Sie sind auf Seite 1von 52

UNIVERSITY OF GOTHENBURG

Department of Earth Sciences


Geovetarcentrum/Earth Science Centre

Compositional variations
between hydrothermal
and magmatic magnetite
and their potential for
mineral exploration
using till from Laver,
Northern Sweden

Jessica Aronsson

ISSN 1400-3821 B915


Master of Science (120 credits) thesis
Göteborg 2016

Mailing address Address Telephone Telefax Geovetarcentrum


Geovetarcentrum Geovetarcentrum 031-786 19 56 031-786 19 86 Göteborg University
S 405 30 Göteborg Guldhedsgatan 5A S-405 30 Göteborg
SWEDEN
Table of content
1. Introduction ............................................................................................................................ 1
1.1 Background ................................................................................................................................... 1
1.2 Laver –The mine that disappeared ................................................................................................ 3
1.3 Regional geology ........................................................................................................................... 5
1.4 Previous studies............................................................................................................................. 7
2. Magnetite geochemistry ......................................................................................................... 9
2.1 Magnetite -a close-up .................................................................................................................... 9
2.2 Hydrothermal Vs Magmatic magnetite ....................................................................................... 10
3. Method ................................................................................................................................. 12
3.1 Sample preparation ..................................................................................................................... 12
3.2 Scanning electron microscope..................................................................................................... 13
3.3 LA-ICP-MS .................................................................................................................................. 14
4. Results .................................................................................................................................. 15
4.1 SEM ............................................................................................................................................. 15
4.2 Trace elements............................................................................................................................. 17
4.3 Multielement variation diagrams ................................................................................................ 21
4.4 Distribution of hydrothermal magnetite ...................................................................................... 24
4.5 Discriminant diagrams ................................................................................................................ 28
4.6 Laver versus Kiruna .................................................................................................................... 30
5. Discussion ............................................................................................................................ 32
5.1 Limitations for sulphide inclusions as tracers for mineralization ............................................... 32
5.2 Compositional variations between hydrothermal and igneous magnetite .................................. 33
5.3 Hydrothermal magnetite as an indicator mineral ....................................................................... 35
5.4 Discerning the deposition type .................................................................................................... 37
5.5 Differences between Laver and Kiruna magnetite ...................................................................... 38
5.6 Sources of error ........................................................................................................................... 38
6. Conclusions .......................................................................................................................... 40
Acknowledgments .................................................................................................................... 42
References ................................................................................................................................ 42
Appendix .................................................................................................................................. 47
Abstract
The omnipresence of magnetite in different geological systems and its resistance to weathering
and glacial transport implicates the value as an indicator mineral (IM) when prospecting in basal
till. The ability of magnetite to incorporate a large number of both minor and trace elements in
its mineral structure makes the mineral a valuable target for geochemical studies. Variations in
the chemical composition of magnetite might be used to “fingerprint” different deposit types
and thus be exploited in mineral exploration. Magnetite extracted from till samples have been
examined using SEM-EDS and LA-ICP-MS and the trace element composition have been
studied in order to separate hydrothermal and magmatic magnetite from a porphyry-Cu deposit
known as Laver. Large amounts of Ti, Al and V has been discovered in magmatic magnetite
while lower levels of Cr have been observed in the magnetite of hydrothermal origin. Plotting
Ti vs. Ni/Cr ratio indicates a decline in hydrothermal magnetite with the increasing distance
from the source, which seems to imply a possible use as an IM. The trace elements that are
present show differences in the concentrations in igneous and hydrothermal magnetite.
Discrimination of magnetite have been performed using Ni/(Cr+Mn) versus Ti+V and
Ca+Al+Mn versus Ti+V diagrams and the differences between the porphyry deposit and other
deposits have been discussed. Further studies should focus on the question if the distinction
between magmatic and hydrothermal magnetite is sufficient or if other important sources, e.g.,
metamorphic magnetites, need to be considered.

Keywords: Magnetite, indicator mineral, prospecting, till, trace elements, hydrothermal, SEM-
EDS, LA-ICP-MS, porphyry-Cu, Laver.
Sammanfattning
Den allestädes närvarande magnetiten i olika geologiska system och dess motståndskraft mot
vittring under transport implicerar dess värde som ett potentiellt indikatormineral under
prospektering av basal morän. Magnetitens förmåga att införliva ett stort antal ämnen som
substitut i dess spinelstruktur uppmärksammar även dess värde i geokemiska undersökningar.
Variationer i magnetitens kemiska sammansättning kan användas för att undersöka om det finns
ett specifikt ”fingeravtryck” hos olika avsättningar som kan användas inom prospekteringen.
Magnetit extraherad från moränprover insamlade i närheten av en kopparrik porfyr känd som
Laver har blivit undersökta med SEM-EDS och LA-ICP-MS för att undersöka om det är möjligt
att skilja hydrotermala och magmatiska magnetiter från varandra. De spårämnen som har
undersökts uppvisar skillnader i sina koncentrationer mellan magmatiska och hydrotermala
magnetiter. En större mängd Ti, Al och V har upptäckts i magmatiska magnetiter och en lägre
mängd Cr har observeras i de hydrotermala magnetiterna. I diagram som visar Ti mot Ni/Cr
kvoten indikerar dessa en minskning i innehållet av hydrotermala magnetiter i förhållande till
ökande avstånd till källan. Detta påvisar att användning som indikatormineral verkar vara
möjlig. Urskiljandet av hydrotermala magnetiter har skett med hjälp av Ni/(Cr+Mn) mot T+V
samt Ca+Al+Mn mot Ti+V diagram och dessutom har skillnader mot andra mineralavsättningar
diskuterats. Framtida studier borde fokusera på frågan om skillnaden mellan magmatiska och
hydrotermala magnetiter är lämplig eller om andra viktiga källor, t. ex metamorfa magnetiter
borde tas under beaktning.

Nyckelord: Magnetit, indikator mineral, prospektering, morän, spårämnen, hydrotermal, SEM-


EDS, LA-ICP-MS, Cu-porfyr, Laver.
1. Introduction
1.1 Background
Terrains that have been effected by glaciations in the past have experienced glacial erosion,
transport and deposition which have had a substantial effect on the distribution of materials in
unconsolidated sediments (Plouffe, 1996). Previous studies have shown that mineral deposits
eroded by glaciers resulted in debris that are enriched in metals (or other material of interest)
which are transported in the ice flow direction and deposited as till. An area with high levels of
the concerned material are thus paralleling the ice flow direction and is known as a “dispersal
train” (Larson and Mooers, 2004; McClengahan, 2005).

The sought commodities are anomalous near the source and steadily decreases in the direction
of the ice flow relatively to the increasing distance to the source. This decrease is possible due
to the mixing and dilution of background material during the transport (Shilts, 1982). The total
distance of a dispersal train varies, the length are often perceived to extend from the source
mineralization and up to a point where it is no longer possible to distinguish an anomalous
concentration from the background values. In mineral exploration is the word “threshold” used
to designate the concentration that serves as the limit between anomalous and background
values of a sought commodity (Plouffe, 1996).

Materials that are often investigated in this context are known as “indicator minerals” (IM)
which are a minor component in sediment and can be defined as minerals that exhibit a specific
signature of lithological, mineralogical or geochemical nature that can be traced back to a
specific source (Larson and Mooers, 2004; McClengahan, 2005). IM (e.g. magnetite, monazite)
are usually more resistant to different types of weathering and methods of transportation than
other co-existing mineral phases (Nadoll et al., 2014a; McClengahan, 2005). Dispersal tracing
have thus been used in studies of till sedimentology and as an exploration tool in glaciated areas
such as Canada and Fennoscandia (Larson and Moores, 2004; McClenaghan et al., 2000).

Deciding which mineral to use as an IM depends on the type of deposit that is sought and in
some cases one specific mineral can be indicative of more than one type of deposit. The majority
of the minerals used as IMs belong to groups that are common in many geological regions (Gent
at al., 2011). Magnetite is one of the most common Fe-oxide minerals in Earth’s continental
crust. It is a common accessory mineral in a wide variety of rocks (Dupuis and Beaudoin, 2011;
Nadoll et al., 2012) and have been considered to be an important IM in both petrogenetic and
geochemical studies for several decades (Nadoll et al., 2014a).

1
The omnipresence of magnetite in different geological systems and its resistance to weathering
indicates a possibility that magnetite might be a useful indicator mineral when prospecting basal
till (Pisiak et al., 2015). Magnetite is known for the ability of incorporating a large number of
both minor and trace elements in its cubic spinel structure during the formation of the mineral
that might take place under a wide range of different geological conditions (Nadoll et al., 2012).

A number of elements that are often seen as substitutes in magnetite from most types of origins
are Mg, Al, Ti, V, Co, Ni, Zn, Cr, Mn, Sn and, Ga (Nadoll et al., 2014a). These elements are
referred to as spinel elements and are useful as indicator elements to potentially distinguishing
magnetite from different sources based on the difference in concentration and the assemblage
of the trace elements (Nadoll et al., 2012; Huang et al., 2014).

In combination with isotopic studies might the varying chemical compositional in magnetite
provide important constrains on different petrogenetic factors such as temperature, pressure and
oxygen/sulphur fugacity (Nadoll et al., 2014a). The magnetic properties of magnetite have also
been extensively used in geophysical studies, mapping and exploration of different kinds of
mineral deposits around the world (Clark, 2014).

Chemical variations in the environment where the magnetite originates most likely controls the
composition of the mineral. Subsequently, differences in physical and chemical conditions
during formation of a mineral deposit might also control the composition of the iron oxides in
the deposits. The compositional varieties in the spinel elements in the magnetite might be
therefore be used as a “fingerprint” for different mineralizations (Dupuis and Beaudoin, 2011).
As magnetite constitute a large part of the heavy mineral fraction, have the interest of using the
trace element signature to identifying the provenance of the magnetite increased as well as the
interest in possible applications in mineral exploration and sedimentology (Dare et al., 2014).

This geochemical study will concentrate on a porphyry copper deposit known as Laver.
Porphyry copper deposits are often considered to be the result of supercritical fluids originating
in the shallow crust which are leaching elements from a crystallizing magma and transports
them away. It is the eventual cooling of the fluid which leads to the deposition and accumulation
of the elements (Berger et al., 2008; Sillitoe, 2010). Magnetite can form as a major
hydrothermal mineral if the total iron content in the host magma contain sufficient amounts of
the necessary elements (Sillitoe, 1997).

2
The age of porphyry deposits varies, however most of the deposits are of Jurassic age or
younger, in Figure 1 are the porphyry deposits in the world summarized based on the estimated
age (Sinclair, 2007). The pre-Jurassic Laver deposit may, in the light of this, be of academic
importance. The scientific descriptions that currently exist of both the geology and the old
Laver deposit itself are both
relatively obsolete and
scarce. The procedures of
separating the diverse types
of porphyries from each
other has evolved, thus what
as interpreted with crude
measurements in the 1940’s
might not be accurate with
today’s methods. Figure 1. Age distribution of porphyry copper deposits in the world
(from Sinclair, 2007).
The original main objective
with this Master’s thesis was to investigate if using the SEM-EDS to detect magnetite
containing sulphide inclusions and to evaluate whether it would be a useful method during
exploration. The magnetite containing inclusions would have functioned as indicators and
would have been easy to find when screening large amounts of material.

