S.A., Sandström, Å., 2010. Selective leaching of arsenic and antimony from a Nickel, E.H., Graeser, S., Karup-Møller, S., 2008. Sulfosalt systematics: a review.
tetrahedrite rich complex sulphide concentrate using alkaline sulphide Report of the sulfosalt sub-committee of the IMA commission on ore
solution. Miner. Eng. 23 (15), 1227–1236. mineralogy. Eur. J. Mineral. 20 (1), 7–46.
Awe, S.A., Sandström, Å., 2012. Antimony electrowinning from alkaline sulphide New Boliden Annual Report 2012, 2013. New Boliden, Stockholm, 132 p.
electrolytes. Internal Report Boliden Mines. Luleå University of Technology, 21 Oxford Instruments NanoAnalysis, 2012. INCAFeature – instructions manual. High
p. Wycombe (UK), 1–124 p.
Awe, S.A., Khoshkhoo, M., Krüger, P., Sandström, Å., 2012. Modelling and process Raat, H., Årebäck, H., 2009. Geology, grade and tonnage estimate of the Rockliden
optimisation of antimony removal from a complex copper concentrate. Trans. VMS deposit, north central Sweden. Int. Rep. Boliden Mines 34, 1–56.
Nonferrous Met. Soc. China 22 (3), 675–685. Roine, A. 2009. HSC chemistry 7.0 user’s guide – chemical reaction and equilibrium
Baum, W., Lotter, N.O., Whittaker, P.J., 2004. Process mineralogy – a new generation software with extensive thermochemical database and flowsheet, simulation,
for ore characterisation and plant optimization. Paper presented at SME Annual vol. 1/2.
Meeting, Preprint 04-12, pp. 1–5. Tian-Cong, Z., 1988. The Metallurgy of Antimony. The Peoples Republic of China:
Benzaazoua, M., Marion, P., Liouville-Bourgeois, L., Joussemet, R., Houot, R., Franco, Central South University of Technology Press, Changsa.
A., Pinto, A., 2002. Mineralogical distribution of some minor and trace elements Walters, S., Kojovic, T., 2006. Geometallurgical mapping and mine modeling (GEM3)
during a laboratory flotation processing of Neves-Corvo ore (Portugal). Int. J. – the way of the future. Paper presented at International Autogenous and Semi-
Miner. Process 66 (2002), 163–181. Autogenous Grinding Technology 2006, Vancouver, Canada, pp. IV-411–425.
Björnberg, A., Holmstrom, S.A., Lindkvist, G., 1986. Method for processing copper Wavra, C.S., Bond, W.D., Reid, R.R., 1994. Evidence from the sunshine mine for dip-
smelting materials and the like containing high percentages of arsenic and/or slip movement during Coeur D’Alene District mineralization. Econ. Geol. 89 (3),
antimony. US Patent, Patent 4,626,279, December 2, 1986. 515–527.
Bolin, N.-J., 2010. Rockliden – mineralogy investigation of products produced from Williams, S., 2011. Why do geometallurgical studies? no surprises! SEG Newsl. Apr.
three different ore samples. Internal Report Boliden Mines, TM_REP2010/043, 2011 (85), 25.
pp. 1–7.
PAPER II

Lithology and mineralisation types of the Rockliden Zn–Cu massive sulphide deposit,
north-central Sweden: Implications for ore processing

Friederike E. Minz, Jonas Lasskogen, Christina Wanhainen, Pertti Lamberg

Applied Earth Science (Trans. Inst. Min. Metall.): Section B (2014), Vol. 123 (1), pp. 2–17
Lithology and mineralisation types of the Rockliden Zn–Cu
massive sulphide deposit, north-central Sweden:
Implications for ore processing

F. E. Minz*1, J. Lasskogen2, C. Wanhainen1 and P. Lamberg1

Department of Civil, Environmental and Natural Resources Engineering, Luleå University of


1

Technology, Luleå, Sweden


2
Exploration Department, Boliden Mines, Boliden, Sweden
*Corresponding author, email friederike.minz@ltu.se

Abstract

The Rockliden Zn–Cu volcanic-hosted massive sulphide deposit is located approximately 150 km
south of the Skellefte ore district, north-central Sweden. Most of the mineralisation is found at the
altered stratigraphic top of the felsic volcanic rocks, which are intercalated in the metamorphosed
VLOLFLFODVWLFVHGLPHQWDU\URFNVRIWKH%RWKQLDQ%DVLQ0D¿FG\NHVFURVVFXWDOOOLWKRORJLFDOXQLWV
including the massive sulphides, at the Rockliden deposit. The relatively high Sb grade of some
parts of the mineralisation results in challenges in handling of the Cu–Pb concentrate in the smelting
process. The aim of this study is to characterise different host rock units and ore types by their main
mineralogy, as well as by their trace mineralogy with focus on the Sb-bearing minerals. Ore types
are distinguished largely on the basis of their main base-metal bearing sulphide minerals, which are
chalcopyrite and sphalerite. Several Sb-bearing minerals are documented and differences in the trace
mineralogy between rock and ore types are highlighted. Based on the qualitative ore characterisation,
rock- and ore-intrinsic parameters, such as the pyrite, pyrrhotite and magnetite content of the massive
sulphides, the trace mineralogy and its association with base-metal sulphide minerals, are outlined
and discussed in terms of relevance to the ore processing.

Key words: Rockliden; North-central Sweden; Base metal sulphide; Ore characterisation;
Process mineralogy; Impurity elements

1
1. Introduction were performed with new samples (Raat and
Årebäck, 2009; Bolin, 2010). There is now an
The Rockliden volcanic-hosted massive sulphide alternative treatment considered to remove Sb
(VHMS) deposit is located in the Västernorrlands from the Cu–Pb concentrate by alkaline sulphide
Län, north-central Sweden, ca. 150 km SW of leaching (Awe, 2013; Bolin, 2010). However, in
the Skellefte ore district and the nearby existing order to determine when it is appropriate to use
processing facilities (e.g. the smelter at Rönnskär, this optional treatment, it is important to know
Fig. 1). The inferred mineral resource is 3.53 Mt what parts of the ore are likely to generate high Sb
with 4.2 wt-% Zn, 1.9 wt-% Cu, 0.7 wt-% Pb and contents in the Cu–Pb concentrate (Minz, 2013).
71 g/t Ag (New Boliden, 2013). The Rockliden Studies of the Sb mineralogy of Rockliden have
deposit is relatively Cu-rich compared to other pointed towards several parameters controlling
Swedish VHMS deposits, e.g. Kristineberg, the distribution of Sb-bearing minerals during
Renström, Maurliden and Rakkejaur (Allen et ÀRWDWLRQDQGKHQFHWKH6EFRQWHQWRIWKH&X±3E
al., 1996; New Boliden, 2013). Similar to other concentrate (Minz et al., 2013). One potential
VHMS deposits of the Skellefte ore district, process-relevant parameter is the interlocking
the Rockliden mineralisation shows high of penalty element-bearing minerals with base-
concentrations of penalty elements such as As, metal and Fe sulphides in mineralogically
Sb and Hg (Allen et al., 1996; Raat and Årebäck, complex particles (Minz et al., 2013).
2009).
Details of the Sb mineralogy of Rockliden are
Shortly after the discovery of the Rockliden published by Minz et al. (2013). A detailed
deposit in the 1980s, the project was put on hold petrographic and mineralogical description
when metallurgical tests indicated a high Sb of the different rock and mineralisation types
content in the Cu–Pb concentrate, thus lowering at the Rockliden exploration project, i.e. the
the quality of this product (Raat and Årebäck, framework to the Sb-bearing minerals, is
2009). In 2007, exploration drilling in the presented in this paper. The main purpose of
Rockliden area restarted, indicating a decrease this study is to characterise and classify the
in the average Sb content at depth and some Rockliden mineralisation qualitatively based
YDULDWLRQ DORQJ VWULNH DQG QHZ ÀRWDWLRQ WHVWV on the occurrence and distribution of the

a b
70ºN
Piteå

N
Boliden 7200000
Lu

Rönnskär /
le
å- lin
Jo e

Skellefteå
kk

Lycksele
m
ok

Skellefte
k

District
Boliden Dorotea
64ºN Rönnskär
Rockliden Åsele
7100000
Rock-
Bas n
nia

Umeå
in

liden
Both

NORWAY FINLAND Solberg


Junsele

Bergslagen
District

7000000

58ºN
SWEDEN
100 km
Sundsvall
1500000 1600000 1700000 1800000
12ºE 20ºE 28ºE

Fig. 1. a) Location of the Rockliden VHMS deposit in the Bothnian Basin limited by the Bergslagen and Skellefte ore
districts (Kumpulainen, 2009), south of the Luleå-Jokkmokk line (Mellqvist et al., 1999; Weihed, 2004), b) road map
locating the exploration site (coordinate system: Swedish grid RT90 2.5 gon west).

2
major sulphide minerals (i.e. pyrite, sphalerite, (2009) proposed an active continental margin
chalcopyrite and pyrrhotite. The parameters VHWWLQJ EDVHG RQ WKH JHRFKHPLFDO FODVVL¿FDWLRQ
used in this petrographic study to characterise of Schandl and Gorton (2002), or an extensional
the massive sulphides include not only the setting within the Bothnian Basin rimmed to the
mineralogical composition but also textural south by the Bergslagen volcanic arc and to the
information such as grain size, oriented north by the Skellefte volcanic arc.
intergrowths and association with penalty and
bonus element-bearing minerals (cf. Walters and Based on geochronological age data the following
Kojovic, 2006). Herein, the current knowledge post-depositional history was proposed. Peak
of the local geology and mineralogy of metamorphism was reached in the Bothnian Basin
Rockliden is compiled, and is discussed in terms during the Svecokarelian orogeny (1.9 to 1.8 Ga)
of petrographic parameters that are expected to and large parts of the basin were metamorphosed
LQÀXHQFH WKH RUH SURFHVVLQJ 7KLV TXDOLWDWLYH to upper greenschist to upper amphibolite facies,
evaluation gives guidance for quantifying and partly migmatised (Lundqvist et al., 1990;
potential process-relevant parameters which Weihed, 2004). Late-orogenic S-type granitoids
might form the background of a process-adopted (1.82 to 1.80 Ga) and post-orogenic A- to I-type
geological model for the Rockliden deposit Revsund granitoids (1.80 to 1.77 Ga) intruded
(Lamberg, 2011; Minz, 2013). into the basin (Claesson and Lundqvist, 1995;
Weihed, 2004; Fig. 2d). Between 1.27 to 1.24 Ga
2. Geological setting WKHEDVLQZDVLQWUXGHGE\PD¿FVLOOVDQGG\NHV
of the Central Scandinavian Dolerite Group
2.1. Regional geology (CSGD) (Söderlund et al., 2006).

The Rockliden massive sulphide deposit is 2.2. Local geology


located within the Bothnian Basin (Welin, 1987;
Lundqvist et al., 1998; Kousa and Lundqvist, The Rockliden massive sulphide deposit is
2000), in the Svecofennian rocks south of located primarily at the contact between the
the Luleå-Jokkmokk line (Weihed, 2004). Härnö sedimentary rocks and the Rockliden
The basin is located south of the Skellefte ore volcanic rocks (Depauw, 2009). On the basis of
district and north of the Bergslagen ore district the geochronological age data of Welin (1987),
(Kumpulainen, 2009; Fig. 1). The supracrustal the Rockliden felsic volcanic rocks are located
rocks of the Svecofennian Domain are dominated in the stratigraphic upper part of the sedimentary
by turbiditic metasedimentary rocks of the Härnö URFNV 'HSDXZ   7KH\ DUH FODVVL¿HG
group which is estimated to be at least 10 km as calc-alkaline rhyolites to rhyodacites and
thick (Lundqvist, 1987; Lundqvist et al., 1990; dacites (Mattsson and Heeroma, 1985) and are
Kousa and Lundqvist, 2000; Kumpulainen, interpreted as quartz–feldspar phyric coherent
2009). Volcanic rocks in the Bothnian Basin are GRPHV WXI¿WLF URFNV ODYDV DQG K\DORFODVWLWHV
divided into basalt-andesites and fractionated DQG YROFDQRFODVWLF PDVV ÀRZV DXWRFODVWLF
basalts to rhyolites (Bergström, 2001). In the breccias, and agglomerates. The stratigraphic
Rockliden area, Mattsson and Heeroma (1985) sequence of these rocks is not well understood.
GHVFULEHGPD¿FURFNW\SHVFRPSULVLQJJDEEURV Metamorphosed turbiditic shales and siltstones
and basaltic rocks, which are found in outcrops stratigraphically overlie the mineralisation
within a distance of 5 to 10 km from the and generally show a sharp contact against the
Rockliden deposit. Acid volcanic rocks occur massive sulphides (Mattsson and Heeroma,
locally. Metarhyolitic to metadacitic volcanic 1985; Raat and Årebäck, 2009; Depauw, 2009).
rocks, located just 1.5 km west of Rockliden, To the north and west, the Rockliden area is
were dated 1875 Ma (Welin, 1987). As with bordered by granitic rocks, whereas to the south
the volcanic rocks of the Skellefte group (Allen and east metagreywackes of the Härnö group are
et al., 1996; Rutland et al., 2001), a number of found as gneisses and migmatites (Mattsson and
different tectonic settings for the Rockliden +HHURPD'HSDXZ 9DULRXVPD¿F
volcanic rocks have been suggested. Depauw intrusions occur in the Rockliden area including

3
decimetre- to metre-wide sills and dykes. They do not appear to be continuous over vertical
mostly occur along structural zones of weakness, distances larger than 50 m and often appear to
such as faults (Raat and Årebäck, 2009). Dolerite coincide with possible fault structures (Raat and
dykes with chilled margins crosscut all rocks of Årebäck, 2009).
the Rockliden area in a WSW–ENE to SW–NE
direction (Fig. 2; Mattsson and Heeroma, 1985; While the metamorphic grade in large parts of
5DDWDQGcUHElFN 0D¿FVLOOVDQGG\NHV the Bothnian Basin reached upper amphibolite
lacking distinct chilled margins, are found close facies conditions during the Svecokarelian
to, or within, the mineralisation (Depauw, 2009). orogeny (e.g. Lundqvist et al., 1998), the host
$W GHSRVLWVFDOH WKHVH PD¿F G\NHV DUH VWHHSO\ rocks to the Rockliden mineralisation have
dipping, mostly lying parallel to bedding or reached only a maximum metamorphic grade
along the foliation of the host rocks, but they of upper greenschist facies (Depauw, 2009).

a b A

N N

7 074 000 mN

7 073 000 mN

7 072 000 mN

7 071 000 mN

100 m

c
B ca. 400 m.a.s.l. A
7 070 000 mN
1 575 000 mE

1 576 000 mE

1 577 000 mE

1 578 000 mE

1 580 000 mE
1 579 000 mE

d
stratigraphic
column: drill core c, Fig. 7

CSDG (Söderlund et al., 2006)


Revsund post orogenic* (Claesson and Lundqvist, 1995)

Härnö late orogenic* (Claesson and Lundqvist, 1995)


early orogenic* (Kousa and Lundqvist, 2000)

*) Svecokarelian orogeny
100 m
legend:
granitic intrusions
felsic volcanic rocks (FVR)
sericite altered FVR
Härnö Group (e.g. Kousa chlorite altered FVR
and Lundqvist, 2000)
massive sulphides
meta-sedimentary rocks
mafic dykes and drill core a (ca. 150 to west)
intrusions, unspecified
dolerite dykes
faults drill core b, Fig. 7

Fig. 2. Geological maps and section of the Rockliden area and mineralisation: a) surface map of the geology of the
5RFNOLGHQDUHD EDVHGRQJHRFKHPLFDOFODVVL¿FDWLRQRISHUFXVVLRQGULOOFRUHVDPSOHVFRRUGLQDWHV\VWHP6ZHGLVKJULG
RT90 2.5 gon west), b) detailed surface map showing the extent and shape of the massive sulphides (enlarged map
from picture a)), c) cross section (see graphic b for location) showing the vertical distribution of alteration zones in the
IHOVLFYROFDQLFKRVWURFNDQGPLQHUDOLVDWLRQ IURP'HSDXZ  GULOOFRUHSUR¿OHVDUHIRXQGLQ¿JXUHd) strati-
JUDSKLFFROXPQ PRGL¿HGIURP'HSDXZ  $EEUHYLDWLRQ&6'*±&HQWUDO6FDQGLQDYLDQ'ROHULWH*URXS

4
According to Raat and Årebäck (2009) and This includes structural repetition of massive
Depauw (2009), the volcanic rocks show, sulphide bodies and remobilisation of ore shoots
locally, impregnations of carbonate minerals and along fault planes, as well as enrichment of
VLOLFL¿FDWLRQLQLUUHJXODU]RQHVLQGLVWDOSDUWVRI galena and chalcopyrite in fold hinges and along
the massive sulphide mineralisation. Sericitic faults and transformation of pyrite to pyrrhotite
alteration is the most prominent type in the (Depauw, 2009; Raat and Årebäck, 2009).
Rockliden volcanic footwall rocks, extending
several tens to hundreds of metres into these 3. Methods
rocks and transgressing into a strong to intense
sericite-quartz dominated alteration within In preparation of drill core sampling for this ore
tens of metres to a few metres of the massive FKDUDFWHULVDWLRQ VWXG\ D VLPSOL¿HG PLQHUDORJ\
sulphides. A zone of chlorite alteration of several (pyrite, chalcopyrite, sphalerite, galena,
metres to only a few decimetres wide occurs in arsenopyrite and bournonite) was calculated
close proximity to the massive sulphides and from chemical assays (S, Cu, Zn, Pb, As and Sb)
locally contains neoblasts of andalusite, biotite, for drill core intervals using the Geo module of
magnetite and red garnet (Depauw, 2009). the HSC Chemistry software (Roine, 2009). The
calculated total for modal pyrite includes other
The contact between the altered volcanic rocks sulphide minerals which are not present in the list
and the metasedimentary rocks of Rockliden is for the element-to-mineral conversion yet might
steeply dipping (Fig. 2c) and the sedimentary be present in the ore, for example pyrrhotite.
rocks are generally younging to the north Bournonite was chosen to represent the Sb-
(Depauw, 2009; Raat and Årebäck, 2009). The bearing minerals in these calculations. Based
geometry of the massive sulphide bodies is on the sphalerite and chalcopyrite content of the
complex due to deformation (Fig. 2b and 2c). plain sulphide fraction (i.e. sulphides calculated
Structural features generally trend E–W (Evins, to 100 wt-%) three preliminary classes were
2011). Two deformation events were proposed GH¿QHGDVSKDOHULWHFKDOFRS\ULWHDQGS\ULWH
E\0DWWVVRQDQG+HHURPD  D¿UVWIROGLQJ rich class. These classes were further subdivided
event with N–S trending fold axis and a second by their Ag and Sb content representing critical
event with E–W trending fold axis. However, bonus and deleterious elements.
the E–W trending fold axes are prominent in
the massive sulphides and the folds are tight, Boliden’s database of chemical assays from the
nearly concentric at the outcrop of the massive Rockliden exploration project were imported and
sulphide mineralisation (Raat and Årebäck, interpolated using a spheroidal model for kriging
2009; Evins, 2011). These folds are occasionally with Leapfrog 3D modelling software (Zaparo
accompanied by axis-parallel shear zones with Ltd., 2009) and qualitatively evaluated together
an apparent sinistral offset of up to 30 m (Evins, ZLWK WKH SUHOLPLQDU\ FODVVL¿FDWLRQ %DVHG RQ
2011). Additionally, NW–SE and NE–SW this evaluation, six drill cores were selected
striking and steeply dipping faults transect the and 59 samples collected from the preliminary
Rockliden area (Mattsson and Heeroma, 1985). GH¿QHG FODVVHV WKURXJKRXW WKH VSDWLDO H[WHQW
Dextral offset was observed on a late shear zone of the deposit for the study of variability in
cutting the massive sulphide mineralisation mineralogy. All 59 samples were assayed by the
(Evins, 2011). ALS Minerals Division (Piteå, Sweden) for 64
major and minor elements. The accuracy and
Mattsson and Heeroma (1985) sketched precision of the data is stated with +/- 5 % for
an idealised mineralogical zoning for the elements at high level concentration and +/- 10
undeformed massive sulphide mineralisation % for elements on trace level concentration.
DW 5RFNOLGHQ 0RGL¿FDWLRQV RI WKH 5RFNOLGHQ The ranges of the analytical precision affect
massive sulphides and deviations from the the recalculated modal mineralogy in the same
idealised zoning pattern are thought to be due to order of magnitude. The geochemical data were
deformation and metamorphism (Depauw, 2009; studied against reference data from the Boliden
Raat and Årebäck, 2009; Minz, 2013; Fig. 2c). database for Rockliden host rock material.

5
Polished thin sections were prepared by silicate minerals (Table 3). In the following these
Vancouver Petrographics Ltd (Canada) and are referred to as host rocks and geochemical
were studied under an optical microscope WRROVDUHHPSOR\HGIRUWKHLUFODVVL¿FDWLRQ
(Nikon ECLIPSE E600 POL, Luleå University
RI 7HFKQRORJ\  WR UH¿QH WKH +6& &KHPLVWU\ $VVKRZQLQ¿JXUHIHOVLFYROFDQLFDQGPHWD
EDVHG SUHOLPLQDU\ FODVVL¿FDWLRQ DQG WR SURYLGH sedimentary rocks form the main host to the
additional textural information. Further, Rockliden massive sulphide mineralisation.
mineralogical studies were continued with The felsic volcanic rocks constitute the
energy-dispersive X-ray spectroscopy (EDS) in stratigraphic footwall and will be described in
the scanning electron microscope (SEM) and the PRUHGHWDLOEHORZ6HYHUDOPD¿FURFNW\SHVDUH
chemical composition of minerals was measured also present, locally in close relationship and
by wavelength-dispersive X-ray spectroscopy SDUWO\LQWHU¿QJHULQJZLWKWKHPDVVLYHVXOSKLGHV
(WDS) at the electron probe micro-analyser. 0D¿FURFNVVKRZLQJDQHOHYDWHG6EFRQWHQWDUH
EDS and WDS analysis were conducted using included in this study.
the following instrumentation with the following
settings: Merlin SEM (Zeiss Gemini, FESEM, 4.1. Petrography and geochemistry of host
Luleå University of Technology, Sweden) with rocks
acceleration voltage of 20 keV and emission
current of 0.9 to 1.1 nA and JEOL JXA-8200 From drill core logging, three main rock types
microanalyser (Center of Microscopy and DUH GLVWLQJXLVKHG IHOVLF YROFDQLF PD¿F DQG
Nanotechnology, Oulu University, Finland; sedimentary host rocks to the massive sulphides.
Table 1) with acceleration voltage of 15 keV, Regarding the base-metal sulphide minerals,
emission current of 15 to 30 nA and a beam PD¿F URFNV WHQG WR FRQWDLQ PRUH FKDOFRS\ULWH
diameter of 5 μm. than the felsic volcanic and sedimentary rocks,
both of which have higher sphalerite content
Table 1. Standards used forEPMA/WDS analysis. (Fig. 3a, cf. Table 2). The three host rock types
element standard are described in more detail below.
Si Wollastonite
Ca Wollastonite 4.1.1. Felsic volcanic rocks
F BaF2 The volcanic rocks cover an area of 25 km2
V V metal (Fig. 3a). The unaltered felsic volcanic rocks
K Orthoclase are characterised by porphyritic textures with
Al Al2O3 variable abundance of feldspar and quartz
Cr Chromite phenocrysts. Feldspar phenocrysts range in
Sn Sn metal length from 1 to 5 mm and, within feldspar-rich
Mg MgO units, they form up to 10 area-% of the rock (Raat
Mn Mn metal et al., 2011). Both plagioclase and alkali feldspar
Ti Ti metal SKHQRFU\VWV DUH LGHQWL¿HG 4XDUW] SKHQRFU\VWV
Na Jade constitute generally less than 5 area-%, with a
Fe Hematite maximum size of ca. 5 mm (Raat et al., 2011).
Zn Zn metal )ROLDWLRQ LQ WKH YROFDQLF URFNV LV GH¿QHG E\
alignment of phyllosilicates (Fig. 4b). In most
4. Results FDVHVWKHJURXQGPDVVRIWKHVHURFNVLVWRR¿QHWR
be resolved by optical microscopy (Fig. 4a and b).
Based on the total calculated sulphide content
the following broad distinction was made for the 7KH YROFDQLF URFNV SORW PDLQO\ LQ WKH ¿HOG RI
drill core samples: massive sulphides (samples rhyolites and dacites in the TAS diagram (Le
with > 50 wt-% total sulphide content) and Maitre et al., 1989; Fig. 3b) and are mostly found
disseminated to semi-massive sulphides (< 50 in the range of rhyolites and dacites in alteration
wt-% total sulphide content). The latter group box-plots, i.e. they plot within the lower part of
is dominated by silicon (Table 2) and hence by the box which composes analysis of the least