However, it was found that using sulphide inclusions were not applicable in this context. The
aim of the thesis was thus shifted to use trace elements in the magnetite collected from the
heavy mineral fraction in till samples as an indicator of provenance. Which is a subject that
have not been widely researched yet. In addition, the variations in trace elements was evaluated
to ascertain their potential importance in mineral exploration. Further will the magnetite of the
old Laver mine and its chemistry has be examined in more depth and an attempt to determine
the possible petrogenisis of the magnetite has been conducted.

1.2 Laver –The mine that disappeared


It all started in the end of the 1920’s when prospectors working for Boliden Minerals AB found
boulders containing chalcopyrite. In the early years of the 1930’s the source of these boulders
was located on the southern slope of the so called “Gråberget”. The test drillings made on the
site showed a low grade of copper, but with the increasing prices (and demand) for copper
during the pre-WWII era’s military rearmament, Boliden decided to move forward with the
plans for a new mine (Holmberg and Korva, 2006 and references therein).
3
The old Laver mine was active 1938-1946. An entire community was built close to the mine
and was considered to be among the most modern mining towns at that time, with both
electricity and remote heating. After the closure of the mine the village was abandoned and
today there are only few ruins left to mark where community was located (Holmberg and Korva,
2006 and references therein). In recent years, new prospecting of the area have revealed a new
potential site for mining a low grade porphyry deposit (Boliden AB, 2012).

Both the old and the new Laver deposit are located in the Fennoscandian shield, in the south-
west part of the Karelian craton to be more correct, and at an approximately distance of 130 km
from the Skellefte District (Allen, 1996; Weihed 1992). The old mine itself is located roughly
120 km west of Luleå in the county of Norrbotten (Holmström et al., 1999; Ljungberg and
Öhlander, 2001) see Figure 2. The concerned study area is situated around the closed Laver
mine and stretches out in the
ice flow direction. Both the
study area and the collection
sites for the till samples was
selected by Boliden Minerals
AB prior to the start of the
project. The foundation of
their selection of collection
sites were mainly based on
maps over the area.

The Old Laver mine were


calculated to have contained
about 1.54 million tonnes
(Mt) of ore and totally were Figure 2. Generalized geologic map over the major tectonic units in the
roughly 1.3 Mt mined with Fennoscandian shield. BB and NC stands for Bothnian Basin and Norrbotten
Craton respectively. The dotted line represent the boundary between the ɛNd-
an average grade of 1.5% Cu, signatures. Laver is marked by the small red square (modified after Bauer et
al., 2013).
36 ppm Ag and 0.2 ppm Au
(Holmström et al., 1999; Ljungberg and Öhlander, 2001; Boliden AB, 2012). The richest parts
of the deposit are comprised mainly of brecciated ore that have been formed though epigenetical
processes. The main ore consist primarily of chalcopyrite and phyrrhotite with minor amounts
of sphalerite, pyrite, galena and arsenopyrite (Du Rietz, 1944; Holmström et al., 1999;
Ljungberg and Öhlander, 2001)

4
The New Laver mineralization in contrast, occurs mainly as impregnation or concentrated in
veins and is considered to be a so called Cu-Au-Mo type. The deposit is calculated to have a
large volume with a low grade. The host rocks of the New Laver deposit are believed to consist
of approximately 1.9 Ga old porphyries of felsic to mafic composition. The majority of the
bedrock are interpreted as dacitic to andesitic feldspar-porphyries with a fine grained matrix.
Volcanoclastic rocks are often associated with the porphyries, in addition are some sedimentary
rocks present in the area. The porphyries that are hosting the mineralization are considerably
altered by silification and further are the rocks penetrated by quartz veins (Boliden AB, 2012).

1.3 Regional geology


Major parts of northern Sweden have been subjected to regional metamorphism of varied
degree and most of them are
related to the Svecofennian
orogeny and the associated
magmatic processes that took
place roughly 2.0-1.85 Ga ago
(Perdhal and Frietsch, 1993).
The northern part of Sweden are
primarily dominated by rocks of
Paleoproterozoic age and the
most common bedrock in the
study area are considered to be
granitic rocks (Lundqvist et al.,
2011), see Figure 3 for a general
representation of the geology in
northern Sweden.

At the start of the Svecofennian


orogeny were the evolution of
the Baltic Shield considered to
be closely tied to the rifting of
the Archaean craton. The rifting
Figure 3. Geological map over northern Sweden (from Perdhal and
resulted in a passive margin Frietsch, 1993) The rectangle represents the study area.
were turbiditic and volcanic material was deposited. These rocks are considered to belong to
the Lapponian and Karealian Supergroups (Öhlander et al. 1999; Perdhal and Frietsch, 1993).

5
Subduction beneath the Archean craton contributed most likely to the volcanic activity during
the Svecofennian orogeny and thus created a vast amount of both volcanic and plutonic rocks
within the edge of the craton. These activities also resulted in rocks with different source
material that are either intermixed with the Archaean crust material or with rocks with a more
primitive subduction signature (Öhlander et al. 1999; Mellqvist et al. 1999). These volcanics
represent one of the few coherent areas in the whole Baltic Shield that are the product of felsic
to intermediate volcanism (Perdhal and Frietsch, 1993).

South of Laver lies the Skellefte sulphide ore district which is believed to have been created by
a volcanic arc that divided the north of Sweden into two different segments. The southern part
are interpreted to be dominated by mafic metavolcanics and marine metasediments belonging
to the Bothnian basin, while in the northern parts, large terrains of terrestrial volcanics known
as the Arvidsjaur formation occur (Öhlander et al. 1999; Mellqvist et al. 1999).

Whereas the submarine volcanic rocks in the Skellefte District are known to host VHMS-type
deposits, the area north of Laver is renowned for the apatite iron ores (e.g. Kiirunavaara and
Malmberget) and copper-deposits of porphyry and IOCG-type (Bergman et al., 2001). In
contrast, the Laver deposit is often associated with porphyries of a more continental character.
The region around Laver are alleged to be dominated by subaeially deposited felsic to
intermediate, 1.88-1.86 Ga metavolcanic rocks (Allen, et al., 1996; Kathol et al., 2012).

These metavolcanic rocks are considered to belong to the Arvidsjaur group and were interpreted
to have been deposited near an active continental margin or a magmatic arc (Allen, et al., 1996;
Lundqvist et al., 2011). The rocks in the Arvidsjaur group are broadly coeval with the Vargfors
group which are dominantly consisting of rocks with a sedimentary and submarine origin. The
Vargfors group is considered to overlie the ore-hosting volcanics of the Skellefte ore district
(Allen, et al., 1996).

The old Laver mine is situated in these metavolcanic rocks (Ödman, 1943) and the study area
is stretching out in the estimated ice flow direction (Boulton et al., 2001). The 19 collection sites
have been placed both above and below the ice flow direction. A more detailed description of
the samples is found in the appendix. Figure 4 is showing a map over the study area together
with the numbered collection sites. The yellow diamond is representing the old Laver mine and
the black arrow represents the estimated of ice flow direction from Boulton et al (2001).

6
Figure 4. The map over the Laver area shows the locations of the collection sites for the till samples (pink dots).
The yellow diamond represent the location of the old Laver mine. The black arrow exhibits the estimated ice flow
direction from Boulton et al (2001).

Further is the Laver deposit located near a geochemical boundary, which are thought to separate
Proterozoic rocks with Archaean crustal signature in the north and the more juvenile rocks in
the south. The boundary is defined by difference in ɛNd-signatures between the rocks where
the rocks in the north are carrying negative ɛNd values while the younger rocks in the south are
exhibiting positive ɛNd signatures (Mellqvist, 1999).

1.4 Previous studies


According to Ödman (1943), the old Laver deposit is hosted by the Arvidsjaur group and the
richest parts of the deposit consisted mainly of epigenetic brecciated ore of Proterozoic age.
Ödman (1943) stated that the ore consisted of chalcopyrite and pyrrothite and a study of the ore
made by Ödman in 1945 uncovers a small mineralization of Ni-Co-Ag inside the main ore. Du
Rietz (1944) presented a description of the bedrock in the area where there are a distinction
between “liparites” in the NE parts of the study area and banded tuffs, agglomerates and granite
porphyries in the south-west. Du Ritz (1944) also describes rocks of dacitic and andesitic
composition together with (steeply dipping) greenstone dykes that cross-cuts the “liparites”.

7
The descriptions of Kathol et al. (2012) are of a more regional character, which is presented in
Figure 5. The black rectangle represents the location of the old Laver mine and the large purple
box signifies the selected study area. Kathol et al. (2012) claim that the surrounding area of the
Laver deposit are predominated by granitoids of the Svecokarealian Perthite-monzonite suite
(1.85-1.80 Ga) and these rocks are intruded by granitoids of the late- or post Svecokarealian
Edefors suite (1.81-1.78 Ga).

Figure 5. Generalized map over the bedrock in the area around the old Laver mine. The black square signifies
the location of the old mine and the large purple box shows the study area (modified from Kathol et al., 2012).

The Perthite-monzonite suite are according to Kathol et al. (2012), roughly the same age as the
Arvidsjaur group and they might be co-magmatic. Both types of rocks have been less affected
by deformation and metamorphism during the Svecokarelian orogeny (1.85-1.80 Ga) compared
with the rocks of the Skellefte District. Most of the deformation of the Laver area have been
interpreted by Kathol et al. (2012) as localized deformation of nearby regional shear zones.

8
2. Magnetite geochemistry
2.1 Magnetite -a close-up
Magnetite has an inverse cubic spinel-type structure with a stoichiometry of XY2O4, where X
represents a divalent cation such as Ni, Co, Fe2+. The Y site represents a trivalent cation such
as Fe3+, Cr, or Al. Provided that there are coupled substitution in order to maintain the charge
balance, it is possible for Ti4+ to occupy the Y site. The inverse cubic structure have often a
fully occupied tetrahedral site, while the octahedral site is distorted and filled with all the X
cations and the remaining Y cations (Levy et al., 2011; Nadoll et al., 2014a; Dupuis and
Bedudoin, 2011).

The substitutions of divalent and


trivalent cations that could take place in
magnetite results in the different phase
transitions with other spinel group
minerals including spinel, ulvöspinel,
titanomaganetite and chromite.
Depending on which minerals that are
involved, the substitution might be
solid, partial or temperature dependent.
Figure 6 shows a simplified schematic
over the different transitions in the
Figure 6. Schematic representation of the spinel group with
spinel mineral group (Dupuis and complete solid solution shown by thick lines and partial solid
solution as thin lines (Dupuis and Beaudoin, 2011).
Bedudoin, 2011; Nadoll, 2011).

The continuous solid solution between magnetite and ulvöspinel (as well as the minerals created
through oxidation) exists when temperatures above 600°C are achieved. Moreover are coupled
substitution of Ti4+ for Fe3+ in the octahedral site and Fe2+ for Fe3+ in tetrahedral sites associated
with the solid solution to ulvöspinel and ilmenite. When reaching temperatures below 600°C is
the thermo-dynamic data of the phase transitions limited and extensive miscability gaps occur
(Nadoll, 2011, Nesser, 2009). A diminishing temperature also leads to the potential exsolution
of ilmenite out of Ti-bearing magnetite, which results in lower levels of Ti in the concerned
magnetite grain (Nesse, 2009).

9
During the cooling of a high temperature melt, combines often the ongoing fractionation with
the specific partitioning behaviour of the elements during crystallization and alters the chemical
composition of the remaining melt. The concentration of an element in a magnetite depends
inter alia on the following factors, (1) the initial concentration of the element in the melt, (2)
competition of accessible elements between different co-crystallizing mineral phases and, (3)
the partitioning coefficient of the element (Dare et al., 2014). The partitioning coefficient of a
specific element depends in turn on the following factors: composition of the melt, temperature,
pressure, oxygen fugacity (fO2) and cooling rate (Nadoll et al., 2014a).

Magnetite that is crystallizing in sulphide melts has a different trace element composition and
partitioning behaviour compered to magnetite that crystallize e.g. in a silicate melt. This is
caused by (1) enrichment of chalcophile elements while the silica melt is enriched in lithophile
elements, and (2) different mineral phases crystallizing. As silicate minerals naturally do not
crystallize in sulphide melts, the lithophile components which are present (even though the
concentration is low) will be incorporated into magnetite and controlled by the fractionation
(Boutroy et al., 2014; Dare et al., 2014). An enrichment of the lithophile elements is visible in
magnetite that are formed early in Fe-rich sulphide melts and as the crystallization process
continuous will the magnetite gradually become depleted (Dare et al., 2014).

Of the chalcophile elements that are present in the melt are only a selected few (including Ni,
Co, Zn, Sn and, Mo) incorporated in the magnetite, in amounts that depends on the composition
of the co-crystallizing sulphides (Dare et al., 2014). Due to the enrichment of chacophile and
(the subsequent) depletion of lithophile elements, shows the magnetite a sawtooth pattern when
normalized against the bulk continental crust in trace element diagrams. Comparing with data
of Fe rich melts exhibits magnetite from a sulphide rich source depleted in lithophile elements
and enriched of chalcophile elements such as Pb, Sn and Ni (Boutroy et al., 2014).

2.2 Hydrothermal Vs Magmatic magnetite


The complexity in the interactions of both geological and mineralogical character for the
hydrothermal magnetite results in the difficulty to discriminate different types of hydrothermal
magnetite from each other (Nadoll et al., 2014a). Improvements in analytical techniques have
made it possible to measure trace elements at lower concentrations and this results in easier
discrimination between hydrothermal and igneous magnetite as well as between hydrothermal
magnetite from different deposit types (Nadoll et al., 2014b; Nadoll et al., 2014a)

10
When dealing with magnetite from a porphyry deposit are the elements Mg, Ti, V, Cr, and, Co
are most useful when discriminating between a magmatic or hydrothermal origin. Ti seems to
be one of the best discriminators as a depletion in Ti is often seen in magnetite from known
hydrothermal porphyries. Only about 10% of the hydrothermal magnetite have more than 1
wt% of Ti, whereas roughly 30% of igneous magnetite holds more than 1 wt% (Dare et al.,
2014; Nadoll et al., 2014a).

In parallel seems lower V concentrations be more indicative of hydrothermal origin, roughly


50% of hydrothermal porphyry magnetite have vanadium concentration below 1000 ppm, while
only about 20% of magnetite with an igneous origin shows similar levels. Further, Cr indicates
that higher levels are present in igneous than in hydrothermal magnetite. While cobalt appears
to exhibit the opposite trend with higher levels in hydrothermal magnetite than in magmatic
type (Nadoll et al., 2014a).

The behaviour of Ni and Cr might be considered to be another important difference between


magmatic and hydrothermal magnetite. Both elements behave compatible during fractionation
of intermediate to felsic melts as their behaviour are considered to be coupled in silicate melts.
This coupled behaviour often exhibits Ni/Cr ratios of ≤1. Nonetheless, in many geological
settings that are affected by hydrothermal processes the behaviour of Ni and Cr are decoupled
and the ratio between them are commonly higher (≥1). This difference is most likely due to
higher solubility of Ni compared to Cr in fluids. These differences might thus be a possible
alternative of distinguishing magnetite originating in hydrothermal deposits from their host
rocks which may exhibit similar trace element signatures (Dare et al., 2014).

Further, hydrothermal magnetite often exhibits a depletion in high field strength elements and
most of these elements are often considered to be immobile during hydrothermal alteration (Van
Baalen, 1993). This unwillingness to move during hydrothermal processes results in discernible
lower concentrations of the immobile elements in most hydrothermal fluids. In addition, high
temperature hydrothermal fluids (about 500-700°C) are often associated with a magmatic-
hydrothermal source (e.g. porphyry and IOCG) and the magnetite that are deposited are
commonly enriched in the compatible elements Ni, V, Co, Zn, Mn and, Sn (Dare et al., 2014).

11
In low temperature environments such as retrograde Fe-skarn where magnetite are most likely
formed from low temperature fluids (<500°C) and sedimentary environments (such as BIF), the
magnetite is often depleted in the above mentioned compatible elements. The most likely cause
of this is the low solubility of these elements in fluids at low temperatures. However, the low
concentrations of trace elements makes them easy to distinguish both from magmatic and high
temperature hydrothermal magnetite (Dare et al., 2014).

Magmatic magnetite might also be separated from hydrothermal magnetite based on textural
evidence as magnetite from a magmatic source often develops oxidation-exsolution of ilmenite
and spinel lamellae. Removal of the elements Ti, Al and Mg to form ilmenite and spinel often
takes place during the oxidation-exsolution and is often accompanied by local redistribution of
Sc, Ti, Mn, Nb, Sn, Ta, and W which are more compatible in ilmenite. The microtextures ranges
from homogenous ulvöspinel grains to through trellis or sandwich intergrowths of ilmenite
lamellae and granules of ilmenite. The different textures are dependent on the varying degree
of oxidation and diffusion that takes place in the mineral (Liu et al., 2015).

Chemical variations in hydrothermal and magmatic magnetite from porphyry deposits can also
be discriminated by using the elements Mg, Al, Ti, V, Mn, Co, Zn and Ga. These elements
appears to exhibit variations in concentration, depending on which deposits the magnetite have
originated. The average concentrations of Al, Ti, V and Ga seems to be higher in igneous
magnetite than in hydrothermal magnetite from porphyry deposits, perhaps reflecting the higher
formation temperature for the magnetite and also variations in composition between host melt
and hydrothermal fluids. Studies also show that igneous magnetite created in porphyry copper
environments might be depleted in V, Mn, Co and Ga (Nadoll et al., 2014b).

3. Method
3.1 Sample preparation
In total, 19 till samples were collected in the study area both above the mineralization and down
in the ice flow direction of the area. The samples were then transported to the University of
Gothenburg for sample preparation and analysis. In order to facilitate the separation of the
sought magnetite was the material was sieved into several different fractions. The material was
then processed using a shaker table in order to concentrate the heavy mineral fraction which
would contain the magnetite.

12
The shaker table concentrates the heavy mineral fraction in sediment samples based on their
density. The basic principle is that particles with a higher density have a greater adherence to
the table than particles of a similar size but with lower densities. The material with a lower
density detach hence easier from the table surface and are ultimately washed away faster by the
flowing film of water, leaving the denser particles behind. (Gent et al., 2011).

As magnetite is a mineral exhibiting high magnetic susceptibility (Hunt et al., 1995), was the
magnetite separated from the rest of the heavy mineral fraction with a handheld magnet. The
magnetite was then mounted in epoxy pucks and polished before they were coated with carbon
for analysis in the scanning electron microscope and further chemical analysis of trace elements
in the LA-ICP-MS.

3.2 Scanning electron microscope


The scanning electron microscope (SEM) allows observations and characterization of samples
of both heterogeneous organic and inorganic materials on a nanometer (nm) to micrometer (μm)
scale (Goldstein et al., 2003). The SEM uses an electron beam to produce different signals at
the surface of a specimen that discloses information such as texture, chemical composition and
crystalline structure. The chemical composition which is the main interest in this study, is
examined by using the energy-dispersive X-ray spectrometer system (EDS) of the SEM, which
is used while analysing the energy spectrum of a specific point of a sample in order to define
the abundance of specific elements (Reed, 2005).

The basics of detecting the dispersing X-rays are based on the electron beams interaction with
the tightly bound inner shell electrons of an atom. The inner shell electrons become excited
when subjected to the electron beam and leaves the atom behind (also in an excited state). The
atom relaxes back to the original state as the outer shell electron(s) fills the gap in the inner
shell. The resulting energy differences in the electron shells are well studied and are thought to
be characteristic for each specific element and might consequently be used in identification of
minerals (Goldstein et al., 2003).

The excess energy which is released during the relaxation of the atoms can occur in two
different ways, (1) the variation in shell energies can be transmitted to another outer shell
electron and ejecting it with an specific kinetic energy, or (2), the excess energy is expressed as
photons of electromagnetic radiation which have a sharply defined energy and is measured by
the EDS system of the SEM (Goldstein et al., 2003).

13
In a pre-study of magnetite samples from both old Laver and the new Laver deposits performed
at the University of Gothenburg, sulphide inclusions were found to be a relatively common
feature in the magnetite. These sulphide inclusions give the magnetite a specific mineralogical
“signature” that are easy to distinguish from magnetite from barren rocks. The first part of this
investigation was focused on establishing whether these inclusions could be useful as an
indicator for similar deposits in this region by using the SEM-EDS.

3.3 LA-ICP-MS
Laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) is an effective
technique for measuring the abundance and distribution of trace elements. Advances within the
field of laser technology, better understanding of the ablation process and new calibration
reference material have improved the accuracy, precision and spatial resolution of the technique
since the late 1980’s (Nadoll and Koenig, 2011).

The LA-ICP-MS generates an aerosol by ablating a solid sample using a high-energy pulsed
laser beam that is focused on the samples surface. Being allowed to concentrate the laser on
small areas on the surface permits spatially resolved analysis of major, minor and trace elements
as well as isotope ratio measurements in most solid materials. The applications range from
geochemistry and biochemistry to medical and forensic studies (Hattendorf and Günther, 2014).

Laser ablation occurs when a material is exposed to a laser pulse of sufficient high energy,
density or fluence (expressed as J/cm2). The ablation results in a swift heating of the sample
material and eventually evaporates parts of the irradiated volume. The evaporated material is
excited either during the ablation or within a localized plasma formation above the ablation
crater. The spectrometer records the emission spectrum from the transient plasma and based on
these emissions it is possible to identify and quantify the elements that are present in the vapour
by using a calibration curve (Hattendorf and Günther, 2014).

A comprehensive and consistent methodology have not been fully developed for geochemical
studies of magnetite, but a somewhat accurate calibration for LA-ICP-MS for magnetite can be
achieved using multiple standards (Nadoll and Koenig, 2011). In this study the machine has
been calibrated using NIST 610 silicate glass, USGS’ GSE-1G synthetic basaltic glass and
BCR-2G basaltic glass and monitored for accuracy using the magnetite BC28 as a secondary
standard. The same epoxy mounts analysed with the SEM-EDS have been investigated with the
LA-ICP-MS and 35 trace elements were analysed.