6
Table 2. Average concentrations of chalcophile and lithophile elements in disseminated and massive sulphide samples.
S Cu Zn Pb As Au Ag Sb Sn Bi Hg Cd In Mo Co Ni W
Classa) No.b)
wt-% wt-% wt-% wt-% wt-% ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm
Py-Mgt-Po-Ccp-Sp 4 35.78 3.12 1.86 0.09 1.45 0.1 39.3 202.5 23.8 12.4 51.9 36.0 21.7 2.7 118.2 11.2 1.7
Po-Ccp/Sp 6 26.72 2.96 13.61 2.29 0.50 0.0 120.7 1823.3 992.9 30.6 389.8 287.2 25.5 5.2 41.7 12.6 1.6
Py 2 46.50 0.49 2.55 0.29 0.12 0.0 42.0 516.5 12.8 7.3 185.3 48.3 13.6 10.8 27.4 9.2 1.5
Py-Ccp-Sp 8 39.23 4.67 3.21 0.32 1.02 0.0 102.8 861.2 59.7 15.7 193.8 60.3 40.1 4.9 117.6 10.3 6.4
Py-Sp-Ccp 5 35.16 3.13 7.45 0.53 0.64 0.0 76.7 596.5 31.5 40.2 261.1 159.2 24.0 3.5 44.2 6.9 4.2
Py-Sp 8 38.61 0.63 10.12 0.79 0.24 0.0 35.6 816.7 30.7 13.0 358.8 192.3 18.4 4.9 31.0 10.5 3.1
felsic (altered) 9 4.61 0.16 0.99 0.36 0.17 0.1 53.7 2804.8 56.9 15.5 27.0 29.6 2.4 3.6 18.0 4.9 5.0
PD¿F 7 3.10 0.93 0.32 4.13 1.24 0.2 279.8 7848.6 104.6 156.9 28.1 10.5 1.3 0.9 21.6 56.0 1.0
sed 3 1.59 0.33 0.14 0.75 0.12 0.0 48.8 1362.0 51.6 28.4 5.4 3.4 1.2 2.6 7.8 17.2 1.8
MgO CaO MnO Fe Zr Nb Y La Cr V

7
SiO2 Al2O3 CO2 Na2O K2O P2O5 TiO2
Classa) No.b)
wt-% wt-% wt-% wt-% wt-% wt-% wt-% wt-% wt-% wt-% wt-% ppm ppm ppm ppm ppm ppm
Py-Mgt-Po-Ccp-Sp 4 5.35 0.37 0.56 0.02 0.02 1.33 0.53 0.04 42.68 0.00 0.00 4.1 1.7 1.5 1.4 20.5 4.8
Po-Ccp/Sp 6 10.66 1.48 3.69 0.15 0.31 2.06 3.43 0.10 28.78 0.01 0.00 17.8 1.4 4.4 4.9 5.8 9.3
Py 2 1.93 0.27 1.70 0.03 0.02 0.81 1.30 0.06 41.80 0.00 0.00 2.2 1.0 1.2 0.7 0.5 5.0
Py-Ccp-Sp 8 6.26 0.21 0.93 0.01 0.01 0.91 0.61 0.05 38.78 0.00 0.00 1.2 0.8 1.1 0.9 1.4 4.8
Py-Sp-Ccp 5 7.83 0.51 1.28 0.04 0.05 1.60 0.85 0.06 35.32 0.00 0.00 5.2 1.1 1.6 1.6 11.4 7.2
Py-Sp 8 3.42 0.31 1.59 0.01 0.04 1.04 1.41 0.09 37.19 0.00 0.00 4.1 1.1 1.6 1.2 1.4 4.4
felsic (altered) 9 64.79 11.32 0.09 0.95 2.60 1.95 1.59 0.06 7.09 0.08 0.42 203.8 9.1 30.3 37.0 21.4 20.7
PD¿F 7 45.38 11.44 0.69 2.00 0.47 9.40 8.06 0.18 7.22 0.14 0.56 59.2 2.9 18.0 10.5 406.1 202.7
sed 3 66.10 13.98 0.27 2.38 2.15 2.74 2.57 0.03 3.49 0.15 0.44 183.0 9.8 25.3 44.9 52.3 62.3
D $EEUHYLDWLRQV&FS±FKDOFRS\ULWH0JW±PDJQHWLWH3R±S\UUKRWLWH3\±S\ULWH6S±VSKDOHULWHIHOVLF DOWHUHG ±IHOVLF DOWHUHG YROFDQLFURFNVPD¿F±PD¿FG\NHVVHG±VHGLPHQWDU\URFNV
b) Number of samples.
Table 3. Relative abundance of minerals in the sulphide fraction and non-sulphide fraction
(xxx – major, xx – common, x – minor, (x) – trace).
sulphide fraction
Classa) No. Py Po Mgt Sp Ccp Apy Sb-bearing minerals
Po–Sp–Ccp 6 x xxx (x) xx xx-x x Ttr, Bour, Gd (in Sp-rich samples), +/- Boul
Py 3 xxx xx (x) xx x (x) Ttr, Gd, Hg-Ag-Sb sulphide minerals
Ttr, Bour, Gd (Gd preferentially in samples from
Py–Sp 9 xxx xx-x x xx x (x)
deeper parts of the deposit)
Py–Sp–Ccp 5 xxx xx x xx x (x) Ttr, Hg-Ag-Sb sulphide minerals
Py–Ccp–Sp 6 xxx xx (x) x xx x Ttr, Gd, Bour
Py–Po–Mgt–Ccp–Sp 3 xxx xx xx x x (x) Gd, Hg-Ag-Sb sulphide minerals
PD¿F 7 xx xx xxx x Bour, Ttr, Mene, Boul, Gd
felsic 1 xxx xx
felsic (altered) 9 xxx xx xx x (x) Gd, Sb-bearing sulphosalts
sed 3 xxx x (x) x (x)
Clay
Classa) No. Qtz Fsp Ser Bt Amp Carb others (minor or trace minerals)
Min
Po–Sp–Ccp 6 xxx (x) x x (x) xxx cassiterite, stannite
Py 3 xxx x xx
Py–Sp 9 xxx (x) x (x) xxx (scheelite)
Py–Sp–Ccp 5 xxx x x (x) xx epidote group minerals
Py–Ccp–Sp 6 xxx x xx (x) xx
Py–Po–Mgt–Ccp–Sp 3 xx xxx (x) xx
PD¿F 7 x xx a)
x x xxx (x) Zn-Chr, Rt, Ilm, Ttn, Ap, Cal
felsic 1 xx xxx b)
xx
felsic (altered) 9 xxx x xx xx (x) (x) (x) Fe-ghanite, Epi, Anh, And, Fl
sed 3 xx xxx (x)
$EEUHYLDWLRQIHOVLF DOWHUHG ±IHOVLF DOWHUHG YROFDQLFURFNVPD¿F±PD¿FG\NHVVHG±VHGLPHQWDU\URFNV$PS±DPSKLEROHV$QG
– andalusite, Anh – anhydrite, Ap – apatite, Apy – arsenopyrite, Bt – biotite, Boul – boulangerite, Bour – bournonite, Cal – calcite,
Carb – carbonate minerals, Ccp – chalcopyrite, Zn-Chr – Zn-rich chromite, ClayMin – phyllosilicates, Epi – epidote group minerals,
)O±ÀXRULWH)VS±IHOGVSDUPLQHUDOV a) plagioclase, b) alkali-feldspar), Gd – gudmundite, Il – ilmenite, Mene – meneghinite, Mgt –
PDJQHWLWH3R±S\UUKRWLWH3\±S\ULWH4W]±TXDUW]6HU±VHULFLWH6S±VSKDOHULWH7WQ±WLWDQLWH7WU±WHWUDKHGULWH

altered rocks (cf. Large et al., 2001; Fig. 3d). The The volcanic rocks plotting outside the box of
JHRFKHPLFDOFODVVL¿FDWLRQRIWKHYROFDQLFURFNV least altered rocks, i.e. AI > 80 (cf. Large et al.,
is supported by the immobile trace element plot 2001), comprise quartz and sericite (Fig. 4c)
by Winchester and Floyd (1977), which shows with varying amounts of phyllosilicates, partly
WZR FOXVWHUV ZLWKLQ WKH UK\ROLWLF±GDFLWH ¿HOG LGHQWL¿HG DV FKORULWH JURXS PLQHUDOV )LJ G 
(Fig. 3c). However, systematic differences in In reference to optical microscopy, they appear
the mineralogical composition, i.e. potential to become more chlorite-rich with increasing
FRUUHODWLRQ ZLWK WKH JHRFKHPLFDO FODVVL¿FDWLRQ CCPI index (Fig. 3d). Gahnite, with 7 to 8 wt-%
and clustering, have not been observed during Fe, is found locally in intensely altered samples
drill core logging. (Fig. 4f). Amphibole minerals were documented
LQRQHVDPSOHRIKRVWURFNFODVVL¿HGDVDIHOVLF
All studied samples have been taken close to volcanic rock (Fig. 4e). Locally, the altered
the massive sulphide mineralisation and the volcanic rocks show elevated values of Sb and
felsic rocks described below are all affected are transected by quartz veins or calcite veinlets
by alteration. Also, those plotting in the box of (Fig. 4b). Besides gudmundite, various Sb-
least altered rocks show features of alteration bearing sulphosalts were documented with EDS
VXFKDVVLOLFL¿FDWLRQ )LJE EXWUHPQDQWVRI analysis.
alkali feldspar crystals are locally preserved.

8
a Legend:
b
Ccp
sedimentary rock 14

volcanic rock
12
volcanic rock, altered

(Na2O + K2O), wt%


10
mafic dykes and rocks
8

0
Py Sp 40 50 60 70 80
Si2O, wt%
c d 100
Ep Cal Dol Ank Chl Py

0.1 80
Zr/TiO2

2
60
CCPI

0.01 40

20 Ill

0 Ab Adu
0.1 0.2 1 2 0 20 40 60 80 100
Nb/Y AI

e f 300

50
250

40
200
La, ppm

V, ppm

30
150

20
100

10 50

0 0
0 1 2 3 4 5 0 150 300 450 600 750
Yb, ppm Cr, ppm

Fig. 3.0LQHUDORJLFDODQGFKHPLFDOFODVVL¿FDWLRQRI5RFNOLGHQKRVWURFNV EDVHGRQVDPSOHVFROOHFWHGIRUWKLVVWXG\


and ca. 380 lithological reference samples from the host rock database (Exploration Department, Boliden Mines)):
a) chalcopyrite(Ccp)-sphalerite(Sp)-pyrite(Py) ternary plot based on mineralogical composition of the sulphide fraction
calculated with HSC Chemistry (Roine, 2009); b)FKHPLFDOFODVVL¿FDWLRQEDVHGRQ/H0DLWUHHWDO  ZLWKYLROHWGRWVUHS-
UHVHQWLQJOHDVWDOWHUHGIHOVLFYROFDQLFURFNV EDVHGRQFODVVL¿FDWLRQE\DOWHUDWLRQER[SORW c) FKHPLFDOFODVVL¿FDWLRQEDVHGRQ
Winchester and Floyed (1977), d)FKHPLFDOFODVVL¿FDWLRQEDVHGRQDOWHUDWLRQER[SORWDIWHU/DUJHHWDO  e) discrimina-
tion of sedimentary rocks based on Yb-La diagram, f)GLVFULPLQDWLRQRIPD¿FURFNVLQ&U9GLDJUDP0LQHUDODEEUHYLDWLRQV
Ab – albite, Adu – “adularia”, Ill – illite, Chl – chlorite, Py – pyrite, Ank – ankerite, Dol – dolomite, Cal – calcite, Ep – epidote.

9
Fig. 4. Volcanic host rock and alteration types (optical microscopy, plane and cross polarised light): a) volcanic host
URFNZLWKIHOGVSDU )VS DQGTXDUW] 4W] SKHQRFU\VWVb)VOLJKWO\VHULFLWHDOWHUHGURFNZLWKTXDUW] 4W] YHLQV PLQRU
alteration minerals are andalusite (And) and calcite (Cal)), c) intensively sericite-altered rock (Ser – sericite), d) alter-
DWLRQDVVHPEODJHFRPSULVLQJTXDUW] 4W] FKORULWHJURXSPLQHUDOV &KO DQGELRWLWH %W e) amphibole crystals (Amp)
and epidote group minerals (Ep), f)TXDUW] 4W] =QVSLQHO =Q6SO)HULFKJDKQLWH DQGELRWLWH %W 

4.1.2. Turbiditic sediments and greywackes zones (decimetre to several metres in width)
The siliciclastic sedimentary rocks comprise RI VLOLFL¿FDWLRQ LQ FRQWDFW ZLWK WKH PDVVLYH
DOWHUQDWLRQVRIVKDOHDQG¿QHJUDLQHGVDQGVWRQH sulphides. However, primary sedimentary
(Fig. 5a). In contrast to the volcanic host textures such as graded bedding, indications
rock, alteration is less pronounced in the of soft-sediment deformation (Evins, 2011)
sedimentary rocks. Drill cores display bleached and fracturing in the form of calcite and quartz

10
Fig. 5.7H[WXUHVLQVHGLPHQWDU\DQGPD¿FKRVWURFNV WUDQVPLWWHGFURVVSRODULVHGOLJKWDQGUHÀHFWHGOLJKWPLFURVFRS\ 
a) variation in the grain size of quartz in the sedimentary rocks. Bedding is transected by calcite veinlets (Cal),
b) PD¿FG\NHZLWKFOXVWHUVRIDPSKLEROH $PS FU\VWDOVLQDQDPSKLEROHELRWLWH $PS%W GRPLQDWHGJURXQGPDVV
c)DPSKLEROH $PS ELRWLWH %W DQGWLWDQLWH 7WQ FU\VWDOVLQDPD¿FG\NHVDPSOHd) Zn-rich chromite (Chr) enclosed
E\PDJQHWLWH 0JW DQGDVVRFLDWHGZLWKDPSKLEROH $PS DQGDUVHQRS\ULWH $S\ LQDPD¿FG\NHVDPSOH

veinlets were documented (Fig. 5a). In contact a similar thickness as the dolerite dykes in drill
with the massive sulphides, the fracturing is core intersections, but show no distinct chilled
locally accompanied by elevated concentration margins. To simplify, they will be referred to as
of Sb in drill core assays. The sedimentary rocks PD¿F G\NHV LQ WKLV SDSHU *HQHUDOO\ WKH PD¿F
are mostly distinct from the (felsic) volcanic dykes are composed of amphibole minerals,
rocks in Yb–La and Cr–V trace element plots phyllosilicates such as biotite, and plagioclase
(Fig. 3e and 3f) and partly distinct in e.g. Al2O3/ IRUPLQJD¿QHJUDLQHGJURXQGPDVV )LJE 7KH
TiO2–Zr/Al2O3 plots. grain size of amphibole varies from 50 to 500 μm
(Fig. 5b and 5c) and chemically it is hornblende,
4.1.3. 0D¿FURFNV with relatively large variation in Fe, Mg, Si and
Dolerite dykes are a few decimetres to metres Al content (Table 4). The main carbonate mineral
wide and are largely aphyric (Raat and Årebäck, LVFDOFLWHLQ¿OOLQJIUDFWXUHVWRJHWKHUZLWKTXDUW]
2009). They have chilled margins and 1 to 5 mm DQGFKORULWHJURXSPLQHUDOV0RVWRIWKHPD¿F
long plagioclase crystals commonly with ophitic dyke samples contain traces of rounded Zn-rich
WH[WXUH DQG FDUERQDWH±TXDUW]¿OOHG DP\JGXOHV chromite grains, with ca. 9 wt-% Zn, surrounded
in the central parts (Raat and Årebäck, 2009). by Cr–V-rich magnetite rims or ilmenite rims
&RPSOHPHQWDU\PD¿FURFNW\SHVWHUPHGPD¿F )LJG7DEOH ,QD&U±9GLDJUDPWKHPD¿F
sills and dykes by Raat and Årebäck (2009), have dykes form a distinct cluster at Cr > 300 ppm and

11
4.2. Massive sulphide mineralisation types
(massive)

91.84

92.18
a $EEUHYLDWLRQV3OJ±SODJLRFODVH$PS±DPSKLEROH%W±ELRWLWH(S±HSLGRWHJURXS*UW±JDUQHWJURXS0JW PD¿F ±PDJQHWLWHLQPD¿FG\NHV&KU±FKURPLWH
Mgt

0.41
-
-
-
-
-
-

-
-
-
-

-
The variation in chemical composition of the
massive sulphide mineralisation is shown in
¿JXUHDDQGWKHPHWDOSURSRUWLRQRIWKHLQIHUUHG
PD¿F

84.92

92.18
Mgt

0.33

0.14
5.53
0.13
0.61
0.38
-

-
-
-
-

-
mineral resource at Rockliden (New Boliden,
 SORWLQWKH=Q±3E±&X¿HOGLQWKH&X±3E±
PD¿F
Table 4. Selected EPMA/WDS analysis showing the compositional variation of some silicate and oxide minerals.

101.22
27.23

50.24

12.56
Zn ternary plot of Franklin et al. (1981).
8.59
0.76
0.35

0.60

0.69
0.19
Chr

-
-

-
The massive sulphides show variable textures
(felsic)

100.94
37.00
21.13

22.74
17.33
1.29

1.31
Grt

in the drill core. Banding is due to the variation


-

-
-
-
-
-
in grain size of pyrite or in the abundance
38.64
27.99

24.66

97.91
of sphalerite in the groundmass surrounding
6.48
Epi

-
-
-

-
-
-
-
-
-
the pyrite grains (Fig. 7). Locally, irregular
chalcopyrite patches (referred to as chalcopyrite
PD¿F
41.09
13.78

20.79
10.60

97.02
Bt 2

0.06

9.01
0.12
0.13
0.93

0.75
-

mesh texture, Fig. 7) or abundant host rock clasts


are observed in the massive sulphides.
PD¿F
37.30
16.51

14.77
10.56

15.28

96.82
Bt 1

0.09
0.27
1.83

,Q ¿JXUH D D FRQWLQXRXV PDVVLYH VXOSKLGH


-

-
-
-

intersection is shown with a gradual change in


IHOVLF±DOWHUHGIHOVLFYROFDQLFURFNVPD¿F±PD¿FG\NHV06±PDVVLYHVXOSKLGHVDPSOHV

Cu/(Cu+Zn) ratio from the sericite and chlorite


Amp 2
(felsic)
50.64

16.95

13.08

97.91
5.99
0.60

0.44

9.63
0.32

0.13
-

-
-
-

altered volcanic host rock into the massive


sulphides. Arsenic is found in relatively high
Amp 1
(felsic)
55.21

20.81

13.59

97.98
2.61
0.24

0.12

5.10
0.29

concentrations in the semi-massive sulphides


-
-
-
-
-

(Fig. 7a). The stratigraphic upper contact of


Amp 2

the massive sulphide intersection with the


PD¿
46.17
12.62

13.63

12.00

98.35
11.23
1.24

0.12

0.28
0.37
0.43
0.19
-
-

sedimentary host rock is sharp (Fig. 7a). At this


contact, the sedimentary host rock is quartz-
PD¿F
Amp 1

53.12

18.70

12.52

98.23
3.97
0.51

8.74
0.30

0.26

veined and characterised by elevated Sb grade.


-

-
-
-

Most other massive sulphide intersections are


GLVUXSWHG E\ PD¿F G\NHV EXW DOVR LQWHUYDOV
PD¿F

100.26
47.37
33.56

16.70
Plg 2

1.97
0.24
0.04

0.34
-
-
-
-
-
-

of altered felsic volcanic rocks. Two of the


b) LLD – lower limit of detection.

GLVUXSWHGPDVVLYHVXOSKLGHSUR¿OHVDUHVKRZQLQ
PD¿F
58.82
25.83

99.70
Plg 1

7.39

0.04
7.43
0.18

¿JXUHEDQGF,QQHLWKHUSUR¿OHDUHV\VWHPDWLF
-

-
-
-
-
-
-

changes in the Cu/(Cu+Zn) ratio of the massive


LLDb)

sulphides observed. Massive sulphides showing


0.15
0.09
0.07
0.07
0.03
0.07
0.09
0.08
0.10
0.06
0.10
0.08
0.17
-

chalcopyrite mesh textures are more common in


minerala)

SUR¿OHVZLWKKLJK&XFRQWHQW FRPSDUH)LJD
Cr2O3
wt-%

Al2O3
Na2O

MnO
MgO

Total
V2O3
CaO

ZnO
TiO2
SiO2

K2O

FeO

and Fig. 7b with Fig. 7c). Some of the interrupted


massive sulphide intersections show high Ag (>
100 ppm) and Sb (> 1000 ppm) contents over
9!SSP )LJI7DEOH 7KHVWXGLHGPD¿F parts or the entire massive sulphide intersections.
dyke samples show elevated concentrations of Relatively narrow (max. 2 m wide) massive
Sb (Table 2), and bournonite, meneghinite and sulphide intervals with abundant host rock
tetrahedrite are the main Sb-bearing minerals clasts are occasionally found (Fig. 7b) and these
7DEOH   0RUH GHWDLOHG UHVHDUFK RQ DOO PD¿F intervals are typically dominated by pyrrhotite.
rock units, occurring within and around the On the basis of optical microscopy, the massive
Rockliden deposit, is needed in order to give a sulphides are grouped according to the dominant
IXOO SHWURORJLFDO GHVFULSWLRQ DQG WR FRQ¿QH WKH sulphide (and oxide) minerals, i.e. pyrite (Py),
WLPLQJ RI IRUPDWLRQ RI GLIIHUHQW PD¿F URFNV pyrrhotite (Po), magnetite (Mgt), chalcopyrite
relative to the ore formation. (Ccp) and sphalerite (Sp) (Fig. 6b). Some
textures of the massive sulphide samples are

12
a Cu / wt% Base metal classes grades: b Ccp
inferred mineral

Cu group
resource at Rockliden
massive sulphides
(analysis from drill core)

Zn-Cu group
Mineral classes in the
massive sulphides:
Py+Po+Mgt-Ccp-Sp
Po-Ccp
Po-Sp
Py
Py-Ccp-Sp
Py-Sp
Py-Sp-Ccp

Pb / wt% Zn-Pb-Cu group Zn / wt% Py Sp


Fig. 6. a) Zn–Cu–Pb ternary plot showing the variation of chemical composition of massive sulphide samples (424
assays with total sulphide content .50 wt-%) and the grade of the inferred mineral resource at Rockliden (New Boliden,
 &ODVVL¿FDWLRQOLQHVIRUPDVVLYHVXOSKLGHGHSRVLWVDVVRFLDWHGZLWKYROFDQLFURFNVDUHWDNHQIURP)UDQNOLQHWDO
(1981) and b)FDOFXODWHGS\ULWHFKDOFRS\ULWHDQGVSKDOHULWHFRQWHQWIRUWKHPLQHUDORJLFDOJURXSVGH¿QHGIRUWKHVWXGLHG
massive sulphide (33 samples).

VKRZQLQ¿JXUH,WPXVWEHQRWHGWKDWWUDQVLWLRQV (Table 2, cf. Fig. 6b). Samples with Cu values


between the various groups are common; there > 2 wt-% are characterised by irregular-shaped
DUH QR ZHOOGH¿QHG ERXQGDULHV EHWZHHQ WKHVH chalcopyrite ± pyrrhotite meshes, imposing
mineralisation types. a deformation texture on these samples (Fig.
8d). Magnetite grains are mostly rounded to
In the pyrrhotite–sphalerite and pyrrhotite– subangular and range from 100 to 200 μm in
chalcopyrite groups (Po–Sp, Po–Ccp), pyrrhotite diameter (Fig. 8d). Occasionally, they contain
is the main sulphide mineral. Locally, isolated inclusions of sulphides, mainly pyrite.
grains of pyrite (up to 500 μm diameter) are
preserved (Fig. 8a). Distinctively high grades of The Py–Sp–Ccp and Py–Ccp–Sp groups can be
more than 1000 ppm Sn are measured in clast- subdivided by their Zn content (Py–Sp–Ccp with
rich, isolated (max. 2 m width), massive sulphide Zn > 5 wt-%, Py–Ccp–Sp with Zn < 5 wt-%, cf.
intersections (Fig. 7b). The Sn-bearing oxide Fig. 6b). The Cu content is generally above 2
and sulphide minerals include cassiterite and wt-% (Table 2, cf. Fig. 6b). Samples with high
stannite, respectively. The silicate clasts form chalcopyrite content tend to also have relatively
so-called augen textures (Fig. 8b; cf. Petruk, high pyrrhotite content. They often lack distinct
2000). The clasts comprise mainly quartz, banding, but deformation textures such as
biotite, chlorite and other phyllosilicates, minor chalcopyrite meshes are documented (Fig. 8d-f,
epidote group minerals and feldspars. Beside Fig. 7). A close association of arsenopyrite with
silicates, carbonates are the main non-sulphide chalcopyrite and pyrrhotite is locally observed
phases in samples of this group (Table 3). Zoning in samples of this group (Fig. 8f). Some samples
of dolomite group minerals is observed (back- of the Py–Ccp–Sp group contain both abundant
scattered electron imaging at the SEM and EDS magnetite and pyrrhotite. Thus, they are allocated
analysis), comprising Fe-rich cores surrounded to a separate group, termed Py–Po–Mgt–Ccp–
by Fe-poor margins, which in turn, might be Sp group (Fig. 6b). However, they do not form
rimmed by calcite. Magnetite is rarely found in a separate cluster in the Py–Ccp–Sp ternary plot
samples of this group. (Fig. 6b).
Massive sulphide samples of the pyrite (Py)
group show banding related to the clustering The pyrite-sphalerite (Py–Sp) group has
and grain size variation of pyrite (50 – 500 μm). relatively low Cu content (generally below 2
The packing of pyrite grains locally reaches wt-%), while the Zn content is relatively high
high density (Fig. 8c). The combined Cu and Zn (ranging from 7 to 15 wt-%, Table 2, cf. Fig.
content of this group is generally below 6 wt-% 6b). Pyrite grains often show embayments of

13
Fig. 7.0DVVLYHVXOSKLGHLQWHUVHFWLRQVLQGULOOFRUH IRUORFDWLRQVHH¿JXUHF a) undisrupted, b) disrupted (massive
sulphide package enclosed by sedimentary rocks), c) disrupted (massive sulphide intersections in contact with brecciat-
ed altered volcanic host rock at ca. 172 m and sedimentary rocks below 245 m). Mineral abbreviation: Ccp – chalcopy-
rite.

sphalerite (atoll texture, Fig. 8g). Grain size of So far, no systematic differences are noticed
pyrite ranges from 200 to 500 μm (up to 1000 μm in the Sb mineralogy of different base-metal
diameter) and it can host inclusions of sphalerite groups of the massive sulphides (Table 3).
and occasionally galena or tetrahedrite (Fig. Based on optical microscopy it is observed
8h). Despite sphalerite forming the dominant that tetrahedrite, bournonite, meneghinite and
mineral embedding pyrite grains, pyrrhotite boulangerite are associated with sphalerite and
is locally present and in this case imposes a galena, as well as pyrrhotite and arsenopyrite,
banded texture on the ore. Locally, pyrrhotite is whereas gudmundite is often found associated
found in diablastic intergrowth in the sphalerite with chalcopyrite, pyrrhotite, silicate minerals
groundmass or as fractures within pyrite grains and calcite (Minz et al., 2013).
(Fig. 8e and h). Magnetite is commonly found in
this group.