14
However, not all magnetite samples were analysed in the LA-ICP-MS due to exceeding of the
provided budget. The samples were selected to give a generic picture of the possible variation
of hydrothermal magnetite in the ice flow direction. Obtaining a relatively acceptable statistical
reliability required that at least 100 gains in each selected sample were analysed. Gains were
chosen randomly to reflect the variation in the quota of magmatic and hydrothermal magnetite
in each sample.

The obtained data were filtered by using the computer program named “GLITTER”. The filter
uses data during a specified time interval during the analysis to recalculate the data. The
integration interval can be changed to avoid heterogeneities and large spikes in the acquired
signal (van Achterbergh et al., n.d.). The data is recalculated until the laser ablation signal is as
smooth as possible.

4. Results
4.1 SEM
While screening magnetite for sulphide inclusions only a few were found thus highlighting the
main problem with the original aim of the thesis. It appears that minimal amounts of undamaged
magnetite grains with inclusions are available in the collected till samples. The only exception
were magnetite collected from the old Laver mine. The magnetite grains was still undamaged
and exhibiting sulphide inclusions.

The results from the analysis performed by the SEM-EDS have been complied in Figure 7. The
images were taken using backscatter imaging (BSE) in order to visualize the different shades
of white, grey and black that represent the differences in atomic density. The sharp boundaries
between the phases in the pictures make it easy to recognize potential different minerals from
each other or inclusions inside the minerals.

Figure 7A shows the examined samples from old Laver. The magnetite are exhibited with a
dark grey colour. Fractures in the magnetite are visible along the boundaries of the inclusions
and through the magnetite itself. The light grey inclusions have been identified as CuFeS2
(chalcopyrite) while the bright white phases are inclusions of ZnS (sphalerite). Figure 7B-C are
from a second sample from the old Laver mine. Figure 7B shows the dark grey magnetite with
the (somewhat) dull white inclusions which are chalcopyrite. It should be noted that sulfide
inclusions are once again associated with fractures in the magnetite however this time the
fractures are slightly harder to detect.

15
The bright white inclusions in Figure 7C have been identified as arsenopyrite and the more dull
white/light grey inclusions are chalcopyrite. The fractures are more easily spotted along the
boundaries of the inclusions and through the host magnetite, thus splitting the magnetite into
smaller pieces which is visible in the right lower corner of the picture. In till sample “M3” taken
roughly 1 km from the old Laver mine, a single magnetite grain with an inclusion was found
which is seen in Figure 7D. The slightly darker gray magnetite grain exhibits one single
inclusion identified as pyrite. Once again it is possible to see fractures along the boundaries of
the inclusion and out towards the edge of the magnetite.

A B
Mgt

Chalco

Mgt
Chalco
ZnS

C D

Py
Aspy

Mgt

Mgt Chalco

Figure 7. The old Laver mine sample is presented in picture A-C. Picture A shows dark grey magnetite (Mgt)
with inclusions of light grey chalcopyrite (Chalco) and bright white sphalerite (ZnS). B) The dark grey
magnetite are once again seen with inclusions of chalcopyrite. In C) exhibits the magnetite bright white
inclusions of arsenopyrite (Aspy) and chalcopyrite that presents a slightly duller shade of white. D) Sample M3
contains one magnetite exhibiting inclusion of pyrite (Py). Small fractures are seen in the grain but they do not
reach the rim of the magnetite. Fractures throughout the magnetite grains out to the rims are otherwise seen
in pictures A-C.

16
The magnetite grains that have been examined with the SEM also exhibit some microtextural
differences between potential hydrothermal and igneous magnetite. Magnetite originating from
magmatic environments often develops lamellae of ilmenite and/or spinel due to oxidation-
exsolution reactions and diffusion of elements that occurs in the magnetite grains as previous
mentioned (Liu et al., 2015). An example of these kind of microtextures are shown in Figure 8
in the form of ilmenite lamellaes within on single magnetite grain.

The ilmenite have been identified as light grey areas and the lamellaes changes direction across
the magnetite grain. In addition, there is a small inclusion of plagioclase (dark grey) present in
the grain. The mineral in the upper right corner of the picture comprises of K-feldspar and
garnet that have been identified as a grossular.

Furtermore, inclusions of
ilmenite have been seen K-feld gnt

in several magnetite
grains in most of the plag
Ilm
samples. Lamellas which
are consisting of ilmenite
were the most commonly
observed microtextures
in magnetite grains from
all samples. The only
variation in extent was
that the lamellas covered
the specific magnetite Figure 8. BSE picture of a magnetite exhibiting lamellae of ilmenite (ilm), light
grey areas,. Furthermore there is an inclusion consisting of plagioclase (plag)
grain. and the mineral in the corner consist of K-feldspar (K-feld) and garnet (gnt).
Further are some fractures visible in the lower part of the magnetite grain.
4.2 Trace elements
The difference between hydrothermal and magmatic magnetite are not only restricted to textural
evidence. The average chemistry seems to differ between the types of magnetite, and the data
from the LA-ICP-MS and according to Nadoll et al. (2014a) might the data be used to divide
the magnetite into two groups based on the concentrations of trace elements that are present.
The used magnetite data in this thesis can be found on the accompanying CD-ROM.

17
Another possibility to discriminate between hydrothermal and magmatic magnetite is to use the
Ti vs. Ni/Cr diagram presented by Dare et al. (2014). This method is based on the assumption
that Ni has higher solubility in fluids and the Ni/Cr ratio is thus higher in magnetite from
hydrothermal settings. While in magmatic settings are Ni and Cr coupled and the subsequent
ratio would be lower (Dare et al., 2014).

In this study have a combination of the suggested cut-off values presented by Nadoll et al.
(2014a) and the Ti vs. Ni/Cr diagram presented by Dare et al. (2014) been used to divide the
analysed magnetite into the hydrothermal or magmatic group. After allocating the magnetite in
the two different groups an average trace element composition were calculated by using the
number of grains (n) analysed for each sample. The perceived hydrothermal magnetite have
been summarized in Table 1. The trace element data for all samples are found in the appendix.

When studying the main discrimination elements of magnetite; Mg, Ti, V, Cr, and, Co,
differences can be discerned in the chemical composition. The magnetite that are considered to
be of hydrothermal origin contains fairly low amounts of Mg, < 200 ppm. Also, relatively low
levels of Ti (< 500 ppm) are observed while V in contrast is generally high, >500 ppm.

In addition, the Cr that are present are showing lower concentrations than 300 ppm. The amount
of cobalt in the possible hydrothermal magnetite is also low and it should be noted that very
low values were observed in the average hydrothermal magnetite in the old Laver mine sample.
The cut-off values presented have been slightly modified from Nadoll et al. (2014a).

Other elements that might contribute to discrimination between hydrothermal and magmatic
magnetite are Ni (<100 ppm), Cu, Zn and Ga. The quantity of Ni in the analysed magnetite
grains are fairly low and low concentrations of Cu are observed in all samples except in the old
Laver mine sample. The amount of Zn that are present seem to be somewhat high while the
concentrations of Ga in the magnetite are relatively low.

The average chemical composition of the potential magmatic magnetite are presented in Table
2. When studying the main discrimination elements, the concentrations of Mg exhibits similar
amounts as seen in the perceived hydrothermal counterpart (less than 200 ppm). It seems that
the igneous magnetite generally contains a considerable larger amount of Ti and V (both > 600
ppm) than what is present in magnetite from a hydrothermal origin. The concentrations of Cr
lies above 400 ppm and seems to be the most distinguishing feature in the collected samples.

18
Table 1. Summary of the LA-ICP-MS results of the magnetite grains with a possible hydrothermal origin. The
values are presented in ppm and as an average of each analysed sample where “n” represent the number of
magnetite grains.

Old laver
Sample M1 M19 M3 M5 M7 M8 M16 M18
mine
n=7 n=7 n=28 n=23 n=18 n=20 n=19 n=24 n=22
Mg 23 32 28 47 88 91 34 16 127
Al 230 693 470 212 400 530 441 353 552
Si 111 236 222 213 373 311 16 27 860
S 201 150 156 59 N.A N.A 112 156 329
K N.A N.A N.A 33 N.A N.A N.A N.A 212
Ca 83 195 89 138 274 272 38 69 118
Sc 0.059 1.3 0.53 0.95 0.66 0.64 0.6 0.32 0.15
Ti 248 407 442 183 276 286 189 326 106
V 945 752 834 648 698 683 794 673 191
Cr 32 42 170 146 265 27 170 85 5.6
Mn 875 1431 809 610 677 748 698 825 327
Co 36 39 35 23 28 25 28 37 0.11
Ni 19 32 28 43 65 26 44 33 3.9
Cu 0.071 0.029 0.18 0.23 0.87 0.56 0.043 0.035 2.1
Zn 171 891 462 191 374 366 352 466 622
Ga 24 45 44 26 45 34 42 51 9.5
Ge 2.3 3.9 3.9 2.4 3.4 3.6 3.6 2.9 1
As 0.2 0.34 0.19 0.55 0.6 0.4 0.2 0.18 1.5
Sr 0.013 0.11 0.046 0.054 1.2 1.5 0.041 0.068 0.34
Y 0.038 0.64 0.12 0.45 0.27 0.3 0.084 0.23 0.4
Zr 0.045 0.11 3 12 5.3 0.52 13 6.1 2
Nb 0.027 0.24 0.76 0.11 0.24 0.47 0.12 0.043 0.11
Mo 0.096 0.22 0.2 0.19 0.21 0.54 0.097 0.46 0.08
Ag 0.012 0.012 0.012 0.008 0.04 0.1 0.012 0.01 0.19
Sn 0.064 2.8 5.4 0.45 2.7 1.4 1.7 1.8 0.19
Sb 0.017 0.24 0.027 0.064 0.14 0.2 0.056 0.024 0.44
Ba 1.5 1.4 0.3 0.24 0.8 1.5 1.4 0.99 1.9
Ce 0.025 0.16 0.24 0.14 0.73 0.57 0.028 0.67 0.18
Hf 0.019 0.003 0.085 0.37 0.44 0.24 0.37 0.16 0.17
Ta 0.008 0.01 0.032 0.025 0.034 0.050 0.035 0.004 0.01
W 0.043 0.097 0.023 0.049 0.85 0.62 0.049 0.014 0.14
Au 0.016 0.003 0.004 0.005 0.003 0.006 0.042 0.002 0.015
Pb 0.045 0.14 0.053 0.22 0.46 0.55 0.13 0.073 0.82
U 0.016 0.017 0.027 0.12 0.3 0.7 0.11 0.043 0.1
*Abbreviation: N.A = Not analysed

19
Table 2. The average chemical composition of magnetite with a probable magmatic origin collected with the LA-
ICP-MS. The elements are shown as an average (in ppm) of each analysed sample, with “n” representing the
number of analysis.

Sample M1 M19 M3 M5 M7 M8 M16 M18 Basement


n=32 n=21 n=29 n=30 n=39 n=37 n=21 n=31 n=8
Mg 63 124 48 40 47 77 80 29 89
Al 485 461 462 423 434 707 621 568 433
Si 488 196 378 251 173 339 396 53 345
S 240 178 161 55 N.A N.A 110 156 41
K N.A N.A N.A 26 N.A N.A N.A N.A 27
Ca 118 114 131 65 254 503 115 97 32
Sc 0.62 1.01 0.45 2.63 0.46 2.6 0.43 0.43 0.29
Ti 3038 808 295 2163 339 615 2300 380 114
V 1128 1172 808 728 898 669 935 770 730
Cr 815 555 3268 625 890 438 1252 1497 432
Mn 567 3002 794 899 483 654 788 768 567
Co 19 35 30 27 24 17 26 22 1.6
Ni 26 62 124 46 34 18 53 42 7.1
Cu 4.7 0.12 0.21 0.22 0.87 0.88 0.83 0.2 0.097
Zn 160 376 178 350 154 200 509 488 163
Ga 31 30 37 36 29 35 35 42 15
Ge 3.8 3.1 4.1 3.5 3.6 4.3 3.4 3.6 0.99
As 0.3 0.25 0.16 0.53 0.53 0.31 0.5 0.22 0.67
Sr 0.17 0.15 0.093 0.082 0.72 1.5 0.057 0.1 0.076
Y 0.38 0.18 0.05 0.26 0.86 0.67 0.51 0.41 0.35
Zr 1.37 0.45 0.66 1.8 0.92 0.78 1.7 0.15 0.35
Nb 0.31 0,063 0.11 3.1 0.38 0.77 0.25 0.066 0.024
Mo 0.2 0,25 0.15 0.35 0.42 1 0.8 0.26 0.045
Ag 0.014 0.011 0.02 0.008 0.042 0.029 0.012 0.009 0.004
Sn 0.99 0.53 1.46 2 0.7 1.7 1.4 2 0.14
Sb 0.056 0.12 0.029 0.42 0.16 0.14 0.051 0.033 0.11
Ba 1.1 0.78 13 1.1 1.3 1.5 1.2 0.33 0.93
Ce 0.31 0.13 0.089 0.24 0.61 0.5 0.17 0.34 0.48
Hf 0.052 0.021 0.15 0.75 0.39 0.063 0.066 0.007 0.27
Ta 0.017 0.024 0.008 0.79 0.057 0.91 0.058 0.004 0.01
W 0.19 0.65 0.048 0.054 0.85 0.54 0.082 0.013 0.014
Au 0.006 0.003 0.016 0.005 0.004 0.007 0.003 0.002 0.002
Pb 0.26 0.11 0.13 0.34 0.84 0.28 0.21 0.14 0.3
U 0.076 0.031 0.064 0.2 0.57 0.092 0.14 0.018 0.12
*Abbreviation: N.A = Not analysed

The amount of Cr are considerable higher for igneous magnetite than seen in hydrothermal
while the as both of Co and Ni exhibit similar low values as observed in the hydrothermal
magnetite. Further, the igneous magnetite contains relatively low levels of the elements Zn and
Ga and thus do not differ considerable from the hydrothermal versions of the magnetite.

20
4.3 Multielement variation diagrams
The data for the trace elements composition of the magnetite have been summarized by using
multielement variation diagrams, constructed by computing the average composition of each
sample. In other studies, (e.g. Dare et al., 2014) multielement variation diagrams have been
used as a tool to ascertain whether the analysed magnetite samples are of high or low
temperature origin. The values for the high temperature field in the diagrams have been taken
from Dare et al. (2014) and the analysed values have been normalized to bulk continental crust
using values presented by Rudnick and Gao (2004). The average trace element composition for
the hydrothermal magnetite is presented in Figure 9.

The old Laver mine sample resides mostly inside the high temperature field or just below. The
most conspicuous difference is the significantly lower concentration of Cr in relation to the
compatibility of the element into the magnetite. Furthermore, the hydrothermal magnetite of
old Laver exhibits considerable higher amount of Cu than the other samples that are presented.
It should also be noted that the old Laver mine magnetite seems to contain high amounts of Zr
and Hf as these elements are plotting above the high temperature field in the diagram. The
samples M1 and M19 are also plotting fairly close to the high temperature field or inside it.
Both samples exhibit a small dip in the Co content, but are not as nearly as notable as in the old
Laver mine sample. Moreover, these two samples contain much lower concentrations of Zr and
Hf than the other samples in average.

All samples are expressing roughly the same depletion of Mg content as well as the dip in the
amount of Al. The Cr concentrations whose average could be considered to be quite low in
hydrothermal magnetite are seen to be plotting inside the high temperature field for all the
analysed samples. Further it should be noted that phosphorus have not been analysed in the LA-
ICP-MS and the Pb values that were analysed have been considered to be untrustworthy and
have thus been excluded in the multielement variation diagrams.

The average trace element composition for the magmatic magnetite are exhibited in Figure 10.
The light pink field in these diagrams represents the average chemical variation in magnetite
that are considered to be of magmatic origin from the old Laver mine sample. Most elements
exhibit small differences in their concentrations except in the content of Zr and Hf which have
resulted in a wide field. The magnetite with possible igneous origin from the old Laver sample
seems to contain lower concentrations of Co than the average magmatic magnetite from the
other collected samples.

21
Figure 9. The diagrams presents the trace element composition of the hydrothermal magnetite from the analysed
till samples. The data have been normalized against bulk continental crust based on values presented by Rudnick
and Gao (2004). Further, have the elements been plotted according to their increasing compatibility into
magnetite. The light blue field signify the compositional range of magnetite from high temperature deposits. The
old Laver mine sample are present in diagram A, B and C for comparative purposes. Diagram A presents the
basement sample as well as the samples M1 and M19. Diagram B shows the samples M3, M5 and M7, finally,
diagram C exhibits the samples M8, M16 and M18 (adapted from Dare et al., 2014).

22
A
Old Laver magnetite

B
Old Laver magnetite

C
Old Laver magnetite

Figure 10. The diagrams presents the trace element composition of the igneous magnetite from the analysed till
samples. The collected data have been normalized against bulk continental crust based on values presented by
Rudnick and Gao (2004). The elements have been plotted according to their increasing compatibility into
magnetite. The light pink field signify the compositional range of the analysed magnetite from the old Laver
sample. Diagram A presents the samples M1 and M19. Diagram B shows the samples M3, M5 and M7, finally is
diagram C exhibiting the samples M8, M16 and M18 (modified from Dare et al., 2014).

23
The average trace element composition of magmatic magnetite seems to be somewhat more
varied than the hydrothermal magnetite. It seems that the most consistent characteristics that all
the samples are exhibiting are found among the more compatible elements, from Ga and
forward. The most evident difference from hydrothermal magnetite are higher concentrations
of Cr. Further, the dip in Mg are also present in the magmatic magnetite, although some samples
are containing more Mg on average than others.

The less compatible elements show a more diverse composition where some of the samples are
plotting high above the compositions of the magmatic magnetite from the old Laver mine
sample. One of the few characteristic that all the samples are exhibiting among these less
compatible elements are the peak concentration of Ge. Most of the samples have also similar
low amounts of Al that are consistent with the magnetite from the old Laver mine.

4.4 Distribution of hydrothermal magnetite


In mineral exploration, an indicator mineral is only considered useful if the collected samples
are exhibiting a decrease in the levels of the sought commodity, relative to the increasing
distance to a (un)known deposit. To examine whether the samples in this study shows this type
of decrease, the data have been plotted by using the Ti versus Ni/Cr ration of Dare et al, (2014).
This diagram seems to easily separate hydrothermal magnetite from igneous and the results are
presented in four diagrams that are compiled in Figure 11.

In Figure 11A the samples M1 and M19 are shown. These samples have been collected outside
the ice flow direction. Sample M1 was sampled above the known mineralization and M19 were
collected roughly 1 km west of the old Laver mine. The amount of magnetite that are plotting
in the hydrothermal part of the diagram are small compared to the quantity of magmatic
magnetite that are present in the other part of Figure 11A.

The samples M3 and M5 were collected close to the old Laver mine (approximately 0.5 to 1
km away respectively) and are presented in Figure 11B. The diagram exhibits a larger
concentration of hydrothermal magnetite compared to Figure 11A. The amount of hydrothermal
magnetite seems to be slightly less in the M5 sample than in M3, and this decreasing trend seem
to continue in sample M7 that was collected about 1.5 km away from the known source.

24
Figure 11. The distribution of the hydrothermal magnetite have been plotted in the Ti versus Ni/Cr diagrams from
Dare et al. (2014). The axis are in logarithmic scale to better present the chemical variation between the magnetite
grains. The dotted line are representing the assumed boundary betwixt hydrothermal and igneous magnetite. The
old Laver mine sample have been included for comparative purposes.

As seen in Figure 11C, sample M7 seems to contain less hydrothermal magnetite than sample
M8 which was located roughly 2 km down the ice flow direction. Thus, visually it seem that
sample M8 shows an increase of hydrothermal magnetite instead of the expected decrease
which is often associated with the longer distance to the source. Further, the samples that were
collected further away from the old Laver mine are seen in Figure 11D. M16 and M18 (sampled
about 6 and 8 km away from the old Laver mine) are more difficult to visually determine
whether a decrease takes place or not. In order to obtain a reliable answer of a possible decrease
associated with the increasing distance to the host mineralization have the changes in the
content of hydrothermal magnetite over the distance been calculated to percentage.

The concentration of hydrothermal magnetite is presented as percentage, relative to the


increasing distance in Figure 12. Samples M1 and M19 that were collected above and beside
the Laver mine contain some of the lowest concentrations of hydrothermal magnetite. While
the highest amount is found in the old Laver mine sample. The decline in the percentage of
hydrothermal magnetite is seen in the samples closest to the mine including sample M7.

25
There seem to be a slight increase in the concentration of hydrothermal magnetite which is seen
in the two samples M8 and M16. The increase in the concentration of the hydrothermal
magnetite may have different causes and will be discussed later. Moreover, the concentration
appear to once again start to slowly decrease in sample M18 that is located furthest away from
the old Laver mine.

Figure 12. The concentration of hydrothermal magnetite have been recalculated and is presented as
percentage relatively to the samples increasing distance to the known mineralization.

The pattern of decreasing amount of hydrothermal magnetite in the ice flow direction, which is
seen with the Ni/Cr vs Ti diagram might also be illustrated when plotting the ratio between V
and Cr. This diagram is based on the statement of Nadoll et al. (2014a), that high Cr and V
concentrations gives a “magmatic signal” and low levels of Cr and V indicates a hydrothermal
origin. When the magnetite from the old Laver mine sample were plotted in a V vs. Cr diagram
it was evident that magnetite was divided into two groups.

When comparing the data plotting in the Ti vs. Ni/Cr diagram with the V vs. Cr diagram, it was
discovered that the magnetite that plotted in the hydrothermal area of the Ti vs. Ni/Cr diagram
corresponds with the group that exhibits lower V/Cr ratios. The same connection were seen
with the magmatic magnetite in the Ti vs. Ni/Cr diagram which shows a higher V/Cr ratio.
Based on these discoveries was a dotted line place in the V vs Cr diagrams to serve as a possible
boundary between hydrothermal and magmatic magnetite. The V/Cr ratio were then plotted for
the remaining samples and is presented in Figure 13.

26
Figure 13. The diagrams are showing the ration between V and Cr and all axes are shown in logarithmic
scale to better visualize the differences in the ratios for the magnetite grains. The old Laver mine sample have
been included in all diagrams for comparative purposes. The dotted line represents the possible boundary
between hydrothermal and igneous magnetite.