14
Fig. 8. 0DVVLYHVXOSKLGHV UHÀHFWHGOLJKWPLFURVFRS\ a) large fractured pyrite grain in a pyrrhotite-chalcopyrite-sphalerite
groundmass and intergrowths of galena and bournonite, b) augen texture of silicate minerals hosting gudmundite, c)
dense packing of pyrite grains, d) pyrite-magnetite dominated sample with chalcopyrite meshes, e) orientated texture
of pyrrhotite and partly also chalcopyrite, f) close association of chalcopyrite, pyrrhotite and arsenopyrite (details given
DWKLJKHUPDJQL¿FDWLRQLQWKHLQVHUW g) pyrite showing atoll texture (white frame) and sphalerite-pyrrhotite diablastic
intergrowth, h) cataclastic pyrite associated with pyrrhotite, galena and native antimony. Mineral abbreviation: Ant – na-
tive antimony, Apy – arsenopyrite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mgt – magnetite, Po – pyrrhotite,
Py – pyrite, Sp – sphalerite.

15
5. Discussion host rocks, and the base-metal content of
Rockliden (Fig. 6a; cf. Barrie et al., 1999), the
Based on the sulphide content, massive (> mineralisation is suggested to belong to the
50 wt-%), semi-massive (15 to 50 wt-%) and bimodal felsic VHMS group (cf. Fig. 3c), in
disseminated (< 15 wt-%) mineralisation types accordance with earlier studies by Depauw
are distinguished in the Rockliden deposit. (2009) and Raat and Årebäck (2009). It is
The latter is dominated by silicate minerals also consistent with Mattsson and Heeroma’s
and forms the host rock to the mineralisation. (1985) schematic for pre-deformational zoning
Several lithologies can be distinguished in the within the Rockliden deposit, which is typical
JHRFKHPLFDO FODVVL¿FDWLRQ GLDJUDPV IHOVLF of massive sulphide deposits associated with
YROFDQLFURFNVVHGLPHQWDU\DQGPD¿FURFNV,Q felsic volcanic rocks (Galley et al., 2007). The
the following the characteristics of the studied felsic volcanic rocks show spatially extensive
lithologies are summarised and the distinguished alteration zones and are dominated by sericite
host rocks are discussed, as is the mineralogy of and quartz with minor chlorite group minerals.
the massive sulphides and the implications on The sequence of alteration zones resembles
mineral processing. the zoning outlined by Galley et al. (2007) for
the bimodal felsic VHMS group (cf. Raat and
5.1. Lithological constraints on the Årebäck, 2009).
FODVVL¿FDWLRQRIWKH5RFNOLGHQGHSRVLW
5.2. Mineralogical indications of
6HYHUDO PD¿F URFN W\SHV KDYH EHHQ GHVFULEHG deformation and metamorphism of the
from the Rockliden area, including gabbro massive sulphides
intrusions, dolerite dykes and basalt (Mattsson
and Heeroma, 1985; Raat and Årebäck, 2009; The metamorphic grade of the host rocks varies
'HSDXZ   0D¿F URFNV VXFK DV JDEEUR from lower greenschist facies, up to amphibole
intrusions and dolerite dykes have been facies (Depauw, 2009; Raat and Årebäck,
suggested to post-date the formation of the 2009). However, the geometry of the ore
massive sulphides at Rockliden (Söderlund et body is complex due to deformation and most
DO'HSDXZ 6RIDURQO\RQHPD¿F massive sulphide intersections are disrupted
URFNW\SHUHIHUUHGWRDVPD¿FG\NHVLVVWXGLHG by thin, partly brecciated host rock intervals.
LQ GHWDLO 7KHVH PD¿F G\NHV DUH GRPLQDWHG E\ Within the massive sulphides, textural features
amphiboles, and contain traces of Cr–V-rich related to metamorphism and deformation are:
magnetite, and can be distinguished in Cr–V-
diagrams (Fig. 3f). They are found in relatively (1) Banding may have been enhanced during
narrow drill core intersections, partly located deformation since galena, sphalerite,
within the massive sulphides. Although they pyrrhotite and chalcopyrite are suggested
represent a volumetrically minor host rock, they to be relatively ductile compared to pyrite
DUHH[SHFWHGWRLQÀXHQFHWKHFRQWHQWRIFULWLFDO at the maximum metamorphic conditions
metals (e.g. Sb and Ag) in the deposit and in for Rockliden (Clark and Kelly, 1973;
WKH UHVSHFWLYH ÀRWDWLRQ SURGXFWV 7KXV WKHLU Raat and Årebäck, 2009). However,
petrographic and geochemical differentiation the extent to which metamorphism is
IURP RWKHU PD¿F URFN XQLWV RFFXUULQJ LQ WKH responsible for this texture is unknown,
Rockliden area and their relative timing in as banding might also be a primary
relation to the massive sulphide mineralisation, feature or a recrystallisation event in
i.e. the potential overprinting effect on the unmetamorphosed “prototype” VHMS
mineralisation, should be studied in more detail. deposits (Eldridge et al., 1983; Mattsson
However, this is beyond the scope of this paper. and Heeroma, 1985).
It is shown that the sedimentary and felsic (2) Corrosion of pyrite grains (by sphalerite,
volcanic rocks can be largely distinguished by chalcopyrite and pyrrhotite) and
their Yb and La contents (Fig. 3e). Given the recrystallisation of pyrite and related
combination of felsic volcanic and sedimentary changes in grain sizes has been suggested

16
to occur during metamorphism (Petruk, WR VWUXFWXUDO PRGL¿FDWLRQ DQG GHIRUPDWLRQ
2000). However, rounding of pyrite at Rockliden (Minz, 2013). Moreover, the
grains and the formation of atoll textures occurrence of pyrrhotite-dominated sulphide
KDYHDOVREHHQDVFULEHGWRPRGL¿FDWLRQ pods within largely unaltered sedimentary rocks
by ore forming solutions (Eldridge et was interpreted as remobilisation along faults
al., 1983). Similarly to banding the cross-cutting the mineralisation, stressing the
impact of metamorphism on this texture importance of deformation on the shape and
is unknown for the Rockliden massive mineralogy of the Rockliden massive sulphide
sulphides. deposit (Depauw, 2009; Fig. 2c). However, the
(3) Locally, pyrrhotite is the predominate above listed metamorphic features are observed
Fe-bearing sulphide mineral in thin, on a microscopic scale and vary at a centimetre
massive sulphide intersections; it is to metre scale. Thus, their implementation into a
found marginal to massive sulphides or process-adapted geological model is challenging
enclosed by sedimentary rocks. Also, (cf. Minz, 2013) and evaluation in terms of
S\UUKRWLWH LV IRXQG WR ¿OO IUDFWXUHV LQ exploitation has not been performed at the
large pyrite grains (Fig. 8h). These current stage.
observations are in accordance with the
interpretation of pyrrhotite formation Regarding the trace mineralogy, the association
during metamorphism and deformation of Sb-bearing minerals is rather complex
(Mattsson and Heeroma, 1985; Depauw, at Rockliden (Minz et al., 2013). Such
2009). Furthermore, pyrrhotite is complex associations are known also for other
documented in diablastic intergrowth polymetallic sulphide ore deposits. For example,
with sphalerite, and pyrrhotite and bournonite is commonly observed adjacent
chalcopyrite often show foliated, to galena and tetrahedrite or meneghinite
elongated and schistose textures. These (Anderson, 1940; Wen et al., 1991; Wagner
features were ascribed to metamorphism and Cook, 1997). Additionally, Sb-bearing
of massive sulphides (Petruk, 2000). minerals are found intergrown with pyrrhotite
(4) Mattsson and Heeroma (1985) suggested at Rockliden (Fig. 8h). It is expected from aS2-
that the formation of magnetite is related temperature diagrams that sulphosalts such as
to secondary (ore-modifying) events as meneghinite and bournonite are stable in the
they observed magnetite bands cross- S\UUKRWLWH¿HOGDWKLJKHUWHPSHUDWXUHVLH!FD
cutting the banding in the massive 200°C (Mookherjee and Mishra, 1984). Given
sulphides. Similar observations were these considerations on the ore formation, a
made in this study. Moreover, magnetite K\GURWKHUPDO RUH PRGL¿FDWLRQ HYHQW FDQQRW
locally contains abundant inclusions be excluded at Rockliden. It was suggested in
of pyrite and pyrrhotite (Fig. 3.5e of earlier studies that remobilisation of Sb occurred
Minz (2013)), resembling magnetite during the intrusion of dolerite dykes and a
porphyroblasts, which were interpreted related increase in temperature (Depauw, 2009).
as metamorphic features in VHMS
deposits studied by Petruk (2000). 5.3. Mineralogical parameters in context of
mineral processing
It is suggested that metal zoning occurs on
the deposit scale, i.e. Cu dominated drill core Based on the qualitative mineralogical results
intersections become more abundant towards of this study, it is proposed that the content of
depth, i.e. ca. 400 m below surface (Raat and chalcopyrite and sphalerite can be calculated
Årebäck, 2009). By correlation this would mean reliably from Cu and Zn assays as they are the
that textures such as the chalcopyrite meshes main carriers of the base metals in the massive
are also expected to be more abundant at depth, sulphides. Exceptions are only Fe-rich gahnite
which is apparent from drill core logging. and Zn-rich chromite for Zn and tetrahedrite,
However, no clear metal zoning pattern was bournonite and meneghinite for Cu. In most
found and the absence of such pattern was related cases, these minerals are not expected to be a

17
VLJQL¿FDQWFDUULHURIHLWKHU&XRU=QDVWKH\DUH A relationship between the Sb mineralogy and
found as minor or trace minerals in the massive the rock types of Rockliden has been proposed
sulphides and host rocks. This means that the ZKHUHE\ WKH PD¿F G\NHV FRQWDLQ PRVWO\
Cu and Zn content of the massive sulphides (cf. bournonite, meneghinite or tetrahedrite; the
)LJD ZLOOEHUHÀHFWHGLQWKHYDULDWLRQRIWKH altered felsic volcanic rocks contain locally
chalcopyrite and sphalerite content. However, gudmundite and various Sb-bearing sulphosalts;
the modal mineralogy does not directly reveal and the massive sulphides contain tetrahedrite,
textural features. Some textural features bournonite or gudmundite as the main Sb-bearing
potentially affect the processing of the massive trace minerals (Minz et al., 2013). There are also
sulphides. For example pyrite–sphalerite indications that the Sb mineralogy of the massive
banding will cause anisotropic mineralogical sulphides might change with depth, i.e. below ca.
hardness contrast, and microtextures will control 400 m below surface (cf. Minz et al., 2013), and
grain and related liberation size (cf. Butcher, towards the northwest of the Rockliden deposit
2010). (Mattsson and Heeroma, 1985). Generally, the
Sb-bearing minerals are associated with base-
From the overlap of Py–Po–Mgt–Ccp–Sp and metal minerals and this association will have
Py–Ccp–Sp massive sulphide groups in the DQLQÀXHQFHLQWKHFRPPLQXWLRQFKDUDFWHULVWLFV
ternary Py–Ccp–Sp plot (Fig. 6b); it is clear that RIWKHRUHDQGWKHLUGLVWULEXWLRQGXULQJÀRWDWLRQ
the entire modal mineralogy cannot be calculated (Minz et al., 2013).
from chemical assays based on the limited
number of elements commonly assayed in drill 6. Conclusions
core. Furthermore, the direct calculation of pyrite
from simple Fe assays is not possible since other Based on the total sulphide content of whole rock
Fe-bearing minerals such as magnetite occur with samples, it is possible to distinguish between
pyrite and pyrrhotite in the massive sulphides. disseminated to semi-massive and massive
The importance of quantitative information on sulphides at the Rockliden deposit. The former
Fe-bearing minerals in massive sulphides was comprise the lithological units which are hosting
illustrated for the magnetite content in ore from the massive sulphides, including felsic volcanic
the Black Mountain polymetallic base metal DQG VHGLPHQWDU\ URFNV DQG PD¿F G\NHV )HOVLF
mine, South Africa (Williams and Holtzhausen, volcanic and sedimentary rocks form the main
2001). The magnetite content was shown to host rocks. The altered felsic volcanic rocks
LQÀXHQFH WKH FRPPLQXWLRQ SHUIRUPDQFH E\ show locally elevated concentrations in Sb,
increasing the hardness of the ore, and the often connected to quartz veining in the host
ÀRWDWLRQSHUIRUPDQFHE\ORZHULQJWKHJUDGHRI URFN $PRQJ WKH PD¿F URFN W\SHV RQH XQLW
valuable minerals (Williams and Holtzhausen, UHIHUUHG WR DV PD¿F G\NHV VKRZV HOHYDWHG
2001). In this study, no systematic variations in concentrations in Sb and, although it is found in
the magnetite content have been documented relatively narrow intervals, it is concluded to be
at Rockliden. In a study by Mattsson and an important rock type when the Sb distribution
Heeroma (1985), it was found that magnetite of the Rockliden ore is considered.
is more abundant in massive sulphides from 0DVVLYH VXOSKLGH VDPSOHV DUH FODVVL¿HG EDVHG
the northwestern part of the deposit. The on their chalcopyrite (Ccp), sphalerite (Sp),
overall magnetite content in the Rockliden ore pyrite (Py), pyrrhotite (Po) and magnetite
is expected to be low compared to the Black (Mgt) content. The following groups were
Mountain ore and might not affect the processing distinguished: Po–Sp, Po–Ccp, Py, Py–Po–Mgt–
RIWKH5RFNOLGHQPDVVLYHVXOSKLGHVVLJQL¿FDQWO\ Ccp–Sp, Py–Ccp–Sp, Py–Sp–Ccp and Py–Sp.
(Bolin, pers. comm., 2013). However, the pyrite The content of pyrite, pyrrhotite and magnetite
content and the silicate mineralogy might have is expected to be important to the ore processing.
DVLJQL¿FDQWLPSDFWLQRUH SURFHVVLQJ HJGXH However, this cannot be directly calculated from
to hardness and grindability characteristics drill core assays. The chalcopyrite and sphalerite
of the ore and therefore controlling the plant content is expected to be approximated from
throughput. chemical assays. However, ore textures cannot

18
be subtracted from chemical assays. Textures and for permission to publish this paper. ALS
which are suggested to be related to deformation Minerals Division (Piteå, Sweden) is thanked
and metamorphism include banding and IRU¿QDQFLDOVXSSRUWIRUSUHSDUDWLRQRIFKHPLFDO
variation of grain size and shape of pyrite on assays and analytical work. Hein Raat (Boliden
PP WR —PVFDOH S\UUKRWLWH¿OOHG IUDFWXUHV Mines, Exploration Department) is thanked for
in pyrite, diablastic intergrowth of pyrrhotite advice about the geology of Rockliden and the
and sphalerite, schistose textures and foliation personnel in the Boliden drill core archive for
of pyrrhotite and chalcopyrite, and magnetite WKHLUKHOSZLWK¿QGLQJDQGPRXQWLQJGULOOFRUHV
with abundant sulphide inclusions. In order to which made it possible to obtain samples in an
evaluate the impact of these textures on the ore HI¿FLHQWZD\
SURFHVVLQJ WKH\ ZRXOG QHHG WR EH TXDQWL¿HG
using (automated) microscopic tools. The authors also acknowledge the comments
Systematic differences in the trace mineralogy, and corrections of three anonymous reviewers
especially the Sb-bearing fraction of the massive and our colleague Riia Chmielowski.
sulphides, could not be determined by qualitative
ore characterisation and the usage of automated
SEM-based mineralogical tools is suggested
IRU WKH TXDQWL¿FDWLRQ RI WKH 6E PLQHUDORJ\
within different rock and ore types. Further,
the content of the Sb-bearing minerals as well
as their association is expected to control their
GLVWULEXWLRQ GXULQJ ÀRWDWLRQ DQG E\ H[WHQW WKH
Sb grade of the Cu–Pb concentrate (cf. Minz et
DO 7KH6EJUDGHRIWKLVÀRWDWLRQSURGXFW
will decide whether it has to be further treated
by hydrometallurgical methods, such as alkaline
sulphide leaching (cf. Awe, 2013), or during
pyrometallurgical processing. In either case
WKHSURGXFWLRQSODQQLQJZRXOGEHQH¿WIURPWKH
forecast of the Sb grade of the Cu–Pb concentrate.
Thus, determining and measuring the rock- and
ore-intrinsic parameters, which control the
distribution of the Sb-bearing minerals, is needed
to develop a model predicting the Sb grade of
the Cu–Pb concentrate (Minz et al., 2013).
SEM-based methods (automated mineralogy)
VKRXOG OHDG WR WKH TXDQWL¿FDWLRQ RI WKHVH
trace minerals. Furthermore, quantifying and
understanding the variation in the Sb mineralogy
in relation to lithological units and structural
features (potentially favouring hydrothermal
PRGL¿FDWLRQ  ZLOO KHOS LQ SUHGLFWLQJ WKH 6E
mineralogy of the ore and plant feed based on a
geological model.

Acknowledgements

7KLV VWXG\ LV SDUW RI D 3K' SURMHFW ¿QDQFHG


by CAMM (Centre of Advanced Mining and
Metallurgy). New Boliden AB is acknowledged
IRU ¿QDQFLQJ WKH DQDO\WLFDO ZRUN LQ WKLV VWXG\

19
References Eldridge, C. S., Barton, Jr P. B. and Ohmoto, H.
1983. Mineral textures and their bearing on
formation of the Kuroko orebodies. Econ.
Allen, R. L., Weihed, P. and Svenson, S. A.
Geol. Monogr., 5, 241–281.
1996. Setting of Zn-Cu-Au-Ag massive
VXO¿GHGHSRVLWVLQWKHHYROXWLRQDQGIDFLHV Evins, P. 2011. New map and structural analysis
architecture of a 1.9 Ga marine volcanic of the Rockliden exposure. Internal Report
arc, Skellefte district, Sweden, Econ. Geol., Boliden Mines, Report No 46346000, 1–23.
91, (6), 1022–1053.
Franklin, J., Lydon, J. and Sangster, D. 1981.
Anderson, R. J. 1940. Microscopic features of 9ROFDQLFDVVRFLDWHG PDVVLYH VXO¿GH
ore from the Sunshine Mine [Idaho]. Econ. deposits. Econ. Geol., 75th Anniversary
Geol., 35, (5), 659–667. Volume, 485–627.
Awe, S. A. 2013. Antimony recovery from Galley, A., Hannington, M. and Jonasson, I.
complex copper concentrates through 2007. Volcanogenic massive sulphide
hydro- and electrometallurgical processes, deposits. Mineral Deposits of Canada: A
Doctoral thesis, Luleå University of Synthesis of Major Deposit-Types, District
Technology. Metallogeny, the Evolution of Geological
Provinces, and Exploration Methods.
Barrie, C. T. and Hannington, M. D. 1999.
Geological Association of Canada, Mineral
&ODVVL¿FDWLRQ RI YROFDQLFDVVRFLDWHG
Deposits Division, Special Publication (5),
PDVVLYHVXO¿GHGHSRVLWVEDVHGRQKRVWURFN
141–161.
composition, Rev. in Econ. Geol., 8, 1–11.
Kousa, J. and Lundqvist, T. 2000. Svecofennian
Bergström, U. 2001. Geochemistry and tectonic
domain, in Description of the bedrock map
setting of volcanic units in the northern
of central Fennoscandia (mid- norden) (ed:
Västerbotten County, northern Sweden.
T. Lundqvist and S. Autio), 47–74, Special
SGU Series, C833, 69–92.
Paper 28 ed. Espoo, Finland: Geological
Bolin, N.-J. 2010. Rockliden - mineralogy Survey of Finland).
investigation of products produced from
Kumpulainen, R. 2009. The Bothnian Basin–its
three different ore samples, Internal
rocks, its age, its origin, FoU-seminarium
Report Boliden Mines, Report No TM_
vid SGU 10–11 mars 2009, 1–45.
REP2010/043, 1–7.
Lamberg, P. 2011. Particles – the bridge
Butcher, A. R. 2010. A practical guide to some
between geology and metallurgy. Preprints
DVSHFWV RI PLQHUDORJ\ WKDW DIIHFW ÀRWDWLRQ
Conference in Minerals Engineering 2011,
in Flotation plant optimisation (ed: C.J.
Luleå, 1–16.
Greet ), 83–93. Australasian Institute of
Mining & Metallurgy. Large, R. R., Gemmell, J. B., Paulick, H. and
Huston, D. L. 2001. The alteration box plot:
Claesson, S. and Lundqvist, T. 1995. Origins
A simple approach to understanding the
and ages of Proterozoic granitoids in the
relationship between alteration mineralogy
Bothnian Basin, central Sweden; isotopic
and lithogeochemistry associated with
and geochemical constraints, Lithos, 36,
YROFDQLFKRVWHG PDVVLYH VXO¿GH GHSRVLWV
(2), 115–140.
Econ. Geol., 96, (5), 957–971.
&ODUN % 5 DQG .HOO\ : &  6XO¿GH
deformation studies; I, experimental Le Maitre, R. W., Bateman, P., Dudek, A., Keller,
deformation of pyrrhotite and sphalerite to J., Lameyre Le Bas, M. J., Sabine, P. A., et.
2,000 bars and 500 degrees C, Econ. Geol., DO $ FODVVL¿FDWLRQ RI LJQHRXV URFNV
68, (3), 332–352. and glossary of terms:recommendation of
the IUGS Subcomission on the Systematics
Depauw, G. 2009. Geology of the Rockliden of Igneous rocks, 193, Cambridge, CUP.
volcanogenic massive sulphide deposit,
north central Sweden, Master thesis, Luleå
University of Technology.

20
Lundqvist, T. 1987. Early Svecofennian Raat, H. and Årebäck, H. 2009. Geology, grade
stratigraphy of southern and central and tonnage estimate of the Rockliden VMS
Norrland, Sweden, and the possible deposit, north central Sweden, Internal
existence of an Archean basement west of Report Boliden Mines, Report No 2009-34,
the Svecokarelides, Precambrian Res., 35, 1–56.
343–352.
Raat, H., Lasskogen, J. and Ronkainen, K. 2011.
Lundqvist, T., Vaasjoki, M. and Persson, P. O. Results Rockliden diamond drilling 2010,
1998. U-Pb ages of plutonic and volcanic Internal Report Boliden Mines, Report No
rocks in the Svecofennian Bothnian Basin, 2011-07, 1–21.
central Sweden, and their implications
for the Palaeoproterozoic evolution of the Roine A. 2009. HSC Chemistry 7.0 User’s
basin, GFF, 120, (4), 357–363. Guide - Chemical Reaction and Equilibrium
Software with Extensive Thermochemical
Lundqvist, T., Gee, D. G., Kumpulainen, R., Database and Flowsheet Simulation
Karis, L. and Kresten, P. 1990. Maps to [computer program]. Version volume 1/2.
the scale of 1:200000. Beskrivning till
berggrundskarta över Västernorrlands län. Rutland, R. W. R., Kero, L., Nilsson, G. and
With english summary. Sveriges Geologiska Stølen, L. K. 2001. Nature of a major
Undersökning Ba 31, 1–429. tectonic discontinuity in the Svecofennian
province of northern Sweden, Precambrian
Mattsson, B. and Heeroma, P. 1985. Rockliden Res., 112, (3–4), 211–237.
prospekteringsresultat 1981-1985, Internal
Report Boliden Mineral AB, Report No GP Schandl, E. S. and Gorton, M. P. 2002.
85 005 2, 1–35. $SSOLFDWLRQRIKLJK¿HOGVWUHQJWKHOHPHQWV
to discriminate tectonic settings in VMS
Mellqvist, C., Öhlander, B., Skiöld, T. and environments, Econ. Geol., 97, (3), 629–
Wikström, A. 1999. The Archaean– 642.
Proterozoic Palaeoboundary in the Luleå
DUHD QRUWKHUQ 6ZHGHQ ¿HOG DQG LVRWRSH Söderlund, U., Elming S.-Å., Ernst, R.E. and
geochemical evidence for a sharp terrane Schissel, D. 2006. The central Scandinavian
boundary, Precambrian Res., 96, (3–4), dolerite Group—Protracted hotspot activity
225–243. or back-arc magmatism?: Constraints from
U–Pb baddeleyite geochronology and Hf
Minz, F., Bolin, N.-J., Lamberg, P. and Wanhainen isotopic data, Precambrian Res., 150, (3),
C. 2013. Detailed characterisation of 136–152.
antimony mineralogy in a geometallurgical
context at the Rockliden ore deposit, north- Wagner, T. and Cook, N. J. 1997. Mineral
central Sweden, Miner. Eng., 52, 95–103. reactions in sulphide systems as indicators
RIHYROYLQJÀXLGJHRFKHPLVWU\DFDVHVWXG\
Minz, F. E. 2013. Mineralogical characterisation from the Apollo mine, Siegerland, FRG,
of the Rockliden antimony-bearing Miner. Mag., 61, (4), 573–590.
volcanic-hosted massive sulphide deposit,
Sweden. Licentiate thesis, Luleå University Walters, S. and Kojovic, T. 2006. Geometallurgical
of Technology. mapping and mine modeling (GEM3) - the
way of the future. International autogenous
Mookherjee, A. and Mishra, B. 1984. ‘Derived’ and semi-autogenous grinding technology
and observed sulphosalt-sulphide phase 2006 Vancouver, Canada, IV-411–IV-425.
assemblages compared—A case study from
Rajpura-Dariba, India. Miner. Depos., 19, Weihed, P. 2004. Overview of the geology and
(2), 112–117. tectonic setting of northern Sweden. Society
of Economic Geologists, Guidebook Series
New Boliden. 2013. New Boliden Annual Report 33, 1–16.
2012, Stockholm, 1–132.
Welin, E. 1987. The depositional evolution of
Petruk, W. 2000. Applied mineralogy in the the Svecofennian supracrustal sequence in
mining industry, Amsterdam, Elsevier Finland and Sweden. Precambrian Res., 35,
Science BV. 95–113.