In Figure 13A, are samples M1 and M19 presented and these samples exhibits similar pattern
as seen in the Ti vs Ni/Cr diagram. Meaning, lower amounts of possible hydrothermal magnetite
than the igneous type. The data from the old Laver mine sample have been included in all
diagrams for comparative purposes. The samples M3 and M5 are presented in Figure 13B. The
hydrothermal quota seem to be larger than seen in Figure 13A. Furthermore, the amount of
hydrothermal magnetite seem to be somewhat lower in the M5 sample than in M3.

The decreasing trend seem to continue in sample M7 in Figure 13C. The slight increase of the
hydrothermal content is once again seen in sample M8. The samples M16 and M18 (seen in
Figure 13D) exhibit similar features as previous seen in the Ti vs Ni/Cr diagram. Referring to
the slightly higher concentration of hydrothermal magnetite in the sample M16 and the
indicated decrease in the sought mineral content in M18. Overall exhibit the V versus Cr plot
similar features with the decrease of the concentration of magnetite with hydrothermal origin
and the slight increase in samples M8 and M18 as seen in the Ti versus Ni/Cr diagram.

27
The distribution between magmatic and hydrothermal
magnetite is somewhat difficult to ascertain by the
naked eye in the diagrams in Figure 13 and to illustrate
the difference in distribution have it been summarized
as percentage in Table 3. The samples collected above
the old Laver mine contain lower concentrations of
hydrothermal magnetite than the samples collected in
the ice flow direction. The highest concentration is
found in the old Laver mine sample (80% hydrothermal
magnetites). A decline in the hydrothermal content
relatively to old Laver mine is seen in samples M3, M5 and M7 as well as the small increase
and decrease in the samples M8, M16 and M18.

4.5 Discriminant diagrams


Discriminant diagrams have often been used in geochemical studies in order to classify rock
types based on major element composition (Rollinson, 1993). This have now been extended to
include trace element composition in minerals. Dupuis and Beaudoin (2011), is one example
were the authors attempts to ascertain possible deposit types in which the minerals have
originated. However, the reliability of these diagrams are still being studied and should be
considered with caution. Neverthless, an attempt to ascertain possible deposit type based on the
trace element composition of the magnetite have been conducted and is presented in Figure 14.

Figure 14. Ti+V versus Ni/(Cr+Mn) discriminant diagram exhibiting a possible origin for all the magnetite samples that have
been analysed. The solid lines are somewhat better defined than the dotted lines (revised from Dupuis and Beaudoin, 2011).

28
The magnetite data plotted in the discriminant diagram have been recalculated from ppm to
weight percentage (Wt%). The different fields inside the diagram have been based upon values
used by Dupuis and Beaudoin (2011). The fields that consists of dotted lines instead of solid
lines, indicate that they are less defined. According to Dupuis and Beaudoin, (2011) experience
these less defined regions in the diagram often overlap between the different deposit types.
Mostly due to similar Ti+V concentrations which might be present in minerals from for example
BIF and porphyries.

As seen in the Ti versus Ni/(Cr+Mn) diagram are the majority of the analysed magnetite plotting
in the area below the solid lines, indicating an skarn origin. A small part of the magnetite gains
are exhibiting higher ratios, between Ti and Ni/(Cr+Mn) which is consistent with magnetite
that have originated in deposits that contain Fe-Ti and V ores. A minor amount are plotting
inside the areas of IOCG and porphyry.

The Ti+V versus Ca+Al+Mn discriminant diagram is often used in combination with the Ti+V
versus Ni/(Cr+Mn) diagram. This diagram might be useful to easily distinguish Fe-oxides
originating in BIF deposits from those that was created in skarn, opemiska type and Archean
porphyry deposits (Dupuis and Beaudoin, 2011). The dotted lines of the BIF field again
representing a less well defined area. In this case it is assumed that the BIF field is sharing the
area with the opemiska deposit type.

The Ti+V versus Ca+Al+Mn diagram is shown in Figure 15. The majority of the magnetite
data are plotting in the area of IOCG instead area of skarn. The same magnetite grains that
plotted in the Fe-Ti, V area of the Ti+V versus Ni/(Cr+Mn) diagram are still plotting in the
same area in the Ca+Al+Mn version of the discriminant diagram. Furthermore, some of the
magnetite are located in the areas that belongs to the deposit types of porphyry and Kiruna.

A relatively large part of the magnetite grains are plotting inside the undefined area of BIF in
the diagram. This is a notable difference compared to the Ti+V versus Ni/(Cr+Mn) diagram,
where almost no magnetite were plotting as BIF. It should also be noted that a few outliers are
present in the diagram, these are located in the uppermost part of the discriminant diagram.
Those magnetite grains that falls outside of the BIF area and below the solid line of skarn might
be of either opemisk or Archean origin.

29
Figure 15. Ti+V versus Ca+Al+Mn discriminant diagram presenting an alternative origin for the analysed magnetite grains
that have been analysed. The solid line represent the slightly better defined areas in the graph and the dotted lines are
representing the less well defined area of BIF (modified from Dupuis and Beaudoin, 2011).

4.6 Laver versus Kiruna


As the area north of the Laver deposit are renowned for the iron-apatite ores are differences in
magnetite chemistry (originating for example Kiruna) of vital importance to separate them from
magnetite from Laver. In this study have trace element analyses of known magnetite from the
Kiruna deposit been provided (by fellow Master’s student Christian Karlsson) for comparison
to the basement sample of the old Laver mine. The composition of the trace elements have been
normalized to the bulk continental crust and plotted in a multielement variation diagram, seen
in Figure 16 together with the average magnetite composition from the old Laver mine.

Figure 16. The multielement variation diagram exhibits the differences in chemistry between hydrothermal
and igneous magnetite from Laver compared to analyses performed on magnetite from the Kiruna deposit.
The light green field represents the chemical variations that is exhibited by the hydrothermal Kiruna
magnetite. The average composition for the hydrothermal magnetite from the old Laver mine is represented
by the black line and the igneous magnetite is presented by the red line (modified from Dare et al., 2014).

30
The light green area in the diagram represent the variation in the trace element composition
between the analysed magnetite from Kiruna. Several differences are visible between the
magnetite from Laver and Kiruna. Firstly, the magnetite from the old Laver mine contains larger
amounts of Al in both the igneous and the hydrothermal magnetite, than what is present in
magnetite originating from the Kiruna deposit. Secondly, the elements Ga, Mn, Mg, Ti and Zn
are present in slightly higher levels in the old Laver mine sample than in the Kiruna magnetite.

Thirdly, the analysed Kiruna magnetite contain higher concentrations of Co than seen in both
igneous and hydrothermal magnetite from the old Laver mine sample. Nevertheless, differences
in concentration between igneous magnetite from Laver and the Kiruna magnetite are typically
less pronounced, than compared to the hydrothermal magnetite from Laver. Finally, the main
difference between magnetite from Laver and Kiruna is the extremely low concentration of Cr
that is present in the Kiruna magnetite. While both igneous and hydrothermal magnetite from
the old Laver mineralization exhibits higher concentrations of the element.

To further display the differences in chemical composition and the different environments that
the magnetite have originated in between the two deposits discriminant diagram have been
applied. The values for the Kiruna magnetite have been recalculated to Wt% and was plotted
together with the values for the magnetite from the Laver deposit.

The recalculated data from the analysis of the provided Kiruna sample have been plotted in a
Ti+V versus Ni/(Cr+Mn) diagram, see Figure 17 below. The data of the old Laver mine sample
have also been included and the difference in indicated deposit type seem to be quite obvious.
The old Laver mine magnetite is plotting as seen before in the less defined area of skarn while
the Kiruna magnetite is located in the field belonging to the IOGC deposit type or just along
the border to the area indicating a Kiruna type deposit.

31
Figure 17. The Ti+V versus Ni/(Cr+Mn) discriminant diagram are used to illustrate the difference in the estimated settings
of the magnetite from both Laver and Kiruna. The values of the magnetite from Kiruna is shown as grey circles and the Laver
magnetite is exhibited as black circles (remade from Dupuis and Beaudoin, 2011).

5. Discussion
5.1 Limitations for sulphide inclusions as tracers for mineralization
As mentioned earlier is the minimal amount of undamaged gains containing sulphide inclusions
in the samples collected downstream in the ice flow direction of old Laver mine the primary
cause that the original aim had to be abandoned. The sulphide inclusions themselves might hold
the answer to why they was not observable in the SEM.

As Blatt et al (2005) explains are mineral grains that contains inclusions (of any kind) weaker
than grains without any inclusions. This weakness exists because minerals are composed of two
different, distinct phases with internal discontinuity surfaces (crystal boundaries) between
them. Exposing these crystal boundaries to any type of stress might result in fractures along
these zones of weakness and eventually lead to the destruction of the grain.

According to Boulton et al. (1974), is material that was picked up by the ice sheet during the
advancing stage of the glacier subjected to shearing from the movement, which could have
resulted in the crushing of the material. The weak boundaries between the magnetite grain and
the inclusion would probably be prone to fracturing and crushing. In combination with shearing
could the existing inclusions weaken the integrity of the magnetite and this is most likely the
main reason to the difficulty to find undamaged magnetite grains with inclusions still intact.

32
The grains were most likely subjected to large amount of stress during the transportation of the
ice. The result was presumably the many fractures along the internal discontinuity surfaces that
are represented by the inclusions such as CuFeS2, ZnS and other identified mineral phases. This
is visible in the Figure 7 where a large number of fractures are exposed along the boundary
between the magnetite and the inclusions and all the way out to the rim of the grain. Some of
the inclusions have even been separated from the host magnetite.

The sample M3 yielded the only positive identification of inclusion in a magnetite of all samples
that was examined outside of the basement sample. This finding can be viewed as an indication
that magnetite with sulphide inclusions (perhaps) from the old Laver mine, might survive a
transport distances of roughly 0.5 km by an ice sheet. The magnetite was found in the 250-180
µm fraction and it is a slight possibility that there might more of these particular inclusion-rich
magnetite in sample M3. Possibly also in the area close to the old Laver mine if large amounts
of magnetite grains in the 250-180 µm fraction and the other smaller fractions are investigated.

Although sample M3 were collected close to the deposit, the magnetite has still been subjected
to stress from the ice (during a shorter amount of time). The shorter exposure time have given
the magnetite a slightly better chance of survival than a grain transported a greater distance.
Being subjected to stress under a longer amount of time increases the likelihood of cracking
along the crystal boundaries. This is a plausible reason to the difficulty to find magnetite with
sulphide inclusions outside of the old Laver mine.

5.2 Compositional variations between hydrothermal and igneous magnetite


As stated by Nadoll et al (2014b) igneous magnetite have often higher Al, Ti, V and Ga
concentrations than hydrothermal magnetite from a porphyry source. In the analyses performed
in this study are amounts of Al, Ti and V higher in the perceived igneous than the hydrothermal
magnetite. However, the amount of V that is present in both types is high but is slightly higher
in the magmatic magnetite and the Ga concentrations are exhibiting almost the same range.