21
Wen, N., Ashworth, J.R. and Ixer, R.A. 1991.
Evidence for the mechanism of the reaction
producing a bournonite-galena symplectite
from meneghinite. Miner. Mag., 55, 153–
158.
Williams, N. and Holtzhausen, S. 2001. The
impact of ore characterization and blending
on metallurgical plant performance. J. S.
Afr. Inst. Min. Metall., 101, (8), 437–446.
Winchester, J. A. and Floyd, P. A. 1977.
Geochemical discrimination of different
magma series and their differentiation
products using immobile elements. Chem.
Geol., 20, 325–343.
Zaparo Ltd. 2009. Leapfrog Tutorial and
Reference Manual [computer program].
Version 2.

22
PAPER III

Distribution of Sb minerals in the Cu and Zn flotation of Rockliden massive sulphide ore


in north-central Sweden

Friederike E. Minz, Nils-Johan Bolin, Pertti Lamberg, Kai Bachmann, Jens Gutzmer,
Christina Wanhainen

Minerals Engineering (2015), Vol. 82, pp. 125–135.


Minerals Engineering 82 (2015) 125–135

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Distribution of Sb minerals in the Cu and Zn flotation of Rockliden


massive sulphide ore in north-central Sweden
Friederike E. Minz a,⇑, Nils-Johan Bolin b, Pertti Lamberg c, Kai Bachmann d, Jens Gutzmer d,
Christina Wanhainen a
a
Luleå University of Technology, Division of Geosciences and Environmental Engineering, Sweden
b
Boliden AB, Division of Process Technology, Sweden
c
Luleå University of Technology, Division of Minerals and Metallurgical Engineering, Sweden
d
Helmholtz Zentrum Dresden-Rossendorf, Helmholtz-Institute-Freiberg for Resource Technology, Germany

a r t i c l e i n f o a b s t r a c t

Article history: The Rockliden massive sulphide Zn–Cu deposit contains minor amounts of Sb minerals. The Sb mineral-
Received 19 December 2014 ogy is complex in terms of composition, micro textures and mineral associations. The main Sb minerals
Revised 3 March 2015 comprise tetrahedrite, bournonite, gudmundite and Sb–Pb sulphides such as meneghinite. The presence
Accepted 5 March 2015
of these minerals is especially critical to the quality of the Cu–Pb concentrate. To study how they are dis-
Available online 20 March 2015
tributed in a simplified flotation circuit and what controls their process behaviour Sb-rich drill core sam-
ples were selected from the Rockliden deposit and a standard laboratory flotation test was run on the
Keywords:
composite samples. Scanning electron microscope-based automated mineralogy was used to measure
Sulphide ores
Antimony
the Sb mineralogy of the test products, and the particle tracking technique was applied to mass balance
Liberation analysis the different liberation classes to finally trace the distribution of liberated and locked Sb minerals. The
Particle tracking mineralogical factors controlling the distribution of Sb minerals are mineral grain size, the degree of
Froth flotation liberation, and associated minerals. Similarities in the distribution of specific particle types from the
tested composites point towards systematics in the behaviour of particles and predictability of their dis-
tribution which is suggested to be used in a geometallurgical model of the deposit.
Ó 2015 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction Rockliden is a polymetallic copper–zinc (Cu–Zn) massive sul-


phide deposit located in north-central Sweden (Minz et al.,
Generally, antimony (Sb) is known to occur as a trace element 2014). The inferred mineral resource is 9.2 Mt with 4.0 wt% Zn,
in massive sulphide ores. It is regarded as penalty element because 1.8 wt% Cu, 0.4 wt% Pb and 48 g/t Ag (New Boliden, 2014). The
it lowers the quality of Cu concentrates produced from ore and Rockliden ore contains not only bonus elements such as silver
sent to the smelters (Minz, 2013 and references therein). (Ag) but also deleterious trace elements such as arsenic (As) and
Typically, penalty charges are imposed for Sb contents of 0.1– antimony (Sb). At Rockliden, drill core samples with Sb grades
0.3 wt% in copper (Cu) concentrates (Larouche, 2001). However, higher than 0.2 wt% represent about 25% of the length of all anal-
Sb has usage, for example as alloy in lead (Pb) batteries, if it can ysed massive sulphide drill core intersections down to a depth of
be separated from massive sulphide ores (Anderson, 2012; Awe, about 900 m below surface. The abundance of Sb-rich massive sul-
2013). In addition to producing valuable products from massive phide drill core intersections is higher in the upper parts of the
sulphide ores, the Sb content of the tailings ought to be considered deposit (i.e. above 400 m below surface). Particularly in the Zn-
since the Sb content of the tailings is a potential environmental dominated massive sulphides from the upper part of the deposit,
hazard, particularly in the discharge to rivers, comparable to the Sb grade is higher than 0.2 wt% in more than half of the anal-
arsenic (Wilson et al., 2010). ysed intervals.
Flotation tests on Rockliden composite samples have shown
that the threshold of 0.2 wt% Sb is exceeded in the Cu–Pb concen-
⇑ Corresponding author at: Division of Geosciences and Environmental Engineer- trate even if the Sb grade of the blended feed material is below this
ing, Department of Civil, Environmental and Natural Resources Engineering, Luleå value (Bolin, 2010; Minz, 2013). This is related not only to the
University of Technology, SE-971 87 Luleå, Sweden. Tel.: +46 920 491068; mobile:
+46 72 5430958.
flotation behaviour of common Sb minerals but also to the mineral
E-mail address: friederike.minz@ltu.se (F.E. Minz). association and the degree of liberation, which are suspected to

http://dx.doi.org/10.1016/j.mineng.2015.03.007
0892-6875/Ó 2015 The Authors. Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
126 F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135

influence the distribution of Sb minerals between different flota- University of Technology, Sweden. Based on similarities in both
tion products (Lager and Forssberg, 1990; Minz et al., 2013). chemical and SEM-based analyses, three individual drill core sam-
Mineralogical studies have shown that the mineralogy of Sb is ples were selected from each set of drill core samples and com-
complex in the Rockliden massive sulphide ore (Minz et al., bined to form composite samples (Fig. 1b and c). The composite
2013). The main Sb minerals are listed in Table 1, together with samples were ground to a P80 of about 50 lm in a laboratory rod
an expected ideal distribution of liberated particles during Cu–Pb mill before the flotation test (Fig. 1c).
and Zn flotation.
In geometallurgical projects it is common to run ore variability 2.2. Laboratory flotation test
tests in the form of laboratory flotation tests in order to study the
variation in the metallurgical response (Baumgartner et al., 2011; A simplified laboratory flotation test, based on the test com-
Lamberg, 2011). Scanning electron microscope (SEM)-based auto- monly used at the Boliden Mining Company (Fig. 1), was carried
mated mineralogy focusing on minor and trace minerals in ore out. The flotation test was conducted using a Magotteaux bottom
products is frequently done for processed material from noble driven flotation machine with a cell size of 2.7 l. The percentage
metal deposits (e.g. Chetty et al., 2009; Goodall, 2008; Zhou of solids was about 47 wt% for both composites. The flotation
et al., 2009). In contrast to the complex Sb mineralogy of the reagents used are indicated in Fig. 1c. The total flotation time for
Rockliden case, the gold (Au) mineralogy of ore deposits is usually each composite was about 20 min.
simple (e.g. Zhou et al., 2009). Complex trace mineralogy is Five flotation products were collected from each test: Cu–Pb
reported, for example, for uranium (U), silver (Ag), rare earth ele- cleaner concentrate (Cu CC), Cu–Pb cleaner tailing (Cu CT), Zn clea-
ment (REE) and platinum group element (PGE) ores (e.g. Bowell ner concentrate (Zn CC), Zn cleaner tailing (Zn CC) and final tailing
et al., 2011; Chetty et al., 2009; Grammatikopoulos et al., 2013; (Fig. 1c). Their chemical composition was analysed by X-ray fluo-
Kelvin et al., 2011; Rohde et al., 2014). In such cases a large number rescence spectrometry (XRF) at Boliden laboratories. The flotation
of mineral grains, and thus a large number of samples, are mea- products of the composite samples were sieved into three size frac-
sured by automated mineralogy tools (Chetty et al., 2009). tions ( 20 lm, +20 to 45 lm, +45 lm) and the chemical
Instead of running the flotation test for a large number of samples, composition of each fraction was constrained by inductively cou-
a different approach was chosen for the work presented here. This pled plasma atomic emission spectroscopy (ICP-AES) for 21 major
study was constrained to a limited number of samples. To allow for and trace elements at ALS Minerals Division, Piteå, Sweden. From
sound statistical interpretation of the distribution of Sb minerals each size fraction about 10 g sample was separated for preparation
into different products, naturally Sb-enriched samples were col- of polished epoxy grain mounts by the Erzlabor, Helmholtz
lected, in an approach similar to that of Byrne et al. (1995). Two Institute Freiberg for Resource Technology, Germany. Polished
composite samples were selected and compared in this study. epoxy grain mounts of the sieved flotation products were prepared
The two composites represent Zn-dominated massive sulphide by mixing of 3 g aliquots with graphite and epoxy resin; and
ore with a Sb grade above 0.2 wt% from different parts of the finally, the epoxy grain mounts were ground and polished.
Rockliden deposit. Given the complexity of the Sb mineralogy, it
was the primary aim of this study to quantify the Sb mineralogy 2.3. Mineral liberation analysis (MLA) for characterisation of flotation
in the ore and in the products of the flotation test using SEM-based products
automated mineralogical tools. Finally, the target is to define the
factors controlling the distribution of Sb minerals between differ- All flotation products were studied using MLA measurements.
ent products like Cu–Pb, Zn concentrates and tailings. Prior to the SEM studies the samples were carbon-coated with a
Leica EM MED020 carbon coater to avoid surface charging due to
electron beam interactions. Analyses were carried out with a FEI
2. Methods
Quanta 650F equipped with two Bruker Quantax X-Flash 5030
energy-dispersive X-ray spectrometers. MLA Suite 3.1.4 software
2.1. Samples
was used for automated data acquisition and processing at the
Geometallurgy Laboratory, Technische Universität Bergakademie
Two sets of drill core samples were collected, representing Zn-
Freiberg, Germany. Consistent operating conditions were applied
rich massive sulphide ore types from the upper (S-7576) and lower
(Table 2). The grain X-ray mapping mode (GXMAP) was used for
(S-7577) part of the deposit (Fig. 1a and b). Each set contains four
MLA measurement (Fandrich et al., 2007).
to five individual drill core samples of about 1 m length. For each
It should be noted that, although distinguishable by careful
drill core sample about 1 kg quartered drill core was collected.
energy-dispersive X-ray spectroscopy (EDX) spot analysis on a
The drill core samples were crushed and about 100 g sample of
SEM and wavelength-dispersive X-ray spectroscopy (WDS) on an
each sample were separated for chemical analysis (Fig. 1c).
electron probe microanalyser (EPMA) (Minz et al., 2013), menegh-
Further, the Sb mineralogy was studied in polished epoxy grain
inite and boulangerite cannot be distinguished by MLA measure-
mounts prepared from each of the crushed drill core samples.
ments under the chosen conditions (Table 2). For this purpose
Qualitative studies were conducted using a Merlin – Zeiss Gemini
the back scattered electron image can be tuned and expanded to
FESEM (Field Emission Scanning Electron Microscope), Luleå
have a difference between grey levels of meneghinite and boulan-
gerite, similarly as recommended for distinction of hematite and
Table 1 magnetite (Bachmann et al., 2014; Figueroa et al., 2012).
Common Sb minerals at the Rockliden ore deposit. However, this was regarded unnecessary in this study since the
Mineral Formulaa Liberation distributionb flotation behaviour of meneghinite and boulangerite is expected
to be similar due to their similar mineral chemistry (Table 1).
Tetrahedrite (Cu, Fe, Ag, Zn)12.1Sb3.89S13 Cu–Pb concentrate
Bournonite Pb1.01Cu1.03Sb0.94S3 Cu–Pb concentrate
It should also be noted that challenges were encountered dur-
Meneghinite Pb13.02Cu1.10Sb6.81S24 Cu–Pb concentrate ing MLA analyses of the particle size fraction 20 lm including
Boulangerite Pb5.16Sb3.80S11 Cu–Pb concentrate misidentification of minerals along rims of grains and within
Gudmundite Fe1.04Sb0.94S Tailing agglomerations. Such analytical errors might have an influence
a
Formula recalculated based on EPMA/WDS analysis (Minz et al., 2013). on the apparent distribution of minor amounts of meneghinite
b
Liberation distribution according to Lager and Forssberg (1990). and bournonite reported to products other than the Cu–Pb cleaner
F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135 127

Fig. 1. Diagram showing the sample collection and experimental procedure. (a) Location of drill core samples within the Rockliden massive sulphide ore body (translucent
shapes). (b) Ore types as described by Minz et al. (2014) are outlined in the Cu–Zn–Sb ternary plot based on chemical assays (wt%). (c) Laboratory flotation test showing all
sampled flotation products. Abbreviations: Cu CC – Cu–Pb cleaner concentrate, Cu CT – Cu–Pb cleaner tailing, Zn CC – Zn cleaner concentrate, Zn CT – Zn cleaner tailing.

concentrate, but also to base metals with relatively low grade 2.4. Balancing of liberation data
(e.g. Cu in S-7577, Fig. 3a). Several filtering tools were used to min-
imise such errors during the processing of MLA data. Further, rim- In mass balancing of liberation data a particle tracking tech-
ming effects are removed during particle tracking as they were nique was used (Lamberg and Vianna, 2007). The standard proce-
mostly below the particle threshold of 5 vol% (Lamberg and dure of particle tracking comprises three major steps: (1) unsized
Vianna, 2007). mineralogical mass balance, (2) mineral-by-size mass balance
128 F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135

Table 2 for each size fraction and stream. Finally, particle bins were
Summary of measurement parameters at the Mineral Liberation Analyzer (MLA). mass-balanced over the flotation circuit.
SEM parameters MLA parameters
Sample batches >20 lm <20 lm
3. Results
Mode GXMAP Scan speed 16 16
Voltage (kV) 25 Resolution in 1000  1000 1000  1000
pixels (px) 3.1. Chemistry and mineralogy of the composite samples
Working 12 Pixel size (lm/ 0.75 0.75
distance px)
The chemical composition of the composite samples is given in
(mm)
Probe current 10 Acquisition time 6 6
Fig. 3a. The modal mineralogy of the two composites (feed samples
(nA) (ms) S-7576 and S-7577) is presented in Fig. 3b. The mineralogy of the
Spot size 5.88 Minimum 2000 2000 composite samples is largely in agreement with the results from
quartz EDX- the optical microscopic observations and qualitative SEM studies
count
on epoxy grain mounts prepared from crushed drill core samples.
Horizontal field 750 Step size (px) 66 66
width (lm) The main sulphide minerals are pyrite (FeS2) and sphalerite
Brightness 83.5 GXMAP trigger 25 25 (ZnS), followed by pyrrhotite (Fe1 xS), with minor amounts of chal-
Contrast 25.9 Minimum 10 4 copyrite (CuFeS2) and galena (PbS) in both composites (Fig. 3b).
particle size (px) Non-sulphide gangue (NSG) is generally dominated by silicate
BSE calibration Au 252 Minimum grain 4 4
size (px)
minerals and thus will be referred to as silicates. In the feed of both
Particle count >45,000 >190,000 composite samples, quartz is dominating the NSG group, with
minor amounts of plagioclase, muscovite, carbonate, amphibole
and chlorite group minerals. In contrast to Sb as a penalty element,
the mineralogy of arsenic (As) is rather simple: Arsenopyrite is the
and (3) mass balance of narrow liberation classes. Prior to the mass main As-bearing mineral and hosts about 99–100 wt% of the
balancing procedure, element to mineral conversion is utilised to arsenic (As) in the composite samples. The arsenic content of tetra-
attain a mineralogical mass balance that is fully consistent with hedrite is low, generally below 1 wt% (Minz, 2013). Tetrahedrite
chemical assays. The HSC Chemistry software provides tools to carries less than 1 wt% of the arsenic in the composite samples.
adjust the mineral liberation data against the modal mineralogy However, arsenopyrite is a minor mineral in the feed and MLA
received from element to mineral conversion (Roine, 2009). This measurements showed that its distribution pattern in the process
approach was also tested in this study but it resulted in erroneous is similar to pyrite. Thus, it was combined with pyrite. As intended
adjustment of mineral contents to zero in several products. This is by selection of drill core samples, the composite samples are dis-
due to the complexity of the studied system – particularly with tinctively different in their Sb mineralogy. Composite S-7576 is
respect to the Sb minerals (Minz et al., 2013). Consequently, the dominated by tetrahedrite and composite S-7577 is dominated
chemical assays were used only to evaluate and confirm the qual- by meneghinite, in both cases followed by gudmundite (Fig. 3b).
ity of the MLA data. Fig. 2 shows that assays back-calculated from The cumulative mineral grain size distribution of the main sul-
MLA data and chemical assays are generally in good agreement for phide minerals (pyrite and sphalerite) is shown in comparison to
Fe, Zn and Sb. This is also true for other elements such as Cu. the Sb minerals in the composite (Fig. 4a and b). The Sb mineral
Variation of the MLA measurements was studied by repeated grains are relatively small compared to pyrite and sphalerite in
analyses on one epoxy mount. For all base metals the coefficient the flotation feed. Moreover, gudmundite has a large portion of fine
of variation values were below 1% and for most remaining ele- mineral grains, especially in the composite from the lower part of
ments below 10%, showing a good reproducibility of the back-cal- the Rockliden deposit (Fig. 4b). Graphic c and d of Fig. 4 shows the
culated assays based on MLA measurements. The accuracy and cumulative mineral liberation yield in weight present. The Sb
precision of the chemical assays (ICP-AES analysis) is stated with minerals are poorly liberated compared to the main sulphide
±5% for major elements and ±10% for elements at trace element minerals, pyrite and sphalerite. Also, chalcopyrite and galena show
concentration (Fig. 2). It should be noted that elements with high relatively poor liberation, particularly in the composite from the
atomic number, such as Sb and Pb, have higher contents (up to lower deposit part (Fig. 3d).
25%) in MLA derived assays than in chemical assays. As illustrated The mode of occurrence of the two most abundant Sb minerals
for Sb assays (Fig. 2b and d), errors in the MLA measurements or found in the ground feed is given in Fig. 3c. A significant portion
chemical assays are unlikely to account for such deviation. (5–10 wt%) of gudmundite is locked with chalcopyrite and galena
However, it indicates that the abundance of minerals containing in binary particles in the feed samples. This is noteworthy as it
elements of high atomic number is apparently overestimated by might cause gudmundite to report to the Cu–Pb cleaner concen-
MLA analyses on epoxy mounts. This is tentatively attributed to trate and lowering the concentrate quality. Further, the association
minor density separation during preparation of polished epoxy of tetrahedrite and meneghinite with sphalerite, pyrite and sili-
blocks, a common problem in preparing epoxy mounts from pro- cates is considered since such particle types potentially subtract
cess samples (Kwitko-Ribeiro, 2012). tetrahedrite and meneghinite from the Cu–Pb cleaner concentrate.
For both Zn-dominated massive sulphide composites, the Although pyrite and silicates are major mineral classes in both
mineralogical composition of the feed was back calculated from composite samples, the amount of tetrahedrite and meneghinite
the products and the distribution (recovery) of minerals and parti- locked with pyrite and silicates is relatively low and the potential
cles with different liberation classes were calculated with HSC of reporting Sb in tailings through such types of locking thus ought
Chemistry using a particle tracking technique (Lamberg and to be low.
Vianna, 2007). Minerals identified by MLA (EDX) measurement
were grouped to reduce the complexity in modal mineralogy.
Subsequently, particles were binned i.e. they were assigned to par- 3.2. Flotation test
ticle classes (binary, ternary and complex particles) and this basic
grouping was adjusted so that mass-balancing was possible by The mineral grades and recovery by size into the main flotation
reaching a sufficient number of particles in each liberation class products are given in Table 3. In both flotation tests chalcopyrite
F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135 129

Fig. 2. Parity chart for Zn, Fe and Sb assays. Assays, back-calculated from MLA modal composition of the products, are shown on the x-axis and chemical results obtained by
ICP-AES analysis on the y-axis. Sb assays (graphic b and d) show error bars for results from MLA measurements a RSD of 2% and for ICP-AES analysis a RSD for 5% for assays
<1 wt% and a RSD of 10% for assays >1 wt%.

and sphalerite show relatively low recoveries into the Cu–Pb and Thus, entrainment effects are unlikely to cause observed behaviour
Zn cleaner concentrates, 63–77% and 47–60%, respectively. of gudmundite, and most likely gudmundite shows slight
The recovery of fully liberated chalcopyrite is high, 96–100% hydrophobicity in the Cu–Pb flotation. Trace amounts of liberated
(Table 4), indicating that flotation selectivity was very good for meneghinite are noted in the Zn cleaner concentrate, especially
chalcopyrite. Similarly, galena and Pb–Cu–Sb minerals (tetra- in the finest (and partly in the coarse) particle size fraction. This
hedrite, bournonite and meneghinite) show high recoveries into might be due to poorer flotation behaviour of very small mineral
the Cu–Pb cleaner concentrate, 94–98% for galena, 96% for tetra- grains.
hedrite and 98% for meneghinite (Table 4). Gudmundite shows in Fig. 5 shows the mode of occurrence of the main Sb minerals in
both flotation tests quite similar distribution: about half of the lib- the different flotation products, for tetrahedrite and gudmundite
erated gudmundite reported to the final tailing, 13% to the Cu–Pb from the upper part of the Rockliden deposit and for meneghinite
cleaner concentrate, 5–9% to the Zn cleaner concentrate and and gudmundite from the lower part of the deposit. Although
remaining gudmundite to the Cu–Pb and Zn cleaner tailings. tetrahedrite and meneghinite report largely to the Cu–Pb cleaner
Differing from chalcopyrite, sphalerite shows lower selectivity in concentrate, a small amount of these Sb minerals is found in the
flotation. The recovery of fully liberated sphalerite into the Zn clea- Zn cleaner concentrate and in the tailings (Table 3). In the latter
ner concentrate is 58–67%. A significant amount (15–18%) of liber- cases they are found associated with sphalerite, pyrite and silicate
ated sphalerite reports to the Cu–Pb cleaner concentrate. The low minerals. About 30% of gudmundite reports to the Cu–Pb cleaner
recovery of liberated sphalerite is discussed in the following concentrate (Table 3), which is high compared to the liberated gud-
section. mundite of which less than 15% report to the Cu–Pb cleaner con-
The relative amount of liberated gudmundite reporting to the centrate (Table 4). In this case the association with chalcopyrite
Cu–Pb cleaner concentrate was found to be much higher compared or galena apparently have an impact on the distribution of gud-
to the amount of liberated silicates (Table 4). The distribution of mundite (Fig. 5).
liberated gudmundite resembles that of pyrite, particularly in As for the distribution of liberated Sb mineral particles (Fig. 6a),
flotation products of S-7576. An enrichment of liberated gud- the distribution of Sb minerals grains, locked in binary particles, is
mundite, pyrite or silicates cannot be observed in the finest frac- systematic (Fig. 6b). Locking with base-metal minerals is relevant
tion of the Cu–Pb cleaner concentrate of the two composites. to the distribution of gudmundite. Fig. 6b shows that the binaries
130 F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135

Fig. 3. Chemical and mineralogical characteristics of the feed back-calculated from MLA data: (a) Chemical composition of composite drill core samples. The mineralogical
distribution of Sb between the Sb minerals (tetrahedrite, bournonite, meneghinite & gudmundite) is derived from modal mineralogy and the Sb content of the corresponding
Sb minerals (Minz et al., 2013). (b) Modal mineralogy and Sb mineralogy of the flotation feed calculated from MLA results of the flotation products. (c) Mode of occurrence of
Sb minerals in the flotation feed (S-7576 and S-7577) and distribution of particle types in different size fractions (%); for liberated Sb minerals the relative standard deviation
is given in brackets. Abbreviations: B Ccp – binary with chalcopyrite, B Gn – binary with galena, Lib – fully liberated, MC – more complex particles, NSG – binary with non-
sulphide gangue (silicates), B Sp – binary with sphalerite, B⁄ – other binaries (with pyrrhotite and remaining Sb minerals), Bour – bournonite, Ccp – chalcopyrite, Gd –
gudmundite, Gn – galena, Mene – meneghinite (with boulangerite), NSG – non-sulphide gangue (for explanation see text), Po – pyrrhotite, Py – pyrite (with arsenopyrite), Sp
– sphalerite, Ttr – tetrahedrite.

gudmundite–chalcopyrite of the composite sample S-7576 and the 4. Discussion


binaries gudmundite–galena from composite sample S-7577
mostly report to the Cu–Pb cleaner concentrate (i.e. to about As expected, almost all liberated main sulphide minerals report
75%). It should be noted that the chalcopyrite content is higher selectively to the respective flotation products. Sphalerite is a nota-
in the feed of composite sample S-7576 than in S-7577 (Fig. 3b). ble exception. The relative large amount of liberated sphalerite in
Thus, it is reasonable that the higher amount of chalcopyrite in the Cu–Pb cleaner concentrate may be related to activation of
the feed of S-7576 will in turn also increase the possibility of lock- sphalerite during Cu–Pb flotation (cf. Lastra, 2007; Chandra and
ing of gudmundite with chalcopyrite. Binary tetrahedrite–pyrite Gerson, 2009). It could also be due to stereological effects; parti-
and meneghinite–silicate particles are found in relatively low cles, which look like fully liberated sphalerite in 2D sections, may
amounts in the feed (Figs. 3c and 6b). The binaries with sphalerite indeed still be intergrown with other minerals. It is known that
report largely to the Cu–Pb cleaner concentrate (85 wt%, without stereological correction the liberation of minerals is over-
Fig. 6XXb). estimated (e.g. Spencer and Sutherland, 2000) and preliminary
F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135 131

Fig. 4. Mineral grain size distribution in the bulk feed, back-calculated from MLA measurements: (a) S-7576, (b) S-7577. Mineral liberation by weight for the feed: (c) S-7576,
(d) S-7577. Abbreviations: Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene – meneghinite, Py – pyrite (with arsenopyrite), Sp – sphalerite, Ttr – tetrahedrite.

Table 3
Distribution (recovery) of minerals into different flotation products, for composite S-7576 and S-7577. Mineral abbreviations: Ccp – chalcopyrite, Gd – gudmundite, Gn – galena,
Mene – meneghinite (with boulangerite), Py – pyrite (with arsenopyrite), NSG – non-sulphide gangue (silicates), Sp – sphalerite, Ttr – tetrahedrite. Cleaner tailings are not shown.
For abbreviation of product names see Fig. 1.

Composite S-7576 S-7577


Mineral NSG (%) Py (%) Ccp (%) Gn (%) Sp (%) Ttr (%) Gd (%) NSG (%) Py (%) Ccp (%) Gn (%) Sp (%) Mene (%) Gd (%)
Cu CC bulk 5.1 10.9 77.6 85.3 27.7 96.2 28.5 5.2 7.8 63.1 86.2 20.3 94.6 28.3
Cu CC 20 lm 1.7 1.7 37.3 54.8 10.1 44.9 9.4 2.1 2.0 37.1 58.6 11.2 58.3 12.7
Cu CC +20 lm 1.1 2.4 21.4 16.2 8.0 27.3 8.0 1.2 1.6 12.3 11.5 4.4 14.7 5.5
Cu CC +45 lm 2.2 6.8 18.8 14.3 9.6 23.9 11.1 1.8 4.2 13.7 16.1 4.7 21.6 10.1
Zn CC bulk 2.8 6.1 1.3 2.6 46.8 0.4 8.5 4.5 4.3 7.8 4.1 59.7 1.1 6.8
Zn CC 20 lm 0.4 0.7 0.4 0.4 21.8 0.1 4.5 0.5 0.8 2.0 0.7 25.2 0.4 3.1
Zn CC +20 lm 0.6 1.1 0.3 0.5 12.1 0.1 1.8 0.8 0.9 0.7 0.6 12.2 0.2 1.2
Zn CC +45 lm 1.8 4.4 0.6 1.7 12.9 0.2 2.2 3.2 2.7 5.1 2.7 22.3 0.6 2.4
Tailing bulk 83.6 66.8 8.2 1.9 1.0 0.2 32.7 81.6 71.5 16.4 2.9 3.4 1.4 35.4
Tailing 20 lm 27.5 21.3 4.7 0.2 0.6 0.1 22.0 22.2 23.5 7.7 0.3 1.9 0.4 23.2
Tailing +20 lm 18.7 12.3 0.9 0.3 0.1 0.0 4.7 8.6 13.9 2.9 0.3 0.2 0.1 6.6
Tailing +45 lm 37.3 33.2 2.6 1.4 0.3 0.1 6.0 50.8 34.1 5.8 2.2 1.3 0.9 5.6

results on the particle behaviour have shown that sphalerite-domi- composite S-7577, lower part of the Rockliden deposit) report
nated particles with small amounts of Cu–Pb bearing minerals will mostly to the Cu–Pb cleaner concentrate, as expected (Lager and
report to the Cu–Pb cleaner concentrate. However, stereological Forssberg, 1990). Liberated tetrahedrite and meneghinite report
correction was not applied because the multiphase system of almost exclusively to the Cu–Pb cleaner concentrates. The amount
Rockliden is complex compared to the two-phase-systems for of liberated gudmundite reported to the Cu–Pb cleaner concentrate
which stereological corrections have been developed (e.g. King is relatively high compared to the study by Lager and Forssberg
and Schneider, 1998). Stereology is, however, not likely the main (1990), who found only about 5% of this mineral in Zn and Cu–Pb
reason since pyrite, the other main sulphide mineral, does show cleaner concentrates. In this study, about 10% of pyrite and gud-
much lower recoveries into the Cu–Pb cleaner concentrate mundite report to the Cu–Pb cleaner concentrate. This cannot be
(Table 4). related to entrainment effects, as no distinct enrichment in the fin-
Generally, the main Sb minerals, tetrahedrite (of composite S- est fraction of the Cu–Pb cleaner concentrate was noticed for the
7576, upper part of the Rockliden deposit) and meneghinite (of distribution of liberated gudmundite in this study (cf. Kursun,
132 F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135

Table 4
Distribution (recovery) of fully liberated particles – particles consisting of only one mineral – into different flotation products, for composite S-7576 and S-7577. A standard
deviation (STD) is given for recoveries of minerals in the bulk of each product; the STD was based on the relative standard deviation = 100/(NOP)0.5 (NOP – number of particles of a
specific particle type in a flotation feed and product) and estimated by using bootstrap Monte Carlo simulation. Mineral abbreviations: Ccp – chalcopyrite, Gd – gudmundite, Gn –
galena, Mene – meneghinite (with boulangerite), Py – pyrite (with arsenopyrite), NSG – non-sulphide gangue (silicates), Sp – sphalerite, Ttr – tetrahedrite. Cleaner tails are not
shown. For abbreviation of product names see Fig. 1.

Composite S-7576 S-7577


Mineral NSG (%) Py (%) Ccp (%) Gn (%) Sp (%) Ttr (%) Gd (%) NSG (%) Py (%) Ccp (%) Gn (%) Sp (%) Mene (%) Gd (%)
Cu CC bulk 2.2 13.3 96.0 93.7 17.5 96.2 13.2 2.8 5.9 99.7 97.5 14.7 97.8 12.8
Cu CC – STD 0.0 0.0 0.2 0.3 0.0 1.5 2.2 0.0 0.0 1.9 0.4 0.0 2.3 2.5
Cu CC 20 lm 0.9 4.0 57.5 71.6 8.1 55.5 8.1 1.6 2.9 73.3 83.4 9.9 77.4 9.9
Cu CC +20 lm 0.5 3.7 26.1 15.6 5.2 27.5 3.3 0.7 1.5 17.9 8.7 3.1 11.9 1.7
Cu CC +45 lm 0.8 5.6 12.4 6.4 4.1 13.2 1.8 0.5 1.5 8.5 5.4 1.8 8.5 1.2
Zn CC bulk 1.3 3.4 0.0 0.4 58.1 0.2 9.2 1.4 0.9 0.0 0.6 66.9 0.6 4.5
Zn CC – STD 0.0 0.0 0.7 0.3 0.0 3.1 1.7 0.0 0.0 4.6 0.5 0.0 3.7 5.3
Zn CC 20 lm 0.3 1.0 0.0 0.2 29.2 0.2 6.3 0.3 0.4 0.0 0.4 31.9 0.5 3.5
Zn CC +20 lm 0.3 0.6 0.0 0.1 15.5 0.0 1.7 0.4 0.4 0.0 0.1 14.3 0.0 0.6
Zn CC +45 lm 0.7 1.9 0.0 0.1 13.4 0.0 1.1 0.8 0.1 0.0 0.1 20.7 0.1 0.4
Tailing bulk 90.4 60.1 0.0 0.1 0.7 0.0 46.3 89.0 81.2 0.2 0.2 2.4 0.3 52.1
Tailing – STD 0.0 0.0 3.8 2.8 0.1 29.0 6.4 0.0 0.0 12.9 3.2 0.0 19.7 11.3
Tailing 20 lm 30.8 34.8 0.0 0.1 0.6 0.0 34.9 28.5 48.6 0.1 0.1 2.0 0.2 40.3
Tailing +20 lm 21.6 15.8 0.0 0.0 0.0 0.0 7.4 10.3 12.7 0.0 0.0 0.1 0.0 9.5
Tailing +45 lm 37.9 9.5 0.0 0.0 0.0 0.0 4.0 50.2 19.9 0.1 0.1 0.3 0.1 2.3

Fig. 5. Mode of occurrence of Sb minerals in different products of the two composites compared with the feed samples (S-7576 and S-7577, cf. Fig. 3c). Abbreviations: B Ccp –
binary with chalcopyrite, B Gn – binary with galena, B NSG – binary with non-sulphide gangue (silicates), B Py – binary with pyrite (with arsenopyrite), B Sp – binary with
sphalerite, Gd – gudmundite, Mene – meneghinite (with boulangerite), Ttr – tetrahedrite, LR – lower deposit part (composite S-7576), UR – upper deposit part (S-7577). For
abbreviation of product names see Fig. 1.

2014; Trahar and Warren 1976; Welsby et al., 2010). Liberated and Forssberg (1990) noticed that the floating of gudmundite in
gudmundite shows behaviour similar to the behaviour of liberated the lime environment (Zn flotation) is slower than in the soda
pyrite. This might be related to similarities in their mineral chem- environment (Cu–Pb flotation). This seems to be supported by
istry. It is assumed that some of the liberated gudmundite and pyr- the findings of this study, where 4.5–9% liberated gudmundite
ite reported to the Cu–Pb cleaner concentrate could have been report to the Zn cleaner concentrate whereas 13% report to the
removed by the use of additional cleaning stages. Further, Lager Cu–Pb cleaner concentrate (Table 4).
F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135 133

Fig. 6. (a) Distribution (%) of liberated Sb minerals into different flotation products and the percentage of liberated Sb minerals in the feed (wt%); in comparison with (b) the
distribution (%) of binary locked Sb minerals into different flotation products and the percentage of liberated Sb minerals in the feed (wt%). Abbreviations: Cu CC – Cu–Pb
cleaner concentrate, Cu CT – Cu–Pb cleaner tailing, Zn CC – Zn cleaner concentrate, Zn CT – Zn cleaner tailing, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene –
meneghinite, NSG –non-sulphide gangue (silicates), Sp – sphalerite, Ttr – tetrahedrite, LR – lower part of the Rockliden ore deposit, UR – upper part of the Rockliden ore
deposit.

The amount of tetrahedrite and meneghinite locked with spha- Misplacement of Sb minerals into valuable flotation products,
lerite was found to be negligible in the Zn cleaner concentrate, but due to locking with base metal sulphide minerals, might not be
relatively high in the Cu–Pb cleaner concentrate (Fig. 6b). Initial adjustable during flotation. Finer grinding could be considered,
studies on particle behaviour have shown that even small amounts but this has cost implications and overgrinding might even have
of Cu- and Pb-bearing minerals, such as tetrahedrite and menegh- negative effects on the flotation of the main base metal sulphide
inite, locked with sphalerite are likely to cause reporting of such minerals (Trahar, 1981; Trahar and Warren, 1976). However, it is
particles into the Cu–Pb cleaner concentrate. Locking of gud- expected that all main Sb minerals from Rockliden can be removed
mundite with chalcopyrite as well as with galena will cause report- from the Cu–Pb concentrate by alkaline sulphide leaching before
ing of gudmundite into the Cu–Pb cleaner concentrate. The types of the product is sent to the smelter (Awe, 2013; Minz, 2013 and
locking of gudmundite – i.e. dominance of chalcopyrite or galena – references therein). Thus, a forecast of the Sb grade of the Cu–Pb
might be related back to the differences in the chalcopyrite and concentrate might help to evaluate an alternative treatment of
galena content of the feed sample. Moreover, it is known that gud- the Cu–Pb concentrate.
mundite can form very fine intergrowths with chalcopyrite, pyr- So far, research was conducted only on selected Sb-rich Zn-
rhotite and galena (Minz et al., 2013), and such fine intergrowths dominated massive sulphide samples from the Rockliden deposit
are found typically in the Cu–Pb cleaner concentrate of composite subjected to a simplified flotation test. These samples represent
S-7577. These intergrowths are too fine (<1 lm) to be correctly two massive sulphide types outlined in qualitative studies on the
resolved in the processed MLA data. However, textural differences Sb mineralogy of the Rockliden mineralisation (Minz et al., 2013).
in the association of gudmundite of S-7576 and S-7577 are sus- Additionally, Sb-rich ore types were described by qualitative ore
pected to have no or little effect (Fig. 6b). characterisation of Rockliden (Minz et al., 2014). Testing of
134 F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135

different ore types might allow linking the behaviour of Sb miner- 5. Conclusions
als in the flotation test with the 3D distribution of ore types in a
geological model. A geological model of the deposit would then The behaviour of Sb minerals from the Rockliden massive sul-
allow to evaluate the impact of Sb-rich ore types on the processing phide deposit was critically assessed based on the study of two
of the Rockliden mineralisation, but such model has been found composite samples using a simplified flotation test. Although the
difficult to implement (Minz et al., 2014). Further, ore-intrinsic main mineralogy of both composites was found similar, the Sb
parameters, i.e. proxies measuring, for example, the recovery, can- mineralogy of the two composite samples is different. The flotation
not be assumed to be additive during blending (e.g. Deutsch, 2013; feed from the upper and lower part of the deposit is dominated by
Dunham and Vann, 2007). Thus, a simple summation of proxies for tetrahedrite and meneghinite, respectively. Additionally to tetra-
the flotation behaviour of different rock and ore units is not hedrite and meneghinite, significant amounts of gudmundite were
expected to predict characteristics of a mining block adequately. found in both composites.
Effects of waste rock dilution on massive sulphides, i.e. flotation Liberated chalcopyrite and galena as well as tetrahedrite and
test of composite samples, for building a 3D process-adopted meneghinite report selectively in the Cu–Pb concentrate of both
model should be assessed separately. tested composites. Liberated gudmundite has a poor selectivity
An earlier laboratory flotation test has been done on composite into the tailings and a weak hydrophobicity both in Cu–Pb and
samples representing different ore lenses, with similar grade as the Zn flotation. The same is valid for pyrite, but the degree of
Rockliden resource, and products were studied with EDX on a SEM hydrophobicity is lower. About 15% liberated sphalerite reports
(Bolin, 2010). The semi-quantitative character of the results of this into Cu–Pb cleaner concentrate indicating slight sphalerite activa-
earlier study does not allow for evaluation and comparison of the tion in Cu–Pb flotation. Gangue minerals show almost no flotation
distribution of Sb minerals with the herein presented results. and effect of entrainment is minimal.
However, it can be assumed that Sb minerals will be equally dis- When Sb minerals occur locked with other minerals they show
tributed even when derived from samples with lower Sb grade. systematic behaviour controlled by the minerals present and their
Similarly, the floatability component method assumes the floata- mass proportions in the particle. Locking effects with, for example,
bility of particles will be constant as long as the flotation condi- sphalerite, had negligible effects on the subtraction of tetrahedrite
tions are the same, i.e. no regrinding, no addition of chemicals and meneghinite from the Cu–Pb cleaner concentrate in the flota-
etc. (Runge et al., 1997). This assumption would need further tion test. However, locking of gudmundite with chalcopyrite and
research. Welsby et al. (2010) pointed out that even floatability galena causes imperfect reporting of up to 10% of gudmundite into
component models have been unable to model satisfactory flota- the Cu–Pb cleaner concentrate.
tion circuits when, e.g. feed grade and particle size distribution Although the two tested samples show differences in the Sb
are changed. The reader should be aware that based on these mineralogy and in the bulk flotation results, initial results show
results of this study no conclusion should be drawn on the varia- that particles with similar size and composition behave in a similar
tion of the Sb mineralogy within Rockliden mineralisation nor way in the used laboratory flotation test. This encourages building
should be speculated whether the particle behaviour will be iden- of a Sb distribution model using systematics in the behaviour of
tical in composite samples representing full-scale mining blocks. particles based on particle properties. Such model should then give
As an alternative to the implementation through ore types, the a possibility to forecast the metallurgical behaviour of the
implementation of results from this study into a resource model mineralogical complex Rockliden ore accurately if the liberation
might be achieved through distribution models of Sb-bearing par- distribution of the flotation feed is available. Further, a Sb dis-
ticles. Primary results on the distribution of Sb minerals have tribution model might help to evaluate treatment options, such
shown to be similar between the tested composites. Thus, the as alkaline sulphide leaching, when the Sb content is predicted to
impact of locked Sb minerals reporting to, or rejected from, the lower the quality of the Cu–Pb cleaner concentrate.
Cu–Pb cleaner concentrate might be evaluated on the basis of par-
ticle behaviour. Using the information obtained on the particle Acknowledgements
behaviour, a metallurgical model for the distribution of Sb miner-
als in different associations and liberation classes could be imple- New Boliden is acknowledged for financing the analytical work
mented in a resource model. The modelling could be done in this study. Travelling costs for MLA studies were partly covered
following the guidelines given by Lund (2013). However, this by ProMinNET (Nordic network of Process Mineralogy and
requires firstly providing the modal composition of each block, sec- Geometallurgy funded by Nordforsk). The Helmholtz Institute
ondly information on mineral association and thirdly a model on Freiberg is acknowledged for funding of MLA analyses presented
liberation distribution of flotation feed with certain particle size in this study. This study is part of a PhD project financed by
distribution. This is labour and time intensive and therefore the CAMM (Centre of Advanced Mining and Metallurgy).
suitability of proxies should be explored. Semi-quantitative studies Staff members of the technical laboratory at the Boliden Mining
relating the Sb mineralogy of drill core samples with chemical Company, of the Erzlabor at the Helmholtz Institute Freiberg and
analysis indicated challenges in using chemical assays to predict Geometallurgy laboratory at the Technische Universität
the Sb mineralogy. However, sequential leaching methods could Bergakademie Freiberg are thanked for their support in sample col-
be developed. Gudmundite is not a sulphosalt whereas the remain- lection and assistance in experimental and analytical work.
ing Cu- and Pb-bearing Sb minerals, i.e. tetrahedrite, bournonite Riia Chmielowski is acknowledged for her comments improving
and meneghinite, are sulphosalts with an oxidation state of +3 of this manuscript.
the contained Sb (Moëlo et al., 2008). The difference in the oxida-
tion state of Sb in these two groups, i.e. gudmundite and the
sulphosalts, would allow to consider selective dissolution methods,
References
e.g. the bromine-methanol method of Penttinen et al. (1977), to
estimate the gudmundite content of a drill core interval. As all Anderson, C.G., 2012. The metallurgy of antimony. Chem. Erde. 72, 3–8.
sulphosalts at the Rockliden deposit have shown to have high Awe, S.A., 2013. Antimony recovery from complex copper concentrates through
hydro- and electrometallurgical processes. PhD Thesis, Luleå Tekniska
recovery to the Cu–Pb cleaner concentrate, generally more than
Universitet, Sweden.
90%, the Sb grade of this product could then be estimated by mod- Bachmann, K., Bartzsch, A., Gutzmer, J., 2014. Discrimination of hematite and
elling the addition of Sb through gudmundite to this product. magnetite in finely intergrown natural iron ores by automated mineralogy. In:
F.E. Minz et al. / Minerals Engineering 82 (2015) 125–135 135

EMAS 2014 11th regional workshop on electron probe microanalysis of Lamberg, P., Vianna, S.M.S., 2007. A technique for tracking multiphase mineral
materials today, Leoben. 290. particles in flotation circuits. In: VII meeting of the southern hemisphere on
Baumgartner, R., Dusci, M., Gressier, J., Trueman, A., Poos, S., Brittan, M., Mayta, P., mineral technology Ouro Preto, Minas Gerais, Brazil, pp. 195–203.
2011. Building a geometallurgical model for early-stage project development –A Larouche, P., 2001. Minor elements in copper smelting and electrorefining. Master
case study from the Canahuire epithermal Au–Cu–Ag deposit, southern Peru. Thesis, Mining and Metallurgical Engineering, McGill University, Montreal,
Conference proceedings of the first AusIMM international geometallurgy Canada.
conference; 5-7 September 2011; Carlton, VIC, Australia: The AusIMM, pp. 53– Lastra, R., 2007. Seven practical application cases of liberation analysis. Int. J. Miner.
59. Process. 84 (1), 337–347.
Bolin, N.J., 2010. Rockliden – mineralogy investigation of products produced from Lund. C., 2013. Mineralogical, chemical and textural characterisation of the
three different ore samples, Internal report Boliden Mineral AB. TM_REP2010/ Malmberget iron ore deposit for a geometallurgical model. PhD Thesis, Luleå
0437. Tekniska Universitet, Sweden.
Bowell, R.J., Grogan, J., Hutton-Ashkenny, M., Brough, C., Penman, K., Sapsford, D.J., Minz, F., Bolin, N.-J., Lamberg, P., Wanhainen, C., 2013. Detailed characterisation of
2011. Geometallurgy of uranium deposits. Miner. Eng. 24 (12), 1305–1313. Byrne, antimony mineralogy in a geometallurgical context at the Rockliden ore
M., Grano, S., Ralston, J., Franco, A., 1995. Process development for the deposit, North-Central Sweden. Miner. Eng. 52, 95–103.
separation of tetrahedrite from chalcopyrite in the Neves-Corvo ore of Somincor Minz, F.E., 2013. Mineralogical characterisation of the Rockliden antimony-bearing
S.A., Portugal. Miner. Eng. 8 (12), 1571–1581. volcanic-hosted massive sulphide deposit, Sweden. Licentiate thesis, Luleå
Chandra, A., Gerson, A., 2009. A review of the fundamental studies of the copper Tekniska Universitet.
activation mechanisms for selective flotation of the sulfide minerals, sphalerite Minz, F.E., Lasskogen, J., Wanhainen, C., Lamberg, P., 2014. Lithology and
and pyrite. Adv. Colloid. Interface. Sci. 145 (1), 97–110. mineralisation types of the Rockliden Zn–Cu massive sulphide deposit, north-
Chetty, D., Gryffenberg, L., Lekgetho, T., Molebale, I., 2009. Automated SEM study of central Sweden: Implications for ore processing. Trans. Inst. Min. Metall. (Sect.
PGM distribution across a UG2 flotation concentrate bank: implications for B: Appl. Earth Sci.) 123 (1), 2–17.
understanding PGM floatability. J. South. Afr. Inst. Min. Metall. 109, 587–593. Moëlo, Y., Makovicky, E., Mozgova, N.N., Jambor, J.L., Cook, N., Pring, A., Paar, W.,
Deutsch, C.V., 2013. Geostatistical modelling of geometallurgical variables – Nickel, E.H., Graeser, S., Karup-Møller, S., 2008. Sulfosalt systematics: a review
problems and solution. The second AusIMM international geometallurgy report of the sulfosalt sub-committee of the IMA commission on ore
conference; 30 September – 2 October 2013; Brisbane, Australia: The mineralogy. Eur. J. Miner. 20 (1), 7–46.
AusIMM, pp. 7–15. New Boliden Annual Report 2013, 2014. New Boliden, Stockholm, 128 p.
Dunham, S., Vann, J., 2007. Geometallurgy, geostatistics and project value – does Penttinen, U., Palosaari, V., Siura, T., 1977. Selective dissolution and determination
your block model tell you what you need to know? Project evaluation of sulfides in nickel ores by the bromine–methanol method. Bull. Geol. Soc. Fin.
conference, Melbourne, Victoria, pp. 189–196. 49 (2), 79–84.
Fandrich, R., Gu, Y., Burrows, D., Moeller, K., 2007. Modern SEM-based mineral Rohde, M., Guresin, N., Johnson N.W., 2014. Characterisation of silver minerals in
liberation analysis. Int. J. Miner Process. 84 (1), 310–320. lead-zinc flotation tailings and their resposne to cyanidation. In: 12th AusIMM
Figueroa, G., Moeller, K., Buholt, M., Gloy, G., Haberlah, D., 2012. Advanced mill operators’ conference 2014, The AusIMM, pp. 163–172.
discrimination of hematite and magnetite by automated mineralogy. In: Roine, A., 2009. HSC Chemistry 7.0 User’s Guide – Chemical Reaction and
Proceedings of the 10th International Congress for Applied Mineralogy Equilibrium Software with Extensive Thermochemical Database and
(ICAM), pp. 197–204. Flowsheet Simulation. Volume 1/2.
Goodall, W.R., 2008. Characterisation of mineralogy and gold deportment for Runge, K.C., Harris, M.C., Frew, J.A., Manlapig, E.V., 1997. Floatability of streams
complex tailings deposits using QEMSCANÒ. Miner. Eng. 21 (6), 518–523. around the Cominco Red Dog lead cleaning circuit. In: Proceedings of the Sixth
Grammatikopoulos, T., Mercer, W., Gunning, C., 2013. Mineralogical Mill Operators Conference, AusIMM, Madang, Papua New Guinea, pp. 157–163.
characterisation using QEMSCAN of the Nechalacho heavy rare earth metal Spencer, S., Sutherland, D., 2000. Stereological correction of mineral liberation grade
deposit, Northwest Territories. Canada Can. Metall. Quart. 52 (3), 265–277. Kelvin, distributions estimated by single sectioning of particles. Image Anal. Stereol. 19,
M.A., Sylvester, P.J., Cabri, L.J., 2011. Mineralogy of rare occurrences of 175–182.
precious-metal-enriched massive sulfide in the Voiey’s bay Ni–Cu–Co Ovoid Trahar, W., 1981. A rational interpretation of the role of particle size in flotation. Int.
deposit, Labrador. Canada Can. Mineral. 49 (6), 1505–1522. J. Miner. Process. 8 (4), 289–327.
King, R.P., Schneider, C.L., 1998. Stereological correction of linear grade distributions Trahar, W.J., Warren, L.J., 1976. The flotability of very fine particles — a review. Int. J.
for mineral liberation. Powder Technol. 98 (1), 21–37. Miner. Process. 3 (2), 103–131.
Kursun, H., 2014. Effect of fine particles’ entrainment on conventional and column Welsby, S., Vianna, S., Franzidis, J., 2010. Assigning physical significance to
flotation. Particul. Sci. Technol. 32 (3), 251–256. floatability components. Int. J. Miner. Process. 97 (1), 59–67.
Kwitko-Ribeiro, R., 2012. New sample preparation developments to minimize Wilson, S.C., Lockwood, P.V., Ashley, P.M., Tighe, M., 2010. The chemistry and
mineral segregation in process mineralogy. In: Proceedings of the 10th behaviour of antimony in the soil environment with comparisons to arsenic: a
international congress for applied mineralogy (ICAM). Springer, pp. 411–417. critical review. Environ. Pollut. 158 (5), 1169–1181.
Lager, T., Forssberg, K.S.E., 1990. Separation of antimony mineral impurities from Zhou, J., Mermillod-Blondin, R., Cousin, P., 2009. Deportment of gold and silver in
complex sulphide ores. Trans. Inst. Min. Metall. (Sect. C: Miner. Process. Extr. the Pinos Altos Composite and leach residues: Implications for recovery
Metall.) 99 (C54), 54–61. improvement. In: World gold conference 2009, SAIMM, pp. 75–84.
Lamberg, P., 2011. Particles – the bridge between geology and metallurgy. In:
Preprints Conference in Minerals Engineering 2011. 16 p.
PAPER IV