According to Nadoll et al (2014a) are Mg, Cr and Co other elements that are considered to be
useful discrimination elements. The amount of Mg in both types of samples is on average low
and the igneous magnetite shows slightly less variation. However the Mg content is not reliable
enough to rely solely on for discriminating between hydrothermal and magmatic grains.

33
Further exhibit Co concentrations that is slightly lower in magmatic magnetite, which is
common trait according to Nadoll et al (2014a). The most striking difference in the measured
concentrations is the significantly lower amount of Cr in the hydrothermal magnetite compared
to the igneous.

In the study performed by Nadoll et al (2014a) the average hydrothermal magnetite have a mean
Cr concentration of 370 ppm and the magmatic magnetite contain higher concentrations than
the hydrothermal magnetite. The average Cr concentration for hydrothermal magnetite in this
study do not even exceeds 50 ppm for any of the samples. In comparison the mean value for
the igneous magnetite varies roughly between 400 ppm to over 3000 ppm.

According to Dare et al (2012) Cr sensitive to the fractionation that takes place in the host
magma. As fractionation occurs, the concentrations of lithophile elements (which includes Cr)
decreasing in the Fe-oxides that are created. During the time when the hydrothermal magnetite
in the Laver area were created, the host magma might already had achieved an advanced
fractionation state. This might be the cause that only magnetite with low concentrations of Cr
were available for transportation and later precipitation by hydrothermal fluids.

Nadoll et al (2014a) claims that the first order of control for concentrations of elements in
hydrothermal magnetite are the availability of the concerned elements. This leads once again
back to the concentrations in the host magma and how the fractionation affects which elements
that are obtainable for the hydrothermal fluid. The original composition of the magma would
most likely affect how much of an element that is accessible for fractionation as the magma
cools and the hydrothermal fluids starts transporting elements.

Further have the prevailing temperature in the magma been observed to affect the partitioning
coefficients of the elements. A lower mean temperature results according to Nadoll et al (2014a)
in hydrothermal magnetite with lower trace element concentrations. However, it is only the Cr
concentration in the hydrothermal magnetite that are very low. One explanation is a
combination of a Cr poor host magma that fractionated heavily before the hydrothermal fluids
started to transport the available elements.

34
The theory seems to be supported by the multivariation diagrams in Figure 9 that shows that
the Cr concentration are still plotting in the areas of the high temperature field, together with
(most of) the other elements that are present in the hydrothermal magnetite. The amounts of Zr
and Hf that are plotting above the high temperature field are most likely due to inclusions in a
few magnetite grains that could be considered as outliers. These diagrams seem to reflect the
trace element composition of the hydrothermal magnetite well. The samples down the ice flow
direction exhibits similar compositions which is coinciding with the high temperature field
while the samples M1 and M19 are indicated to have been created during lower temperatures.

It should be noted that the area were Laver is located have been subjected to metamorphism of
varying degree according to Perdhal and Frietsch (1993) along with the rest of northern Sweden.
This event might have caused an overprint in the chemical composition of the magnetite. A
possible overprint of the magnetite might cause an originally magmatic magnetite to exhibit a
similar trace element composition as a hydrothermal magnetite (or vice versa). Alteration of
the chemical composition after the creation of the mineral makes the discrimination between
magmatic and hydrothermal magnetite even more complex.

5.3 Hydrothermal magnetite as an indicator mineral


By correlating a changing ratio between hydrothermal and igneous magnetite in samples that
are located on an increasing distance from a mineralization, have Dare et al. (2014) shown that
it might be possible to use this as an indicator mineral for mineral deposits. In this study have
the same Ti vs. Ni/Cr as presented in both Dare et al. (2014) and Huang et al. (2014) been used
to examine if there are a decrease in the hydrothermal magnetite content in till samples that
befits a indicator mineral.

The characteristic of an IM is the gradual decrease in the ice flow direction as stated both by
Larson and Mooers (2004) and McClengahan (2005). The Ti vs. Ni/Cr diagram exhibits few
magnetite grains of hydrothermal origin when plotting the samples M1 and M19 as it should.
These two samples was, as previous mentioned collected above and the roughly 1 km west of
the old Laver mine as reference samples. Showing how the assembly between magmatic and
hydrothermal magnetite in till, not containing material from the deposit might look like.

As stated by Pluoffe (1996) is indicator minerals useful until it is no longer possible to


distinguish the anomalous concentration from the background value. The background signal
were not completely identified as both the references samples contained small amounts of
hydrothermal magnetite.

35
More reference samples should have been collected further away from the old Laver mine to
determine the real background signal in the area. The bedrock in the area might also effect the
background signal some magnetite might have originated from the surrounding bedrock and
not the old Laver mine. This influence the reliability of using indicator minerals and the results
should be considered with some caution.

The diagrams seen in Figure 11 that are showing the samples collected in the ice flow direction
seems to, visually, decrease in the amount of hydrothermal magnetite compared with their
igneous counterpart. As seen in other studies of other types of indicator minerals (e.g.
McClengahan et al., 2000) is large initial amounts of the sought mineral found in samples
gathered close to the source and at greater distances the concentration diminish.

When examining compositional variations between hydrothermal and magmatic magnetite in


the old Laver mine sample a difference between the concentrations of V and Cr were noted.
When plotting the elements against each other it was seen that the data were divided into two
clusters. These groups could be linked with the two groups in the Ti vs. Ni/Cr diagram were the
data from one cluster corresponds with the data exhibited as either hydrothermal or magmatic.

Plotting V versus Cr for all the samples presents a similar trend as exhibited by the Ti vs. Ni/Cr
diagram. The amount of hydrothermal magnetite seem to decrease in the ice flow direction. To
clarify whether there really is a decrease have the change in percentage for both types of
diagrams been calculated for each sample. As expected constitutes the hydrothermal magnetite
only small parts of the analysed amount of magnetite in the samples M1 and M19. Further, the
line is showing a negative trend as the samples further away from the source are presented.

However, from sample M8 and continuing to M16 is there a small increase in the amount of
hydrothermal magnetite and the curve starts once again to diminish in sample M18. This
increase may be attributed to an unfortunate sampling in another till layer, which were created
during one of the many other glaciations that have affected Scandinavia during the last 2.5 Ma
years according Wohlfarth et al (2008). This layer might contain higher levels of hydrothermal
magnetite in respect to the amount igneous that are present. As in Larson and Moores (2004)
have the expected distribution of material been based on an assumed steady state glacial
erosion, which is not always the case when dealing with natural processes. The increase of
hydrothermal magnetite in the distal part of the sampling area might simply be attributed to an
uneven deposition by the ice sheet.

36
Although there is a slight increase in the concentration of hydrothermal magnetite. It is still a
smaller amount that is present compared to the samples that are located closer to the old Laver
mine. Thus, it seems that using the Ti vs. Ni/Cr diagram and hydrothermal magnetite as a
method might be beneficial in this case. The samples collected in this study exhibits the typical
decrease in indicator mineral with the distance. The V vs. Cr diagram needs on the other hand
to be exposed to more thorough examination before it could be considered credible but it seems
to exhibit some potential.

5.4 Discerning the deposition type


Discerning the deposit type of the magnetite present in the till samples might give a hint of the
host deposit. Further it might be helpful if there are several different types of deposits in the
area with similar chemical fingerprint. The Ti+V vs. Ni/(Cr+Mn) diagram have been used e.g.
in studies of Nadoll et al (2014a), Huang et al., (2014) and Dupuis and Beaudoin (2011) to
ascertain a possible deposit type for the analysed magnetite. The fields for IOCG, Kiruna,
porphyry-Cu and Fe-Ti and V are somewhat well defined. Because magnetite from these types
of deposits often do not exhibit extensive variation in either ratios that are used. However, the
data that was collected during this study mostly plots in the undefined area of skarn, BIF,
opemiska type and Archean porphyry.

As Dupuis and Beaudoin (2011) explains is the combination both the Ti+V vs. Ni/(Cr+Mn)
with the Ti+V vs. Ca+Al+Mn discriminant diagram important. This seems to facilitate the
distinction between iron oxides from BIF environments from those of skarn, opemiska and
Archean porphyry deposits. According to Liu et al (2015) it is not uncommon that data from
magnetite porphyry and BIF origins are plotting outside the suggested fields. Other types of
diagrams and combinations of diagram have been suggested by Nadoll et al. (2014), however
no extensive studies examining the reliability of these diagrams have been conducted and they
should be viewed with a critical mind.

The Laver mineralization have previously been considered to be a porphyry Au-Mo-Cu deposit
and these types would be expected to plot in the porphyry field of the discriminant diagrams
provided by Dupuis and Beaudoin (2011). However it is seen in Figures 14-15 that only a few
magnetite grains are plotting in the porphyry field and none of them belongs to the known
sample from the old Laver mine. As stated by Dupuis and Beaudoin (2011) have Phanerozoic
porphyry copper deposits higher Ti+V ratios than the data collected from Archean porphyries.
Based on these observations might it be possible that the low Ti+V ratios of the magnetite from
the old Laver mine indicates that they originated in an Archean porphyry.
37
5.5 Differences between Laver and Kiruna magnetite
The ability to separate magnetite from different and perhaps already known deposits in the area
(and at a distance) is as previously explained important. In this study might the possible
presence of magnetite from the large iron-apatite deposits north of Laver to be considered as a
source of contamination. Fortunately is the chemistry of the provided data of the hydrothermal
Kiruna magnetite quite different from the Laver deposit.

The main difference are the higher amount of Co and the considerable lower concentration of
Cr that are present in the Kiruna samples. This seems to indicate that there are differences in
the composition of the host magmas of the different deposits. Further might differences in the
fractionation during the creation of different minerals have affected the amount of available
litophile elements. The Kiruna magnetite is mostly plotting below the high temperature field of
the multivariant diagram. Which leads to the interpretation that the hydrothermal magnetite of
Kiruna might have been created in lower temperatures than the old Laver magnetite.

Another difference is the suggested deposit type when plotting the supplied data in a Ti+V vs.
Ni/(Cr+Mn) diagram. The magnetite from Kiruna exhibits slightly higher Ti+V ratios than the
Laver magnetite. In addition are the Kiruna samples showing considerable higher Ni/(Cr+Mn)
ratios than what are presented in the magnetite from the Laver deposit. The ratios locates the
Kiruna sample in the IOCG field of the diagram while the Laver sample are plotting as
mentioned before in the undefined field of skarn. These diagrams are, as already mentioned,
not full credible when plotting data from porphyry or BIF deposits and this difference should
be viewed critically and differences in the chemical data should prevail.

The chemical differences between these two deposits confirms further what have been shown
in studies by Nadoll et al (2014a), Huang et al (2014) and Dupuis and Beaudoin (2011). Using
geochemical data from different deposits and plotting trace elements against each other seem
to be a reliable method to discriminate between different deposits as long as the elements that
are used are of relevance.

5.6 Sources of error


One example of a possible source of error is the sampling of the till samples might have taken
place in different till layers. As there have been several glaciations that have occured in the area
it is hard to determine which layer that corresponds to which glaciation. Samples might also
have been contaminated by another layer(s) during sampling as material from overlying layers
fell down during the sampling.