Particle-based Sb distribution model for Cu–Pb flotation as part of geometallurgical


modelling at the polymetallic Rockliden deposit, north-central Sweden

Friederike E. Minz, Nils-Johan Bolin, Pertti Lamberg, Christina Wanhainen, Kai Bachmann,
Jens Gutzmer

manuscript to be submitted (to: Mineral Processing and Extractive Metallurgy: Transactions


of the Institutions of Mining and Metallurgy: Section C).
3DUWLFOHEDVHG6EGLVWULEXWLRQPRGHOIRU&X±3EÀRWDWLRQDVSDUWRI
JHRPHWDOOXUJLFDOPRGHOOLQJDWWKHSRO\PHWDOOLF5RFNOLGHQGHSRVLW
north-central Sweden

Friederike E. Minz1, Nils-Johan Bolin2, Pertti Lamberg3, Christina Wanhainen1,


Kai Bachmann4, Jens Gutzmer4

1
Luleå University of Technology, Division of Geosciences and Environmental Engineering, Sweden
2
Boliden Mineral AB, Division of Process Technology, Sweden
3
Luleå University of Technology, Division of Minerals and Metallurgical Engineering, Sweden
4
Helmholtz Zentrum Dresden-Rossendorf, Helmholtz-Institute-Freiberg for Resource Technology, Germany

Abstract

The polymetallic Cu–Zn ore of the Rockliden massive sulphide deposit in the Skellefte District
in north-central Sweden contains a number of deleterious elements in relevant concentrations. Of
particular concern is the amount of antinomy (Sb) reporting to the Cu–Pb concentrate. The aim
of this study is to compare different model options to simulate the distribution of Sb minerals in a
ODERUDWRU\ÀRWDWLRQWHVWEDVHGRQGLIIHUHQWGHJUHHRIGHWDLOVLQWKHPLQHUDORJLFDOLQIRUPDWLRQRIWKH
ÀRWDWLRQIHHG([SHULPHQWDOGDWDREWDLQHGIURPIRXUFRPSRVLWHVFROOHFWHGIURPGULOOFRUHVZDVXVHG
IRUWKHPRGHOOLQJDQGVLPXODWLRQ$OOÀRWDWLRQPRGHOVDUHGHVFULEHGE\¿UVWRUGHUNLQHWLFSURFHVVDQG
ÀRWDWLRQUDWHFRQVWDQWV NYDOXHV GHULYHGIURPPLQHUDOUHFRYHULHV0LQHUDOVSHFL¿FNYDOXHVZHUH
selected from two composites. The following different simulation levels were run (sorted from least to
highest level of detail of their mineralogical information): chemical assays, unsized bulk mineralogy,
sized bulk mineralogy and particle information. For the latter, k values were modelled from recovery
of liberated particles. It was shown that recoveries simulated based on bulk mineralogy are mostly
ZLWKLQWKHHUURUPDUJLQDFFHSWDEOHLQWKHH[SORUDWLRQVWDJHRIWKH5RFNOLGHQGHSRVLW8QH[SHFWHG
high deviation in the simulation using particle information from the original recovery has been partly
attributed to the fact that recovery of non-liberated particles cannot be modelled appropriately in the
SUHVHQWYHUVLRQRIWKHPRGHOOLQJDQGVLPXODWLRQVRIWZDUH,WLVH[SHFWHGWKDWWKHLPSOHPHQWDWLRQRI
full particle information in simulation will improve the Sb distribution model for the mineralogically
FRPSOH[5RFNOLGHQGHSRVLW

.H\ZRUGV)ORWDWLRQNLQHWLFVPLQHUDOÁRDWDELOLW\PRGHOOLQJVXOSKLGHRUHVDQWLPRQ\

1
1. Introduction 7KHUDWHFRQVWDQWLVH[SUHVVHGDV

An important aim in geometallurgy is to forecast (2) k = kc . Rf = P . Sb . Rf


the concentrate grades and recoveries of target
minerals in processing products based on the with: k – overall rate constant (s-1), kc – collection
information available for the ore. This might be zone rate constant (s-1  3 ± RUH ÀRDWDELOLW\
done through quantifying relevant parameters (dimensionless); Sb ± EXEEOH VXUIDFH DUHD ÀX[
or indicators in the ore (e.g. Chauhan et al. (s-1), Rf – froth recovery factor (e.g. Gorain et
2013). Several studies have been conducted al., 1998).
demonstrating the impact of ore characteristics
on the processing of base metal and precious Several input factors must thus be considered
metal ore (e.g. Bojcevski, 2004; Cropp et al., ZKHQEXLOGLQJDPRGHOIRUDSDUWLFXODUÀRWDWLRQ
2013; Dzvinamurungu et al., 2013; Lastra, process. These factors can be distinguished
2007; Petruk and Hughson, 1977). Laboratory in two main groups: characteristics of a cell
ÀRWDWLRQ WHVWV KDYH EHHQ XVHG WR HYDOXDWH WKH IHHG VWUHDP UHSUHVHQWHG E\ ÀRDWDELOLW\ 3  LQ
UHFRYHULHVEDVHGRQWKHÀRDWDELOLW\RIGLIIHUHQW
Eq. (2), and cell operating conditions (Ralston
ore types or geometallurgical units within an ore
et al., 2007; Runge et al., 2003). The latter is
body (Brough et al., 2010; Chauhan et al., 2013;
represented by Sb*Rf in Eq. (2) and comprises
Lotter et al., 2003; Lotter et al., 2011; Suazo et
ÀRWDWLRQFHOOFKDUDFWHULVWLFVIURWKUHFRYHU\ 5f),
al., 2010). Suazo et al. (2010) developed unit
DQG WKH EXEEOH VXUIDFH DUHD ÀX[ 6b), which
SDUDPHWHUVUHSUHVHQWLQJWKHLQKHUHQWÀRDWDELOLW\
of geometallurgical units. They used these include parameters such as entrainment, bubble
parameters and the particle size distribution to GLDPHWHU VXSHU¿FLDO JDV YHORFLW\ EXEEOH ULVH
predict the recoveries for each unit and then velocity and turbulent energy dissipation (e.g.
IRUHFDVWHGWKHUHFRYHU\RIDÀRWDWLRQSODQW7KH Ralston et al., 2007). The bubble surface area
approach was successfully applied in the case ÀX[ LV NQRZQ WR EH SDUWO\ GHSHQGHQW RQ IHHG
study of a porphyry copper deposit (Suazo et stream parameters, such as particle size (Gorain
al., 2010). A different approach was suggested et al., 1999). Feed characteristics (P), on the
by Bulled and McInnes (2005), who aimed other hand, include particle size distribution,
to measure the pulp kinetics for each of the degree of liberation or mineralogical (surface)
PLQHUDO VSHFLHV LQGHSHQGHQW RI SUHGH¿QHG RUH composition of particles, degree of surface
W\SHV 7KH PLQHUDOVSHFL¿F SXOS NLQHWLFV ZHUH R[LGDWLRQ FRQWDFW DQJOH RI PLQHUDOV W\SH DQG
applied for a whole mine block model (Bulled H[WHQWRIUHDJHQWFRYHUDJHSDUWLFOHGHQVLW\DQG
and McInnes, 2005). For such an approach to shape and degree of aggregation of the particles
be successful, the particle size distribution and in the system (Niemi, 1995; Runge et al., 2003;
liberation distribution should not change within Welsby et al., 2010 and reference therein).
WKHÀRWDWLRQIHHG)XUWKHULWLVDVVXPHGWKDWWKH
GHFRXSOLQJRIWKHPLQHUDOVSHFL¿FSXOSNLQHWLFV As the feed characteristics (P) cannot be assumed
and the ore type is admissible. to be constant, but instead will vary in relation to
WKHYDULDELOLW\DQGFRPSOH[LW\RIDQRUHERG\LW
5HFRYHULHV REWDLQHG LQ D ÀRWDWLRQ SURFHVV DUH KDVEHHQVXJJHVWHGWRGHVFULEHWKHÀRWDWLRQIHHG
RIWHQGHVFULEHGE\¿UVWRUGHUUDWHSURFHVV HJ E\DVSHFWUXPRIÀRDWDELOLWLHVGLVWLQJXLVKLQJIDVW
Niemi, 1995; Runge et al. 7KH¿UVWRUGHU VORZDQGQRQÀRDWLQJFRPSRQHQWV 5XQJHet al.,
rate equation for the mineral recovery in a plug 2003). The classes include particles of variable
ÀRZ ZLWK D SHUIHFWO\ PL[HG HQYLURQPHQW LV size and composition. If the data is recorded
written as: only on unsized basis, the underlying physical
(1) R = mi . (1 – e(-k t))
i
.
properties cannot be directly related to fast, slow
DQGQRQÀRDWLQJFRPSRQHQWV,WLVNQRZQWKDWWKH
with: R – recovery, ki – kinetic constant or rate of
ÀRWDWLRQFRPSRQHQWPRGHOLVRQO\YDOLGZKHQQR
ÀRWDWLRQIRUWKHFRPSRQHQWLPi – mass fraction
of component i, t – discrete time in a batch cell VHYHUHFKDQJHVLQWKHÀRWDWLRQUHJLPHRFFXULH
(Runge et al. 1997; Welsby, Vianna, Franzidis QR UHDJHQW DGGLWLRQ DJJUHJDWLRQ R[LGDWLRQ RU
2010). regrinding in the circuit, and it faces challenges

2
when feed parameters, such as head grade and presented in previous contributions (Minz et al.,
particle size distribution, change (Runge et al., 2015a; Minz et al., 2015b). The main focus of the
1997; Welsby et al., 2010). Welsby et al. (2010) present study is to formulate a model, which can
showed that a narrow particle size fraction and forecast the Sb recovery and grade of the Cu–
liberation interval of the valuable mineral could Pb concentrate, based on the distribution of Sb
EHPDWFKHGZLWKWKHIDVWÀRDWLQJFRPSRQHQWDV PLQHUDOVLQSDUWLFOHVLQWKHIHHGWRÀRWDWLRQ7KLV
GH¿QHG E\ WKH ÀRDWDELOLW\ FRPSRQHQW PHWKRG study should also show whether the prediction
Ideally, it means that physical particle parameters of Sb recovery and grade is more precise when
FRXOGEHDVVLJQHGWRWKHÀRDWDELOLW\FRPSRQHQWV WKH ÀRDWDELOLW\ RI 6E PLQHUDOV LV GHULYHG IURP
Thus, if the relevant physical parameters of the analysis providing particle information (mineral
ÀRWDWLRQ IHHG FDQ EH H[WUDFWHG IURP WKH RUH liberation level) than by using bulk mineralogy
this would allow a connection of the variability (sized and unsized). For comparison with
DQG FRPSOH[LW\ IRXQG LQ DQ RUH ERG\ DQG WKH PLQHUDORJLFDO GDWD WKH ÀRDWDELOLW\ ZDV DOVR
IRUHFDVWLQJRIWKHÀRWDWLRQHI¿FLHQF\LQWHUPVRI modelled based on chemical composition.
grades and recoveries (cf. Lamberg, 2011).
2. Methods
In this study, the particle composition and size
were selected as physical parameters relevant
WR ÀRWDWLRQ PRGHOOLQJ DQG VLPXODWLRQV )RU 2.1. 0DWHULDODQG([SHULPHQWDO6HWXS
VLPSOLFLW\ RQO\ RQH ÀRDWDELOLW\ FRPSRQHQW This study is based on the results from laboratory
was used. The assumption behind this is that ÀRWDWLRQ WHVW GRQH RQ IRXU FRPSRVLWH VDPSOHV
WKH FRPSOH[LW\ LQ UHFRYHULHV ZKLFK LV PHW representing four Sb-rich mineralisation
by different components (fast, slow and non- types of the Rockliden deposit three massive
ÀRDWLQJ  LQ WKH ÀRDWDELOLW\ FRPSRQHQWV PRGHO sulphide samples (two Zn-dominated and
FDQEHVLPSOL¿HGWRRQHFRPSRQHQWZKHQHDFK RQH &XGRPLQDWHG  DQG D PD¿F G\NH VDPSOH
FRPSRQHQW LV GH¿QHG E\ LWV PLQHUDORJ\ DQG Detailed descriptions of the sampling, analytical
degree of liberation for narrow size fractions. SURFHGXUH DQG WKH UHVXOWV RI ÀRWDWLRQ WHVWV DUH
7KH ÀRDWDELOLW\ ZDV PRGHOOHG EDVHG RQ WKH reported in Minz et al. (2015; 2015). Only a
recovery of minerals, on unsized and sized basis, summary is provided here. The samples were
DQG ¿QDOO\ EDVHG RQ WKH UHFRYHU\ RI OLEHUDWLRQ ground in a laboratory rod mill to a P80 of about
classes in narrow particle size fractions. In the  —P7KH ÀRWDWLRQ WHVW ZDV FRQGXFWHG XVLQJ
ODWWHU FDVH LW LV DVVXPHG WKDW ÀRDWDELOLW\ RI D D0DJRWWHDX[ERWWRPGULYHQÀRWDWLRQPDFKLQH
pure liberated particle in a narrow size range is 7KHFHOOVL]HRIWKHÀRWDWLRQPDFKLQHZDVO
virtually constant in different ore types at similar the percentage of solids used was 46 to 49 wt%
ÀRWDWLRQFRQGLWLRQV *RUDLQet al., 2000). IRU WKH IRXU FRPSRVLWHV DQG WKH XVHG ÀRWDWLRQ
UHDJHQWV DUH VKRZQ LQ ¿JXUH  7KH IURWK ZDV
The mineralogical information used in this removed with a scraper at constant pace and depth
VWXG\ ZDV FROOHFWHG RQ ÀRWDWLRQ SURGXFWV RI JLYHQE\WKHGHVLJQRIWKHÀRWDWLRQPDFKLQH7KH
the polymetallic Cu–Zn massive sulphide WRWDOÀRWDWLRQWLPHLQWKHURXJKHUVWDJHVZDV
Rockliden deposit. Antimony (Sb) is a minor minutes and 3.5 minutes for the cleaner stages
to trace element in the Rockliden ore, which (Fig. 1). Samples were dried, weighted, and
potentially lowers the quality of the Cu– sieved into three size fractions, and polished
Pb concentrate since it is a penalty element HSR[\ PRXQWV IRU PLQHUDO OLEHUDWLRQ VWXGLHV
at Cu smelters (Larouche, 2001). The Sb were prepared. It should be noted that the test
PLQHUDORJ\RIWKH5RFNOLGHQGHSRVLWLVFRPSOH[ for the composites was not repeated and hence
comprising tetrahedrite (Cu12Sb4S13), bournonite WKHH[SHULPHQWDOHUURUFDQQRWEHVWXGLHGGLUHFWO\
(PbCuSbS3), meneghinite (Pb13CuSb7S24) and (cf. Chauhan et al., 2013).
gudmundite (FeSbS) as the main Sb carriers
(Minz et al., 2013). Four mineralogically A mineral liberation analyser (MLA) at
distinct composite samples were subjected to the Geometallurgy Laboratory, Technische
D ODERUDWRU\ ÀRWDWLRQ WHVW DQG WKH UHVXOWV ZHUH 8QLYHUVLWlW %HUJDNDGHPLH )UHLEHUJ *HUPDQ\

3
Cu-Pb rougher Zn rougher
ground
composite 5 flotation products for each
5.5 min 5.5 min Zn RT composite sample
(ca. 2 kg
feed material)

Cu-Pb cleaner Zn cleaner sieving into 3 size fractions


(-20 μm, -45 μm, + 45 μm)
3.5 min Cu CT 3.5 min Zn CT

epoxy mounts for


MLA measurements
Cu CC Zn CC

1.5 g/kg Zn-sulphate slaked lime (pH ca. 12)


1 g/kg soda ash 0.3 g/kg Cu-sulphate
K-amyl-xanthate Na-isobuthyl-xanthate particle tracking
Dowfroth 250 Dowfroth 250
Fig. 1.)ORZVKHHWRIWKHODERUDWRU\ÀRWDWLRQWHVWVKRZLQJDOOVDPSOHGÀRWDWLRQSURGXFWV&X&&±&X±3EFOHDQHU
concentrate, Cu CT – Cu–Pb cleaner tailing, Zn CC – Zn cleaner concentrate, Zn CT – Zn cleaner tailing, Zn RT – Zn
URXJKHUWDLOLQJ7KHÀRWDWLRQUHDJHQWVDUHOLVWHGIRUWKH&X±3EDQG=QÀRWDWLRQDQGWKHWRWDOÀRWDWLRQWLPHLVJLYHQLQ
the symbols of the rougher and cleaner stages.

was used to study the mineralogical composite 2009) and then used similar to MLA-based bulk
RI SDUWLFOHV LQ WKH VL]HG ÀRWDWLRQ SURGXFWV mineralogy in the modelling of k-values and
by GXMAP analysis (Fandrich et al., 2007). simulation of grades and recoveries of the four
Results were treated by a particle tracking composites. It should be noted, however, that
technique (Lamberg and Vianna, 2007). Before the bulk mineralogy calculated from chemical
balancing of particles in narrow size fractions, assays does not allow the distinction of different
PLQHUDOVZHUHJURXSHGWRUHGXFHWKHFRPSOH[LW\ Sb minerals.
in mineralogy given in MLA measurements for
the Rockliden case. A total of 10 mineral groups 2.2. Modelling
ZHUH GH¿QHG IRU HDFK FRPSRVLWH FRPSULVLQJ All modelling described in this manuscript
Sb minerals, base-metal sulphides, Fe sulphides is empirical, i.e. it is based on the results of
and non-sulphide gangue minerals. The main Sb WKH H[SHULPHQWDO GDWD IRU WKH IRXU VWXGLHG RUH
minerals were kept separately as they show a FRPSRVLWHV 2QH ÀRWDWLRQ UDWH FRQVWDQW ZDV
GLIIHUHQWÀRWDWLRQGLVWULEXWLRQGXULQJ&X±3EDQG PRGHO¿WWHG IRU HDFK PLQHUDO E\ VL]H IUDFWLRQ
=QÀRWDWLRQ /DJHUDQG)RUVVEHUJ0LQ]et (-20 μm, +20 to -45 μm, +45 μm). The models
al., 2015). The grouped MLA data was binned, of this study are developed for assay-based bulk
i.e. particle classes based on their composition mineralogy (grades and recoveries of minerals
ELQDU\ WHUQDU\ DQG FRPSOH[ SDUWLFOHV  ZHUH derived from element to mineral conversion,
GH¿QHG IRU HDFK VL]H IUDFWLRQ 7KH GDWD ZDV SL0), bulk mineralogy (mineral grade and
further adjusted by advanced binning, so that recoveries derived from MLA data, on unsized
mass-balancing was possible by reaching a and sized basis, SL1 and SL2, respectively) and
VXI¿FLHQW QXPEHU RI SDUWLFOHV LQ HDFK SDUWLFOH for fully liberated particles (derived from mass
group of each product (Lamberg and Vianna, balancing of MLA data, SL3) (cf. Table 1).
2007). The processed MLA results will be 8VLQJLQIRUPDWLRQRQVL]HGEDVLVJLYHVDWRWDORI
UHIHUUHGWRDVH[SHULPHQWDOGDWD  PLQHUDOV [ VL]HIUDFWLRQV  ÀRWDWLRQ
UDWHV LQ RSWLRQ 6/ DQG 6/  PLQHUDOV  [
As an alternative approach, a bulk mineralogy  VL]H IUDFWLRQ    ÀRWDWLRQ UDWHV LQ RSWLRQ
was computed from chemical assays provided 6/ DQG  FDOFXODWHG PLQHUDOV  [  VL]H
IRUDOOÀRWDWLRQSURGXFWV7KH&X=Q3E$V6E IUDFWLRQV  ÀRWDWLRQUDWHVLQRSWLRQ6/7KH
and S assays were converted into chalcopyrite, ÀRWDWLRQUDWHVDUHPRGHOOHGERWKIRU&X±3EDQG
sphalerite, galena, bournonite, quartz, =Q ÀRWDWLRQ )RU DOO PLQHUDOV 6/ WR 6/  RU
arsenopyrite, pyrite and quartz in the Geo module liberated particles of each size fraction (SL3),
of the HSC Chemistry software (cf. Roine, WKH ÀRWDWLRQ UDWHV ZHUH FDOFXODWHG XVLQJ QRQ

4
OLQHDU¿WWLQJWHFKQLTXHVE\PLQLPLVLQJWKHPHDQ
VTXDUH HUURU 06(  RI WKH H[SHULPHQWDO DQG ACV = 100 . [2 / N .Ȉi=1N{(ai – bi)2 /
modelled recoveries. The modelled recoveries (ai + bi)2}](1/2)
were calculated based on Eq. (1), applying the
ÀRWDWLRQWLPHVIRUURXJKHUDQGFOHDQHUÀRWDWLRQ It should be noted that the ACV values were
DVVKRZQLQ¿JXUH developed originally for assay data in ore
geology applications and not for recoveries
2.3. Simulation and model validation (Abzalov, 2008).
'XULQJWKHODERUDWRU\ÀRWDWLRQWHVWVRQWKHIRXU
composite samples, the cell operating conditions 3. Results
were kept constant. The cell parameters were not
considered in modelling, i.e. the bubble surface 3.1. Mineralogical characterisation of test
DUHDÀX[ 6b) and froth recovery were set to the material and distribution of Sb
value of “1” in the simulations leading to k = P
in Eq. 2. Table 2 gives an overview of the bulk
mineralogy of the four composites as derived
)RUHDFKPLQHUDO RUPLQHUDOJURXSDVGH¿QHGLQ from MLA measurements. It shows that the feed
WKHH[SHULPHQWDOGDWD JOREDONYDOXHV ÀRWDWLRQ is dominated by sulphide minerals including
UDWHV  ZHUH FKRVHQ IURP WZR FRPSRVLWHV 85 pyrite, sphalerite and chalcopyrite for all tested
=Q RU /5=Q FRPSRVLWH ERWK DUH FODVVL¿HG PDVVLYH VXOSKLGH FRPSRVLWHV 7KH PD¿F G\NH
as Zn-dominated massive sulphide ore, cf. composite is different as it is dominated by non-
7DEOH   7KH PRGHOOHG PLQHUDOVSHFL¿F N sulphide gangue, mostly comprising amphiboles,
values were used in HSC Chemistry Sim 7.1 phyllosilicates, feldspar minerals and quartz.
modelling and simulation software to forecast More details on the mineralogical characteristics
WKHGLVWULEXWLRQRIWKHPLQHUDOVLQWRWKHÀRWDWLRQ are found in Minz et al. (2015a; 2015a).
products (Roine, 2009). As described above,
IDVW VORZ DQG QRQÀRDWLQJ FRPSRQHQWV ZHUH Table 3 shows the deportment of antimony
not distinguished; instead one component was (Sb) between the Sb minerals and the degree of
XVHG FRQWDLQLQJ PLQHUDOVSHFL¿F N YDOXHV IRU liberation in the feed samples. The data in Table
narrow size classes. Accordingly to the depth  DQG  GHPRQVWUDWHV VLJQL¿FDQW GLIIHUHQFHV
of information on which the k values have in the mineralogy between the feed samples.
been modelled, different detail of mineralogical Systematic correlation between head grades
information was used to describe the feed in the of Sb minerals (Table 2) and their degree of
simulation of the four composites (Table 1). liberation (Table 3) is not apparent. However, it
might be noted that at low grades (< 0.15 wt%)
7KH DYHUDJH FRHI¿FLHQW RI YDULDWLRQ $&9 FI
the degree of liberation of Sb minerals is highly
Abzalov, 2008) has been used to estimate the errors
between the Sb grade and recovery given by the
variable, ranging between 18 and 48 %; with
simulated data (ai  DQG H[SHULPHQWDO GDWD Ei) for
increasing grade (> 0.15 wt%) the degree of
the pairs (a and b) of the rougher and cleaner liberation of Sb minerals narrows to a range of
FRQFHQWUDWHLQ&X±3EÀRWDWLRQ 1   35 to 45 %.

Table 1. 3DUDPHWHURSWLRQVFKRVHQIRUVLPXODWLRQ$EEUHYLDWLRQV6/±VLPXODWLRQOHYHO85=Q±=QGRPLQDWHG
massive sulphide from the upper part of the Rockliden deposit, LR-Zn – Zn-dominated massive sulphide from the
lower part of the Rockliden deposit.

Depth of information Bulk mineralogy Bulk mineralogy Bulk mineralogy Liberated


for feed to simulation computed from derived from derived from particles,
and k values selected chemical assays, MLA data, MLA data, sized sized
IURP85=QDQG/5 unsized unsized
Zn
Simulation option SL0 SL1 SL2 SL3

5
Table 2. Mineralogical composition of the studied composites based on MLA measurements. Abbreviations: Bour
– bournonite, Ccp – chalcopyrite, Gd – gudmundite, Gn – galena, Mene – meneghinite (with boulangerite), NSG –
non-sulphide gangue (with remaining trace minerals), Po – pyrrhotite, Py – pyrite (with arsenopyrite), Sp – sphalerite,
7WU±WHWUDKHGULWH&RPSRVLWHV85&X±&XGRPLQDWHGPDVVLYHVXOSKLGHIURPWKHXSSHUSDUWRIWKH5RFNOLGHQGHSRVLW
85=Q±=QGRPLQDWHGPDVVLYHVXOSKLGHIURPWKHXSSHUSDUWRIWKH5RFNOLGHQGHSRVLW/5=Q±=QGRPLQDWHGPDV-
VLYHVXOSKLGHIURPWKHORZHUSDUWRIWKH5RFNOLGHQGHSRVLW0'±PD¿FG\NH
3\ 6S NSG Po &FS Gn Ttr Bour Mene Gd
&RPSRVLWH
(wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%)
85&X 51.9 4.5 8.6 14.1 19.6 0.6 0.40 0.04 0.02 0.24
85=Q 55.9 14.7 13.4 7.9 4.1 2.5 0.85 0.13 0.05 0.45
LR-Zn 53.6 19.0 13.6 9.1 1.3 2.3 0.01 0.19 0.60 0.35
MD 4.5 0.4 82.6 0.4 2.1 3.6 0.45 1.25 4.58 0.07

Table 3. Deportment of Sb between Sb minerals. For abbreviations see Table 2.


'HSRUWPHQWRI6E Degree of liberation in feed (P80 ca. 50 μm)
Ttr Bour Mene Gd Ttr Bour Mene Gd
85&X 43.2 3.4 1.3 52.1 37.6 18.5 34.2 38.2
85=Q 46.4 5.7 1.8 46.2 43.4 34.1 47.9 44.3
LR-Zn 0.6 12.5 32.1 54.8 29.6 40.9 43.4 36.6
MD 9.6 22.4 64.4 3.7 39.2 43.6 44.0 28.6

Table 4 summarises the distribution of the Sb are recovered mostly into the Cu–Pb cleaner
PLQHUDOV EHWZHHQ GLIIHUHQW ÀRWDWLRQ SURGXFWV concentrate (recoveries > 85 %) if head grades
both for bulk minerals and for liberated particles. of the Cu- and Pb-bearing Sb sulphosalts
,QGLYLGXDO 6E PLQHUDOV VKRZ VLJQL¿FDQW are higher than 0.15 wt% (compare Table 2
differences in their distribution. The distribution and 4). Gudmundite reports largely to the
of liberated Sb minerals is more pronounced tailings (Table 4). The difference in distribution
than the distribution of bulk minerals. Liberated between bulk and liberated gudmundite are
particles of Cu- and Pb-bearing Sb sulphosalts attributed to effects of locking, in particular
(tetrahedrite, bournonite and meneghinite) of locking in chalcopyrite-gudmundite and

Table 4. Distribution of Sb minerals (Gd – gudmundite, Bour – bournonite, Mene – meneghinite, Ttr – tetrahedrite)
LQWRWKUHH¿QDOÀRWDWLRQSURGXFWV &X&&±&X±3EFOHDQHUFRQFHQWUDWH=Q&&±=QFOHDQHUFRQFHQWUDWH=Q57±URXJK-
HUWDLOLQJ RIHDFKFRPSRVLWH 85&X±XSSHU5RFNOLGHQ&XGRPLQDWHGPDVVLYHVXOSKLGH85=Q±XSSHU5RFNOLGHQ
=QGRPLQDWHGPDVVLYHVXOSKLGH/5=Q±ORZHU5RFNOLGHQ=QGRPLQDWHGPDVVLYHVXOSKLGH0'±PD¿FG\NH %ROG
are the two main Sb host minerals of each composite.
'LVWULEXWLRQRIEXONPLQHUDOV 'LVWULEXWLRQRIIXOO\OLEHUDWHGSDUWLFOHV
Sb% Ttr % Bour % Mene % Gd % Ttr % Bour % Mene % Gd %
85&X CuCC 24.2 82.0 69.9 54.9 14.0 85.9 46.4 72.3 8.3
85&X ZnCC 6.7 0.9 2.3 5.5 10.1 0.4 2.9 3.3 10.4
85&X ZnRT 42.4 0.4 13.7 22.4 36.9 0.3 30.8 11 41.6
85=Q CuCC 63.1 96.2 83.0 70.1 28.5 96.2 79.7 78.4 13.2
85=Q ZnCC 4.3 0.4 1.2 3.1 8.5 0.2 0.7 0.9 9.2
85=Q ZnRT 16.3 0.2 7.8 7.0 32.7 0.0 10.6 3.4 46.3
LR-Zn CuCC 56.8 93.5 87.5 94.6 28.3 95.9 90.0 97.8 12.8
LR-Zn ZnCC 4.3 2.0 1.6 1.1 6.8 2.0 0.5 0.6 4.5
LR-Zn ZnRT 21.1 0.4 5.4 1.4 35.4 0.4 5.2 0.3 52.1
MD CuCC 82.9 66.0 80.6 77.3 13.4 86.6 91.1 87.2 9.6
MD ZnCC 5.8 9.9 7.1 5.9 37.5 1.0 0.5 0.3 3.0
MD ZnRT 1.5 12.4 3.4 3.3 16.5 3.6 2.2 2.2 45.8

6
galena-gudmundite binary particles (Minz et al., 3.2. )ORWDWLRQUDWHFRQVWDQWVDQGUHFRYHU\RI
2015a; Minz et al., 2015b). The hypothesis that binaries
similar particle types (in terms of composition 7KHÀRWDWLRQUDWHFRQVWDQWV NYDOXHV KDYHEHHQ
and size, e.g. chalcopyrite-gudmundite and calculated for all minerals in the four composites
galena-gudmundite binary particles) show according to the simulation options listed
VLPLODUÀRWDWLRQEHKDYLRXU±UHJDUGOHVVRIEXON LQ 7DEOH  7KH ÀRWDWLRQ UDWH FRQVWDQWV ZHUH
composition, location or ore type of the feed – calculated for liberated particles per size fraction
was supported by quantitative mineralogical LQWKHWKLUGVLPXODWLRQOHYHO 6/ ,Q¿JXUH
VWXGLHVRQWKHIHHGDQGÀRWDWLRQSURGXFWV 0LQ] the size-recovery curves (SL3 and SL2) and
et al., 2015a; Minz et al., 2015b). unsized values (SL1) are shown for chalcopyrite,

Flotation rates for main minerals in Cu-Pb flotation Flotation rates for Sb minerals in Cu-Pb flotation
1.6
1.8 Ttr
Ccp

0.8

0.9
unsized -20 μm +20 μm +45 μm

1.4
Bour

0.0
0.7
unsized -20 μm +20 μm +45 μm

0.2
Sp
unsized -20 μm +20 μm +45 μm

1.4
Mene

0.1
0.7

unsized -20 μm +20 μm +45 μm

0.0
unsized -20 μm +20 μm +45 μm Gd
0.2
0.06
NSG

0.1

0.0
0.03
unsized -20 μm +20 μm +45 μm

selected k (SL3 = {-20,+20,+45 μm})

unselected k (SL3 = {-20,+20,+45 μm})

selected k (SL2 = {-20,+20,+45 μm}, SL1 = unsized)


0.00
unsized -20 μm +20 μm +45 μm unselected k (SL2 = {-20,+20,+45 μm}, SL1 = unsized)

Fig. 2. Flotation kinetics (k values) for chalcopyrite (Ccp), sphalerite (Sp), non-sulphide gangue (NSG) and Sb miner-
DOV 7WU±WHWUDKHGULWH%RXU±ERXUQRQLWH0HQH±PHQHJKLQLWH*G±JXGPXQGLWH LQ&X±3EÀRWDWLRQ%ODFNOLQHVVKRZ
WKHVL]HUHFRYHU\FXUYHVRIPRGHO¿WWHGNYDOXHVVHOHFWHGIURP=QGRPLQDWHGPDVVLYHVXOSKLGHFRPSRVLWHV 85=QRU
LR-Zn). Grey lines show size-recovery curves for the remaining composites to illustrate the range of k values.

7
sphalerite, non-sulphide gangue and Sb minerals Fig. 2). This is related to the fact that chalcopyrite
LQ&X±3EÀRWDWLRQ$VH[SHFWHGFKDOFRS\ULWHDV includes liberated and non-liberated particles
well as Cu- and Pb-bearing Sb sulphosalts show in mineral-by-size modelling and simulation
KLJK N YDOXHV LQ &X±3E ÀRWDWLRQ )LJ   7KH RSWLRQ 6/  DQG LW UHÀHFWV WKH PRUH
VDPHLVWUXHIRUVSKDOHULWHLQ=QÀRWDWLRQ7KHN pronounced distribution of liberated particles
values of the medium size fraction (20 to 45 μm) into Cu–Pb cleaner concentrate on which the
DUHRIWHQKLJKHVWZKLFKLVWKHH[SHFWHGSDWWHUQIRU k values of option SL3 are based (cf. Table 4).
size-recovery curves (e.g. Trahar, 1981; Trahar Correspondingly, for sphalerite the k values of
and Warren, 1976; Fig. 2). It should be noted liberated particles are lower than for the bulk
that composite LR-Zn has a low chalcopyrite mineral in size-recovery curves (Fig. 2).
grade. The highest k value of chalcopyrite in the
smallest size fraction is thus attributed due to The behaviour of binaries is derived from
the fact that in the case of the LR-Zn composite H[SHULPHQWDOGDWDDQGVHOHFWHGW\SHVDUHVKRZQ
OLEHUDWHGFKDOFRS\ULWHDOPRVWH[FOXVLYHO\UHSRUWV LQ ¿JXUH  *HQHUDOO\ LW FDQ EH VHHQ IURP WKH
into the Cu–Pb cleaner concentrate. bent shape to the recovery curves (Fig. 3), that a
small amount of Cu- or Pb-bearing minerals (5 to
7KH ÀRWDWLRQ UDWH FRQVWDQWV N YDOXHV  IRU 10 wt%) will give high recovery of such binary
chalcopyrite in option SL3 are higher than in particles into the Cu–Pb cleaner concentrate.
option SL2 (compare size-recovery curves, +RZHYHUWKHÀRWDWLRQPRGHODYDLODEOHLQ+6&

100 100
CuCC (UR-Zn) CuCC (LR-Zn)

80 80
recovery %

recovery %

60 NOP in feed: 60 NOP in feed:


-20 μm : 4025 -20 μm : 5600
+20 μm : 5723 40 +20 μm : 8054
40
+45 μm : 3454 +45 μm : 6920
20
20

0 20 40 60 80 100 0 20 40 60 80 100
Sp wt% in Ccp-Sp binaries Sp wt% in Gn-Sp binaries
100
CuCC (UR-Zn) ZnCC (UR-Zn)
60
80
NOP in feed:
recovery %

recovery %

60 40 -20 μm : 839
NOP in feed:
+20 μm : 824
-20 μm : 139
40 +45 μm : 430
+20 μm : 239 20
+45 μm : 128
20

0 20 40 60 80 100 0 20 40 60 80 100
Gd wt% in Ccp-Gd binaries Ttr wt% in Sp-Ttr binaries

ZnCC (UR-Zn) ZnRT (LR-Zn)


60 80

NOP in feed: 60
recovery %

recovery %

-20 μm : 4025 NOP in feed:


40 -20 μm : 1577
+20 μm : 5723
+45 μm : 3454 40 +20 μm : 4685
+45 μm : 7773
20
20

0 20 40 60 80 100 0 20 40 60 80 100
Ccp wt% in Sp-Ccp binaries Gn w% in Py-Gn binaries
-20 μm +20 μm +45 μm

Fig. 3.([DPSOHVRIEHQGHGVKDSHRIUHFRYHU\FXUYHVIRUVHOHFWHGELQDULHVLQVHOHFWHGSURGXFWV;D[HVZHLJKW ZW 


RIRQHPLQHUDOLQELQDU\SDUWLFOHV WKHZHLJKWRIRWKHUPLQHUDOLVWKHGLIIHUHQFHWR <D[HVUHFRYHU\RIWKH
ELQDU\SDUWLFOHDVGH¿QHGE\H[SHULPHQWDOGDWD$EEUHYLDWLRQV123±QXPEHURISDUWLFOHV E\VL]HIUDFWLRQ IRURWKHU
abbreviations see Table 2.

8
Chemistry Sim 7.1 uses linear relation for simulation options SL0 to SL2 generate higher
recovery curves of binary particles. It is clear recoveries for Cu and Sb than actually observed.
IURP¿JXUHWKDWIRUWKHVLPXODWLRQRIÀRWDWLRQ The opposite is true for the simulation options
to be successful non-linearity must be included SL3. Recovery-curve of option SL3 tend to
in model option SL3. The implementation of UHÀHFW WKH VKDSH RI WKH H[SHULPHQWDO FXUYH RI
a non-linearity model is currently not possible FRPSRVLWH85=QDQG/5=QPRUHFORVHO\WKDQ
by the HSC Chemistry Sim 7.1 modelling and option SL2 to SL0, and in some cases give even
simulation software. recovery-grade curves apparently closest to
WKHFXUYHVEDVHGRQWKHH[SHULPHQWDOGDWDHJ
3.3. &RPSDULVRQRIGLIIHUHQWOHYHOVRI the Sb grade-recovery curve for option SL3 of
simulation FRPSRVLWH85&X )LJ 7KHUHFRYHU\JUDGH
To critically assess the different simulation curves for simulation based on the sized and
options recovery-grade curves are shown for unsized bulk mineralogy (SL2 and SL1) are
Cu and Sb in the Cu–Pb circuit (Fig. 4). The very similar (Fig. 4). The results obtained for

Cu Sb
35 1
UR-Cu UR-Cu
30
0.8
25
grade wt%
grade wt%

20 0.6
15 0.4
10
0.2
5
0 0
70 80 90 100 20 40 60 80 100

16 4
UR-Zn UR-Zn
12 3
grade wt%
grade wt%

8 2

4 1

0 0
60 70 80 90 100 40 60 80 100

5 3
LR-Zn LR-Zn
4 2.5
2
grade wt%
grade wt%

3
1.5
2
1
1 0.5
0 0
40 60 80 100 20 40 60 80 100

10 12
MD MD
8 10
grade wt%
grade wt%

8
6
6
4
4
2 2
0 0
60 70 80 90 100 50 60 70 80 90 100

Cu recovery % Sb recovery %
SL0 SL1 SL2 SL3 EXP

Fig. 4.5HFRYHU\JUDGHFXUYHVFRPSDULQJUHVXOWVIURPWKHVLPXODWLRQZLWKWKHH[SHULPHQWDOGDWD (3; 6LPXODWLRQ


RSWLRQV 6/6/6/DQG6/ DUHOLVWHGLQ7DEOH&RPSRVLWHV85&X±XSSHU5RFNOLGHQ&XGRPLQDWHGPDVVLYH
VXOSKLGH85=Q±XSSHU5RFNOLGHQ=QGRPLQDWHGPDVVLYHVXOSKLGH/5=Q±ORZHU5RFNOLGHQ=QGRPLQDWHGPDVVLYH
VXOSKLGH0'±PD¿FG\NH

9
simulation based on unsized bulk mineralogy option (Table 5). The ACV values consider only
(SL1) deviates only slightly from that based on WKH UHVXOWV IURP &X±3E ÀRWDWLRQ 6LPXODWLRQV
size-by-size bulk mineralogy (SL2, Fig. 4). The based on bulk mineralogy (SL1 and SL2) give
recovery-grade curves of simulation option SL0 similar, in some cases even smaller ACV values
DUH VWLOO FORVH WR H[SHULPHQWDO GDWD EXW VKRZ WKDQIRUPRGHORSWLRQ6/ HJFRPSRVLWHV85
more variable shapes in comparison to curves of Zn and LR-Zn, Table 5). It should be noted that
simulation option SL1 to SL3. LWLVWKHVHWZRVDPSOHV 85=QDQG/5=Q IURP
which k values have been selected for modelling.
7KH QRQOLQHDU ÀRWDWLRQ EHKDYLRXU RI VHOHFWHG The observation that ACV values, in particular
binary particle types was implemented in IRU WKH 6E UHFRYHU\ RI WKH 85=Q DQG /5=Q
simulation using k values for the liberated composites are partly smaller for model option
particles from the same composite. The recovery- SL1 or SL2 than for SL3 might be attributed to
grade curves of the tested cases were closer to the the fact that the distribution of binary and more
H[SHULPHQWDOGDWDLQFRPSDULVRQZLWKWKHFXUYHV FRPSOH[ SDUWLFOHV LV QRW FRUUHFWO\ PRGHOOHG
of simulation option SL3, but did not come to with option SL3. The k values based on the
FRQJUXHQFH ZLWK WKH H[SHULPHQWDO GDWD 0RUH bulk mineralogy in option SL2 and SL1 might
research is required to enable incorporation of compensate the missing modelling of binary and
all non-linear behaviour of binary and more FRPSOH[SDUWLFOHVLQVLPXODWLRQ6/+RZHYHU
FRPSOH[SDUWLFOHVLQVLPXODWLRQ ACV values for option SL3 are partly lower for
WKH 85&X DQG 0' FRPSRVLWH WKDQ YDOXHV RI
7DEOHFROODWHVH[SHULPHQWDODQGVLPXODWHG&X option SL2 and SL1. The plain bulk mineralogy
Pb, and Sb grades and recovery of the Cu–Pb 6/ DQG 6/  GRHV QRW GLUHFWO\ WDNH WH[WXUDO
cleaner concentrate. In addition ACV values IHDWXUHVLQWKHRUHLHWKHIHHGWRÀRWDWLRQLQWR
are also given in Table 5, comparing grades and consideration. The slightly higher precision of
UHFRYHULHV RI WKH H[SHULPHQWDO DQG VLPXODWHG 6/DQG6/FRPSDUHGWR6/IRUWKH85=Q
GDWDLQWKH&X±3EÀRWDWLRQIRUHDFKVLPXODWLRQ DQG/5=QFRPSRVLWHVWKDQIRUWKH85&XDQG

Table 5. 7KHGLVWDQFHEHWZHHQH[SHULPHQWDOO\DQGVLPXODWHGGDWDSRLQWVRI&X3EDQG6EJUDGHVDQGUHFRYHULHV
in the grade-recovery curves (cf. Fig. 4) is given for the Cu–Pb cleaner concentrates for the respective simulation
option (SL0 to SL3). ACVYDOXHV  DUHJLYHQIRUFRPSDULVRQRIH[SHULPHQWDODQGVLPXODWHGDVVD\VDQGUHFRY-
HULHVLQWKHURXJKHUDQGFOHDQHUFRQFHQWUDWHRIWKH&X±3EÀRWDWLRQFLUFXLW6LPXODWLRQRSWLRQVDUHOLVWHGLQ7DEOH
For abbreviations see Fig. 4.
Distance $&9YDOXHVIRUDVVD\V ACV values for recoveries
D(Cu) D(Pb) D(Sb) Cu wt% Pb wt% Sb wt% Cu % Pb % Sb %
85&X SL3 6.9 14.9 2.9 15.3 2.9 11.1 2.9 14.2 6.5
85&X SL2 9.4 7.4 9.4 13.3 22.9 28.6 4.6 5.3 11.4
85&X SL1 10.8 7.2 9.2 13.7 24.2 30.2 5.7 5.3 11.6
85&X SL0 8.5 14.1 15.9 14.6 8.3 38.3 3.3 10.3 21.7
85=Q SL3 16.5 17.4 17.4 28.6 28.5 22.6 12.1 12.3 18.0
85=Q SL2 5.7 7.2 2.8 25.8 26.2 24.8 3.1 0.9 3.1
85=Q SL1 5.8 7.4 3.4 26.0 26.5 24.6 3.2 1.1 3.7
85=Q SL0 6.4 9.0 4.1 25.7 24.5 18.2 3.6 3.7 7.8
/5=Q SL3 21.3 25.5 17.7 22.3 26.5 22.8 22.7 17.3 19.3
/5=Q SL2 12.2 7.6 4.2 28.1 21.5 17.5 17.0 3.1 7.8
/5=Q SL1 13.1 7.7 4.7 28.8 21.9 17.0 17.6 3.3 8.5
/5=Q SL0 13.1 8.8 2.4 30.2 22.5 17.5 17.9 3.3 14.8
MD SL3 4.5 14.5 8.1 8.5 16.5 4.2 3.4 6.5 7.1
MD SL2 11.5 32.8 14.6 10.8 21.5 11.0 11.1 21.4 10.7
MD SL1 12.2 33.3 14.8 11.5 21.9 11.2 11.5 21.7 10.8
MD SL0 2.0 27.5 15.0 9.0 23.5 4.8 5.4 15.8 11.3

10
MD composites indicates limits in the portability interaction (e.g. slime coating). They pointed out
(or global usage) of k values used for forecasting that the number of such interactions is largely
of Sb recoveries (see subsection 4.2). unknown. However, such interactions cannot be
discerned by the approach taken in this study.
4. Discussion
4.2. 8VDELOLW\RIVLPXODWLRQUHVXOWVIRU
4.1. *OREDOXVDJHRINYDOXHV geometallurgical modelling
0RGHOOLQJ WKH PLQHUDOVSHFL¿F N YDOXHV According to purpose of sampling different
ÀRWDWLRQUDWHFRQVWDQWV IURPH[SHULPHQWDOGDWD tolerance levels were suggested for the
showed that, although the mineralogy of the fundamental error in sampling (sFE) by Pitard
composites is different (cf. Table 2 to 4), the k (1993): sFE”IRUFRPPHUFLDOSXUSRVHVFE
values are similar for liberated particles in all ”IRUWHFKQLFDOVDPSOLQJDQGSURFHVVFRQWURO
IRXU FRPSRVLWHV 85&X 85=Q /5=Q DQG and sFE”IRUH[SORUDWLRQDQGHQYLURQPHQWDO
MD, Fig. 2). For simulation options SL3 to SL1, VDPSOLQJ ([SDQGLQJ WKLV FODVVL¿FDWLRQ WR WKH
PLQHUDOVSHFL¿FNYDOXHVZHUHFKRVHQIURPWZR ACV values, these thresholds can be used to
FRPSRVLWHV 85=QDQG/5=Q ,WZDVVKRZQ evaluate the applicability of different simulation
that the precision is similar for results from options in the forecasting of grades and recoveries
VLPXODWLRQV FRQGXFWHG RQ WKH FRPSRVLWHV 85 by geometallurgical models. The results indicate
=QDQG/5=Q IURPZKLFKWKHPLQHUDOVSHFL¿F that the model, based on bulk mineralogy of the
k values were selected and for simulations using IHHGDQGPLQHUDOVSHFL¿FNYDOXHVGHULYHGIURP
feed information from the remaining composites WZR WHVWHG FRPSRVLWHV LV VXI¿FLHQW WR IRUHFDVW
85&X DQG 0' 7DEOH   7KLV LQGLFDWHV WKDW the Sb recoveries at a precision of less than 16%
using k values derived from distribution of (Table 5). Thus, simulation options SL2 and
OLEHUDWHG SDUWLFOHV IURP WKH ÀRWDWLRQ WHVW RQ D 6/FRXOGEHXVHGIRUPRGHOVLQWKHH[SORUDWLRQ
different sample is permissible, i.e. mineral- stage or a scoping study. The studied simulation
VSHFL¿FNYDOXHVPLJKWEHXVHGJOREDOO\ZLWKLQ options (SL0 to SL3) are not recommended to be
a deposit. DSSOLHGEH\RQGWKHH[SORUDWLRQVWDJH

However, it should be noted that the mineralogy For simulation option SL3, the assumption of non-
of tested composite samples represents, by linear recovery behaviour for binary particles is
LQWHQWLRQ H[WUHPH FDVHV ERWK LQ WKH PDLQ currently not correctly implemented in modelling
mineralogy and in the Sb trace mineralogy and simulation with HSC Chemistry Sim 7.1.
in respect to ore from the Rockliden deposit This is a reason for relatively high ACV values
(Minz et al., 2015b; Minz et al., 2014). Holding compared to other simulation options (SL2 to
H[SHULPHQWDOFRQGLWLRQVFRQVWDQWLQWKHÀRWDWLRQ 6/ +RZHYHULWLVH[SHFWHGWKDWDIXOOSDUWLFOH
WHVWGRHVQRWPHDQWKDWWKHÀRWDWLRQFRQGLWLRQV based simulation, considering simulation of
for each composite were equal, e.g. caused by ELQDULHV DQG PRUH FRPSOH[ SDUWLFOHV ZRXOG
different reagent consumptions. It is observed have considerable advantages especially (but
that the Cu recovery of liberated chalcopyrite is not only) when trace minerals are concerned.
relatively low for the Cu-rich massive sulphide This is reasonable under the assumption that at
FRPSRVLWH 85&X  DOWKRXJK WKH OLEHUDWLRQ decreasing grade and grain size, trace minerals
degree of chalcopyrite is highest in this sample. might be found more frequently locked in
In order to account for the close relation of particles and consequently less governed by
ÀRWDWLRQ FHOO SDUDPHWHUV DQG WKH ÀRDWDELOLW\ RI the behaviour of fully liberated particles. By
particles, modelling of micro-kinetics would be implementing the correct behaviour of binary
required (Danoucaras et al., 2013 and references DQGPRUHFRPSOH[SDUWLFOHVHJE\WDNLQJLQWR
therein). Guy and Trahar (1985) have listed account the bent shape of recovery curves (cf.
VHYHUDOLQWHUDFWLRQVLQWKHÀRWDWLRQFHOOWULJJHUHG Fig. 3), models based on particle information
E\FKDUDFWHULVWLFVRIWKHIHHGVXFKDVR[LGDWLRQ may be considerably improved (referred to as
precipitation, catalytic reactions and solid-solid SL4 in Table 6).

11
'XHWRWKHFRPSOH[LW\RIWKH6EPLQHUDORJ\DW term portability means that model parameters
5RFNOLGHQ LW ZDV H[SHFWHG WKDW WKH PRGHOOLQJ can be derived from one sample or small
and simulation using only assays (SL0) would number of samples and can be globally used
result in much poorer forecast of Sb grades in a large number of samples or ore blocks
and recoveries than using simulation options, in a geometallurgical model or program, if
especially since Sb minerals are not discerned appropriate input information for simulation is
in option SL0. This was indeed the case for the available, e.g. in form of bulk mineralogy and
85&XFRPSRVLWHZKHUHWKHHUURUIRU6EJUDGH WH[WXUDOLQIRUPDWLRQLQFDVHRIVLPXODWLRQRSWLRQ
and recovery is higher in option SL0 compared SL3 and SL4 (cf. Lund, 2013).
to the other options. Based on this observation,
the plain chemical assays should not be used for 4.3. 5HPDUNVRQPRGHOH[SDQVLRQ
the simulation of product quality in the case of A complete decoupling of the feed stream
PLQHUDORJLFDOFRPSOH[RUH characteristics and the cell operating conditions
is not possible (cf. Danoucaras et al., 2013). For
Table 6 compares the advantages and D WHFKQLFDO DSSURDFK LW PLJKW EH VXI¿FLHQW WR
disadvantages of different modelling levels for DGMXVWWKHÀRDWDELOLW\RISDUWLFOHVE\GHWHUPLQLQJ
the simulation of the behaviour of minerals in WKH HIIHFWLYH ÀRDWDELOLW\ RI D VSHFL¿F ÀRWDWLRQ
DJHRPHWDOOXUJLFDOSURJUDP,QWKLVFRQWH[WWKH cell and by controlling the operation conditions

Table 6. Suggested application of mineralogical information for different simulation options


6/WR6/DUHGHVFULEHGLQ7DEOHIRUH[SODQDWLRQRI6/VHHWH[W 
Model 5HIHUHQFHVFDVH
Advantage Disadvantage 5HFRPPHQGDWLRQ
RSWLRQ studies
8VDEOHLQKLJKWRPRGHUDWH
grade ore material, with
Portably might be possible,
simple mineralogy (i.e.
Information of but not recommend with Magnetite ore
mineral grades can be
SL0 assay-derived bulk changes in mineralogy (Lund, 2013;
estimated from chemical
mineralogy needed. (such as it is the case of Sb Parian, 2015).
assays), high liberation
mineralogy).
degree and small variation
LQPLQHUDOWH[WXUHV
Portability in ore bodies
Information of bulk 8VDEOHLQPDWHULDOZLWK
showing only variation
mineralogy needed, FRPSOH[PLQHUDORJ\EXW
in mineral grades. Leads This study, SL1
SL1 no sizing and low variability in mineral
easily to a big number of level.
liberation analysis WH[WXUHV HJJUDLQVL]HDQG
samples to be tested (e.g. in
needed. mineral association).
ÀRWDWLRQWHVWV 
As for SL1. Portability over 8VDEOHIRURUHVZKHUH
Information of bulk
high variation in mineral PLQHUDOWH[WXUHVKDYHD
mineralogy by size This study, SL2
SL2 WH[WXUHVGRHVQRWH[LVW low variability and particle
needed, no liberation level.
Requires sizing and thus size distribution is to be
analysis needed
tedious. optimised.

Labour-intensive, requires
Enables simulation Ores showing high variation This study, SL3
liberation analysis and
of different particle LQPLQHUDOWH[WXUHV HJ OHYHO([DPSOHVIRU
SL3 HVWDEOLVKPHQWRIWH[WXUDO
sizes and mineral grain size and mineral magnetite ore (Lund
classes (arche-types). Needs
WH[WXUHV association). 2013).
model development.

Global k values See disadvantages of SL3. For simulation of trace


and bending factors &RPSOH[PLQHUDORJ\ZLOO elements (< 1 wt%), Sb minerals
SL4 for modelling H[SDQGLQWRODUJHQXPEHU FRPSOH[WUDFHPLQHUDORJ\ in Rockliden
distribution of of mineral combinations in poorly liberated particles of mineralisation.
FRPSOH[SDUWLFOHV particles. trace minerals.

12
carefully. The application of results from one DFFXUDWHO\ UHÀHFW WKH PLQHUDORJLFDO YDULDELOLW\
ÀRWDWLRQ FRQGLWLRQ WR DQRWKHU HJ ZLWK WKH SRWHQWLDOO\ LQÀXHQFLQJ SURFHVVLQJ SDUDPHWHUV
purpose for up-scaling) requires studying Moreover, only Sb-rich composite samples were
WKH FRUUHODWLRQ RI ÀRWDWLRQ FHOO SDUDPHWHUV collected to study the mineralogical end-members
LQÀXHQFLQJ WKH ÀRDWDELOLW\ RI D SDUWLFOH LQ of the Rockliden massive sulphide deposit, but
different environments (e.g. laboratory test and were not intended to cover the variation in the
ÀRWDWLRQ SODQW  *RUDLQ et al., 2000; Niemi, Sb grade at Rockliden and related difference in
1995). This is beyond the scope of this study. It the Sb mineralogy (including differences in bulk
is recommended to conduct research that would 6EPLQHUDORJ\DQGWH[WXUDOSDWWHUQV RISRWHQWLDO
allow scaling-up the results of this study. mining blocks.

4.4. Alternative model 5. Summary and Conclusions


Statistical modelling, commonly used in
In this study several modelling and simulation
geometallurgy, could be chosen as an alternative
options were tested to forecast grades and
approach. By taking this approach, processing
recoveries (of Cu, Pb, Sb, etc.) in the Cu–Pb
attributes such as recovery can be estimated
concentrate of the polymetallic Rockliden
based on geological data. A statistical approach
massive sulphide deposit. Simulations were
would require a large number of laboratory
EDVHGRQPLQHUDOVSHFL¿FÀRWDWLRQUDWHFRQVWDQWV
tests to build a model that forecasts grades and
N YDOXHV  PRGHOOHG DFFRUGLQJ WR D ¿UVWRUGHU
recoveries as a function of ore properties: e.g.
kinetic process. Mineralogical information was
CuRec=f(Cu, Sb, S, Zn head grade, etc.) (e.g.
DYDLODEOH IURP H[SHULPHQWDO GDWD FROOHFWHG
Bulled and McInnes, 2005). Keeney et al. (2011)
for four composite samples. Four levels of
showed that comminution models based on
simulation were distinguished by the details in
chemical assays results in poorer accuracy than
mineralogical information describing the feed
those which also use additional information such
used for simulation and by the corresponding
as petrophysical data. Similarly, it is assumed that
PLQHUDORJLFDOGHWDLOVDYDLODEOHIRUPRGHO¿WWLQJ
only using drill core assays would give poorer
of k values. The following levels were tested
DFFXUDF\ IRUHFDVWLQJ WKH 6E JUDGH RI ÀRWDWLRQ
in the simulation, from most to least detailed:
products than using additional mineralogical
full particle information (option SL3), sized
or petrographical information. However, other
bulk mineralogy (option SL2) and unsized bulk
than the drill core assays, quantitative data is
mineralogy (option SL1), as well as simulation
not consistently available in the Rockliden case.
based on chemical assays only (option SL0).
This means that mineralogical, petrographical
For each mineral a k value was selected from
or petrophysical data and test parameters would
two composites. The application of the selected
need to be collected for a large number of samples
PLQHUDOVSHFL¿FNYDOXHVLQVLPXODWLRQRIJUDGHV
to build a 3D model for simulation of process
and recoveries of the remaining two composites
parameters such as comminution indices and
indicated the portability or global usability of
recoveries, which is more accurate than a model
WKHPLQHUDOVSHFL¿FNYDOXHV
based on assays only (e.g. Bulled and McInnes,
2005). The low number of samples (four) tested
Comparing the recovery-grade curves of the
in the Rockliden case study, does not permit a
simulation results of option SL3 (mineral
model forecasting the recovery based on sound
liberation) to SL0 (chemical assays) showed
statistics connecting geological and processing
that neither simulation option consistently gives
parameters. Variations in the mineralogy
D FORVH PDWFK WR WKH H[SHULPHQWDO GDWD 7KRVH
within the Rockliden massive sulphide deposit
model options based on bulk mineralogy (SL2
ZHUH LGHQWL¿HG LQ SDUWLFXODU FRQFHUQLQJ WKH
and SL1) and on chemical assays (SL0) tend
Sb mineralogy (Minz et al., 2014). Thus, it is
to overestimate recoveries of Cu and Sb, while
not recommended to base a model on a direct
option SL3 tends to underestimate recoveries.
relation between drill core assays and processing
,W ZDV IRXQG VXI¿FLHQW WR XVH EXON PLQHUDORJ\
parameters since the chemical assays might not
data as a basis for simulation of Sb recoveries

13
IRU WKH H[SORUDWLRQ DQG VFRSLQJ VWXG\ VWDJH RI $FNQRZOHGJHPHQWV
a mining project. The simulation options (SL2
and SL1) require less effort for data acquisition, %ROLGHQ0LQHUDO$%LVWKDQNHGIRU¿QDQFLQJWKH
processing and building of a model since no analytical work in this study, and for permission
complete particle analysis is implemented in the to publish this paper. The Helmholtz Institute
simulation. Freiberg is thanked for funding of MLA analyses
presented in this study. Travelling costs for MLA
A large deviation between measured and studies were partly covered by ProMinNET
simulated Sb grades and recoveries was noted (Nordic network of Process Mineralogy and
for one composite using simulation option Geometallurgy funded by Nordforsk). This study
SL0 and thus it is not recommended to use LV SDUW RI D 3K' SURMHFW ¿QDQFHG E\ &$00
only assays for the simulation in the case (Centre of Advanced Mining and Metallurgy).
RI D FRPSOH[ PLQHUDORJ\ 7KH VWXGLHG IHHG
samples all had relatively high Sb grades (about All staff members of the various technical
ZW )XUWKHUWKHSURSRUWLRQRIIDVWÀRDWLQJ laboratories involved in this study are thanked
Sb minerals (Cu- and Pb-bearing Sb sulphosalts) for their support in sample collection and
DQGVORZO\ÀRDWLQJJXGPXQGLWHLVVLPLODUIRUDOO DVVLVWDQFHLQH[SHULPHQWDODQGDQDO\WLFDOZRUN
the massive sulphide composites. Thus, there Colleagues of the Division of Minerals and
is no basis on which the results for simulation Metallurgical Engineering are thanked for their
RSWLRQ6/FRXOGEHH[WUDSRODWHGWRIHHGVDPSOHV help in modelling and simulation work of this
ZLWK GLIIHUHQW UDWLR LQ IDVW DQG VORZÀRDWLQJ study.
Sb minerals or feed samples with different Sb
grades for which also a lower degree of liberation Riia Chmielowski is thanked for proofreading
IRU6EPLQHUDOVLVH[SHFWHG0RUHH[SHULPHQWDO and corrections.
work should be conducted on this issue.

Generally, it is perceived that with increasing


depth of information a more precise modelling
DQG VLPXODWLRQ LV DFKLHYHG LQ WKH FRQWH[W RI
geometallurgy (e.g. Lishchuk et al., 2015). In this
study it could not be shown that a particle-based
simulation (SL3) gives consistently better result
than simulation based on e.g. bulk mineralogy
(SL2 or SL1). This was partly attributed to the
IDFWWKDWUHFRYHULHVRIELQDU\RUPRUHFRPSOH[
particles are not correctly implemented in
the particle-based simulation option (SL3).
Continued research on this issue should reveal
whether the implementation of full particle
LQIRUPDWLRQLQFOXGLQJWKHÀRWDWLRQGLVWULEXWLRQ
RIELQDU\RUPRUHFRPSOH[SDUWLFOHVPD\SURYLGH
better quality forecasts on grades and recovery
for implementation in a geometallurgical model
XVDEOHEH\RQGWKHH[SORUDWLRQVWDJH

14
5HIHUHQFHV Fandrich, R., Gu, Y., Burrows, D. and Moeller,
K. 2007. Modern SEM-based mineral liberation
Abzalov, M. 2008. Quality control of assay analysis, Int. J. Miner. Process., 84, 310–320.
data: A review of procedures for measuring and
monitoring precision and accuracy, Explor. Min. Gorain, B., Napier-Munn, T., Franzidis, J. and
Geol., 17, 131–144. Manlapig, E. 1998. Studies on impeller type,
LPSHOOHUVSHHGDQGDLUÀRZUDWHLQDQLQGXVWULDO
Bojcevski, D. 2004. Metallurgical VFDOH ÀRWDWLRQ FHOO SDUW  9DOLGDWLRQ RI N6b
FKDUDFWHULVDWLRQRI*HRUJH)LVKHUPHVRWH[WXUHV relationship and effect of froth depth, Miner.
DQGPLFURWH[WXUHVMaster Thesis, University of Eng., 11, 615–626.
Queensland, Australia.
Gorain, B., Franzidis, J. and Manlapig, E. 1999.
Brough, C., Bradshaw, D. and Becker, M. 2010. The empirical prediction of bubble surface area
$FRPSDULVRQRIWKHÀRWDWLRQEHKDYLRXUDQGWKH ÀX[LQPHFKDQLFDOÀRWDWLRQFHOOVIURPFHOOGHVLJQ
effect of copper activation on three reef types and operating data, Miner. Eng., 12, 309–322.
from the Merensky reef at Northam, Miner.
Eng., 23, 846–854. Gorain, B., Franzidis, J., Ward, K., Johnson, N,
and Manlapig, E. 2000. Modelling of the Mt Isa
Bulled, D. and McInnes, C. 2005. Flotation URXJKHUVFDYHQJHUFRSSHUÀRWDWLRQFLUFXLWXVLQJ
plant design and production planning through size-by-liberation data, Miner. Metall. Proc., 17,
geometallurgical modelling, SGS MINERALS 173–180.
SERVICES, TECHNICAL BULLETIN 2005-3,
1–8. Guy, P. and Trahar, W. 1985. The effects of
R[LGDWLRQ DQG PLQHUDO LQWHUDFWLRQ RQ VXOSKLGH
Chauhan, M., Napier-Munn, T., Keeney, L. and ÀRWDWLRQ in: Flotation of sulphide minerals,
Bradshaw, D. J. 2013. Progress in developing Elsevier Amsterdam, 6, 91–110.
D JHRPHWDOOXUJ\ ÀRWDWLRQ LQGLFDWRU in: The
second AusIMM international geometallurgy Keeney, L. and Walters, S. G. 2011. A
conference, 30 September – 2 October 2013, methodology for geometallurgical mapping
Brisbane, Australia, 201–206. and orebody modelling, LQ 7KH ¿UVW $XV,00
international geometallurgy conference; 5-7
Cropp, A. F., Goodall, W. R. and Bradshaw, D. September 2011; Brisbane, Australia, 217–225.
-7KHLQÀXHQFHRIWH[WXUDOYDULDWLRQDQG
gangue mineralogy on recovery of copper by Lager, T. and Forssberg, K. 1990. Separation
ÀRWDWLRQIURPSRUSK\U\RUH±$UHYLHZin: The RI DQWLPRQ\ PLQHUDO LPSXULWLHV IURP FRPSOH[
second AusIMM international geometallurgy sulphide ores, Trans. Inst. Min. Metall. (Sect. C:
conference; 30 September – 2 October 2013, Miner. Process. Extr. Metall.), 99, 54–61.
Brisbane, Australia, 279–291.
Lamberg, P. 2011. Particles – the bridge between
Danoucaras, A. N., Vianna, S. M. and Nguyen, geology and metallurgy, in: Conference in
A. V. 2013. A modeling approach using back- Minerals Engineering 2011, 1–16.
calculated induction times to predict recoveries
LQ ÀRWDWLRQ Int. J. Miner. Process., 124, 102– Lamberg, P. and Vianna, S. M. S. 2007. A
108. technique for tracking multiphase mineral
SDUWLFOHVLQÀRWDWLRQFLUFXLWVin: Proceedings VII
Dzvinamurungu, T., Viljoen, K., Knoper, M. and Meeting of the Southern Hemisphere on Mineral
Mulaba-Bafubiandi, A. 2013. Geometallurgical Technology (eds: R M F Lima, A C Q Ladeira, C
FKDUDFWHULVDWLRQ RI 0HUHQVN\ UHHI DQG 8* DW A Da Silva), 195–202.
WKH 0DULNDQD PLQH %XVKYHOG FRPSOH[ 6RXWK
Africa, Miner. Eng., 52, 74–81.

15
Larouche, P. 2001. Minor elements in copper Minz, F. E., Lamberg, P., Wanhainen, C., Bolin,
VPHOWLQJ DQG HOHFWURUH¿QLQJ Master Thesis, N.-J., Bachmann, K. and Gutzmer, J. 2015b.
McGill University, Montreal, Canada. Mineralogical controls on the distribution of
DQWLPRQ\ LQ D EDVHPHWDO ÀRWDWLRQ WHVW DW WKH
Lastra, R. 2007. Seven practical application cases Rockliden massive sulphide deposit, north-
of liberation analysis, Int. J. Miner. Process., 84, central Sweden, in: Proceedings of the 13th
337–347. biennial SGA meeting (Mineral resources in a
sustainable world), Nancy, France, 1439–1442.
Lishchuk, V., Lamberg, P. and Lund, C. 2015.
&ODVVL¿FDWLRQ RI JHRPHWDOOXUJLFDO SURJUDPV Niemi, A. J. 1995. Role of kinetics in modelling
based on approach and purpose, in: Proceedings DQGFRQWURORIÀRWDWLRQSODQWVPowder Technol.,
of the 13th biennial SGA meeting (Mineral 82, 69–77.
resources in a sustainable world), Nancy,
France, 1431–1435. Parian, M. 2015. Development of the
mineralogical path for geometallurgical
Lotter, N., Kormos, L., Oliveira, J., Fragomeni, modeling of iron ores, Licentiate thesis, Luleå
D. and Whiteman, E. 2011. Modern process University of Technology, Sweden.
mineralogy: Two case studies, Miner. Eng., 24,
638–650. Petruk, W. and Hughson, M. R. 1977. Image
analysis evaluation of the effect of grinding
Lotter, N., Kowal, D., Tuzun, M., Whittaker, P. PHGLD RQ VHOHFWLYH ÀRWDWLRQ RI WZR =Q3E&X
DQG .RUPRV /  6DPSOLQJ DQG ÀRWDWLRQ ores, Can. Min. Metall. Bull. (CIM Bulletin), 70,
testing of Sudbury basin drill core for process 128–135.
mineralogy modelling, Miner. Eng., 16, 857–
864. Pitard, F. F. 1993. Pierre Gy’s sampling theory
and sampling practice: Heterogeneity, sampling
Lund, C. 2013. Mineralogical, chemical and correctness, and statistical process control, CRC
WH[WXUDOFKDUDFWHULVDWLRQRIWKH0DOPEHUJHWLURQ Press.
ore deposit for a geometallurgical model, PhD
Thesis, Luleå University of Technology, Sweden. Ralston, J., Fornasiero, D., Grano, S., Duan,
J. and Akroyd, T. 2007. Reducing uncertainty
Minz, F., Bolin N.-J., Lamberg, P. and Wanhainen, LQ PLQHUDO ÀRWDWLRQ²ÀRWDWLRQ UDWH FRQVWDQW
C. 2013. Detailed characterisation of antimony prediction for particles in an operating plant ore,
PLQHUDORJ\LQDJHRPHWDOOXUJLFDOFRQWH[WDWWKH Int. J. Miner. Process., 84, 89–98.
Rockliden ore deposit, north-central Sweden,
Miner. Eng., 52, 95–103. 5RLQH$+6&&KHPLVWU\8VHU¶V*XLGH
- Chemical Reaction and Equilibrium Software
Minz, F. E., Lasskogen, J., Wanhainen, C. and ZLWK ([WHQVLYH 7KHUPRFKHPLFDO 'DWDEDVH DQG
Lamberg P. 2014. Lithology and mineralisation Flowsheet Simulation [computer program].
types of the Rockliden Zn–Cu massive sulphide
deposit, north-central Sweden: Implications for Runge, K. C., Franzidis, J. P., Manlapig, E. V.
ore processing, Trans. Inst. Min. Metall. (Sect. DQG +DUULV 0 &  6WUXFWXULQJ D ÀRWDWLRQ
B: Appl. Earth. Sci.), 123, 2–17. PRGHO IRU UREXVW SUHGLFWLRQ RI ÀRWDWLRQ FLUFXLW
performance, in: Proceedings XXII International
Minz, F. E., Bolin, N.-J., Lamberg, P., Bachmann, Mineral Processing Congress, Cape Town, South
K., Gutzmer, J. and Wanhainen, C. 2015a. Africa, 973–984.
Distribution of Sb minerals in the Cu and Zn
ÀRWDWLRQ RI 5RFNOLGHQ PDVVLYH VXOSKLGH RUH LQ
north-central Sweden, Miner. Eng., 82, 125–135.

16
Runge, K. C., Harris, M.C., Frew, J.A. and
Manlapig, E.V. 1997. Floatability of streams
around the Cominco Red Dog lead cleaning
circuit, in: Proceedings of the 6th annual mill
operators conference, Madang, Papua New
Guinea, 157–163.

Suazo, C., Kracht, W. and Alruiz, O. 2010.


Geometallurgical modelling of the Collahuasi
ÀRWDWLRQFLUFXLWMiner. Eng., 23, 137–142.

Trahar, W. J. 1981. A rational interpretation of


WKHUROHRISDUWLFOHVL]HLQÀRWDWLRQInt. J. Miner.
Process., 8, 289–327.

Trahar, W. J. and Warren, L. J. 1976. The


ÀRWDELOLW\RIYHU\¿QHSDUWLFOHV²$UHYLHZInt.
J. Miner. Process., 3, 103–131.

Welsby, S, Vianna, S., and Franzidis, J. P. 2010.


$VVLJQLQJ SK\VLFDO VLJQL¿FDQFH WR ÀRDWDELOLW\
components, Int. J. Miner. Process., 97, 59–67.

17

Вам также может понравиться