38
Another source of possible errors is the shaker table that was used to enrich the till samples in
magnetite and separate the heavy mineral fraction from the lighter minerals. Once again have
the effectiveness of the shaker table been proven that it is highly dependent on the skills of the
operator, mostly due to the difficulty to optimizing the settings for the different grain sizes.
Still, the ability of processing large quantities of material easily and relatively fast gives the
shaker table an advantage of other more time consuming methods.

The shaker table might also be another source of contamination but this time between samples.
Because after some time operating the shaker table it was observed that some material from a
sample stuck on the table despite cleaning the shaker table before a new sample was processed.
The contamination were minimal as this problem were revealed and rectified early using a towel
to dry the shaker table by hand between processing of the samples. Further, when collecting the
material from the shaker table some of the material were swept away by the water before it
could be collected. However the loss of this material are considered to be negligible.

The source of error that originates from the LA-ICP-MS is firstly from the analysing of the
magnetite samples. Some of the grains might contain small inclusions of other minerals or have
been subjected to processes that have resulted in the development of lamellaes. This might give
a misleading analysis. Minimal fractures and cracks in the magnetite might also contribute to
ambiguous analysis. Sometimes was the timing off when analysing, rendering the concerned
analysis worthless. However these specific analyses were filtered out when reviewing the
acquired data and subtracted from the total grains that was analysed.

Secondly is the filtrating of the raw data another source of potential error. When the data is
filtrated are the values calculated based on a specific given time interval during the analysis.
Sometimes it was difficult to find a large enough area that gives a fair representation of the
grain that was analysed. Further it was sometimes not possible to filter enough times to make a
smooth analysed line. However, to filter to many times may also contribute to data that are
showing an erroneously concentration of the concerned element.

39
6. Conclusions
Magnetite containing sulphide inclusions in magnetite does not appear to be able to survive
longer transport distances than 0.5 km under stress. The most likely cause to their destruction
is the weaker areas along the boundaries of the inclusion and the host mineral. The reason for
using uncommon (or any other type) inclusions in minerals as a cause to implicate that they
might be used as an indicator mineral is faulty at best. The conclusion in this thesis is thus,
using inclusions as tracers is not a useful method when prospecting.

The compositional variations that are present between hydrothermal and magmatic magnetite
have been decided by comparing the content of trace elements in magnetite. Using the more
common discrimination elements, exhibits the igneous magnetite often higher concentrations
of Al, Ti and V than the hydrothermal magnetite. The amount of Ga that is present shows
approximately the same range for both igneous and hydrothermal magnetite and thus, is not a
discriminant elements that could be used with reliability.

The very low concentrations of Cr that is present in the hydrothermal magnetite is the most
striking chemical feature. One of the main controls on the concentrations of an element is the
availability. Which in turn is effected by other factors such as (1) the fractionation in the host
magma, (2) the initial chemical composition of the magma and (3) the temperature that is
prevailing in the area as the hydrothermal processes starts. All these factors could individually
be responsible for the low concentrations of Cr. However it could also be a combination of all
these factors that have led to the Cr poor hydrothermal magnetite that is seen in the Laver area.

The Ti vs. Ni/Cr diagram was used to examine the distribution of hydrothermal magnetite which
seemed at first sight to be a steady decline in the amount of this sought type of magnetite. An
additional diagram plotting V against Cr exhibited a similar pattern. Calculating the percentage
of hydrothermal magnetite in the samples showed first a decrease in hydrothermal magnetite
and then a small increase in the samples furthest away from the known source. Despite the
slight increase in the more distal samples it still exhibits a smaller concentration than what is
present in samples located closer to the old mine.

Using hydrothermal magnetite as an indicator mineral appears to have been somewhat


successful. This might be used as a first step to confirm whether there is something worth
examining further when searching for undiscovered deposits. In addition, it seems that the V
vs. Cr diagram is presented in this study might have potential. However, the diagram needs to
be exposed to more thorough examinations before it could be considered to be credible.

40
When discerning a possible origin of the old Laver mineralization and the magnetite that was
collected from till samples is the combination of two diagrams necessary to make any type of
assumptions. Unfortunately are the discriminant diagrams that are in use today of somewhat
questionable character, where the possible deposit type is based on the observations of the low
ratios of Ti+V. Which is considered to be a more common trait in Archean porphyries than in
the Phanerozoic types.

Separating the Laver magnetite from magnetite from the iron-apatite deposits of Kiruna and
Malmberget are important in order to show that each type of deposit have different composition,
which could be made visible using simple diagrams. In the case of the Kiruna magnetite the
main difference lies in the concentrations of Cr and Co when compared to the magnetite from
Laver. Moreover, the differences in chemical composition might be explained by the initial
chemical composition in the host magma. Another possible suspect is the fractionation of
different mineral phases which results in variations in the concentrations of elements that are
available for transport by hydrothermal fluids.

To conclude, it should be said that here are still much to discover and examine in the field of
magnetite geochemistry and all the applications that it could be used in. One example is studies
to further define the fields in the discriminant diagrams and hopefully conducted in the near
future. Moreover, additional studies of Laver might be needed to geochemically separate the
old Laver mineralization from the new deposit that are to be mined by Boliden Minerals AB.
These two geographically close mines might exhibits geochemical variation which could be of
both academically and practical interest.

It will be of interest to examine magnetite collected in bedrock outside the mineralization area
to study chemical differences of non-mineralized magmatic and metamorphic rocks and to
ascertain a more reliable background signal in the area. It has to be stressed that no data on trace
elements in metamorphic magnetite exists. The effects of detailed metamorphism on magnetite
chemistry also need to be evaluated. However these are subjects that requires further studies in
the future.

41
Acknowledgments
Firstly I would like to thank Boliden Minerals AB for the financial support and for introducing
me to the subject, special appreciation goes to Joachim Albrecht for assistance in the field.
Further, I would like to express my gratitude to my advisor Johan Hogmalm, for his time,
assistance and in particular for all the helpful and encouraging advice, without you I would not
have been able to finish this Master’s thesis. I would also like to express my very deep
appreciation to Thomas Zack for reviewing this manuscript and for his constructive and useful
comments. My special thanks goes to David Cornell for his advice concerning the SEM and
Christian Karlsson for volunteering his data of the Kiirunavaara magnetite. Special thanks goes
to my opponent Samuel Martinsson for volunteering his time in order to review a draft of this
thesis and for his constructive ideas. Finally, I would like to acknowledge the support I have
received from my family and friends.

References
Bauer. T. E., Tavakoli. S., Weihed. P., Skyttä. P., Hermansson. T., Allen. R., Dehghannejad.
M., García Jutanatey. M. A. & Juhlin. C. (2013). A regional scale 3D-model of the Skellefte
mining district, northern Sweden. Mineral deposits for a high-tech world: Proceedings of the
12th SGA Beinnial Meeting 2013, 12-15 August, Uppsala Sweden: Svergies Geologiska
Undersökning [SGU], available only at
http://pure.ltu.se/portal/files/56064142/ACR72019.6E63E_Bauer_etal_revised.pdf, retrieved
2015-05-14

Berger. B. R., Ayuso. R. A., Wynn. J. C. & Seal. R. R. (2008). Preliminary model of porphyry
copper deposits. U.S. Geological Survey [USGS], Open-File report 2008-1321, 1-55.

Blatt. H., Tracy. R. & Owens. B. (2005). Petrology. Igneous, sedimentary and metamorphic
(3rd edition). United States of America: W. H. Freeman

Boliden Minerals AB. (2012). Beräkning av mineraltillgång av Laver i Norra Sverige.


Boliden Mineral AB Press Release, available only at
http://www.boliden.com/Documents/Investor%20Relations/CMD2012/Laverrapport_SVE-
Final.pdf, retrieved 2015-05-14.

Boulton, G. S., Dongelmans, P., Punkari, M. & Broadgate, M. (2001). Palaeoglaciology of an


ice sheet through a glacial cycle: the European ice sheet through the Weichselian. Quaternary
Science Reviews, 20, 591-625.

42
Boulton, G.S., Dent, D. L. & Morris, E. M. (1974). Subglacial shearing and crushing, and the
role of water pressures in tills from south-east Iceland. Geografiska Annaler, Series A,
Physical Geography, 56(3-4), 135-145.

Boutroy. E., Dare. S. A. S., Beaudoin. G., Barnes. S-J. & Lightfoot. P. C. (2014). Magnetite
composition in Ni-Cu-PGE deposits worldwide: application to mineral exploration. Journal
of Geochemical Exploration, 145, 64-81.

Clark. D. A. (2014). Magnetic effects of hydrothermal alteration in porphyry copper and iron-
oxide copper-gold systems: A review. Tectonophysics, 624, 46-65.

Dare. S. A. S., Barnes. S-J., Beaudoin. G., Méric. J., Boutroy. E. & Potvin-Doucet. C. (2014).
Trace elements in magnetite as petrogenetic indicators. Mineralium Deposita, 49(7), 785-796.

Dare, S. A. S., Barnes, S-J. & Beaudoin, G. (2012). Variations in trace element content of
magnetite crystallized from a fractionating sulphide liquid, Sudbury, Canada: Implications
for provenance discrimination. Geochimica et Cosmochimica acta, 88, 27-88.

Dupuis. C. & Beaudoin. G. (2011). Discriminant diagrams for iron oxide trace element
fingerprinting of mineral deposit types. Mineralium Deposita, 46, 319-335.

Du Rietz. T. (1944). The alteration of the rocks in the copper deposit at Laver. Sveriges
Geologiska Undersökning [SGU] C467, 1-38.

Gent. M., Menendez. M., Toraño. J. & Torno. S. (2011). A review of indicator minerals and
sample processing methods for geochemical exploration. Journal of Geochemical
Exploration, 110, 47-60.

Goldstein. J. I., Newbury. D. E., Echlin. P., Joy. D. C., Lyman. C. E., Lifshin. E., Sawyer. L.
& Michael. J. R. (2003). Scanning electron microscopy and X-ray microanalysis (3rd edition).
Kluwer Academic Publishers, United States of America.

Hattendorf. B. & Günther. D. (2014). Laser ablation inductively coupled plasma mass
spectrometer (LA-ICPMS) in Gauglitz. G. & Moore. D. S. ed., Handbook of Spectrocopy (2nd
edition). Wiley-VCH Verlag GmbH & Co, Germany.

Holmberg. A. & Korva. N. (2006). Laver –Orten som försvann (Master’s thesis). Luleå
Technical University. Institution of Industrial economy and social science, available only at
http://epubl.ltu.se/1402-1773/2006/033/LTU-CUPP-06033-SE.pdf, retrieved 2015-05-14.

43
Holmström. H., Ljungberg. J. & Öhlander. B. (1999). Secondary copper enrichment in
tailings at the Laver mine, northern Sweden. Environmental Geology, 38(4), 327-342.

Huang. X., Qi. L. & Meng. Y. (2014). Trace element geochemistry of magnetite from the Fe
(-Cu) deposits in the Hami region, eastern Tianshan Orogenic Belt, NW China. Acta
Geologica Sinica, 88(1), 176-195.

Hunt. P. C., Moskowitz. B. M. & Banerjee. S. K. (1995). Magnetic properties of rocks and
minerals. In Ahrens. T. J ed., Rock physics and phase relations: A handbook of physical
constrains. American Geophysical Union, AGU Reference Shelf, 3, 189-204.

Kathol, B., Jörnberger, J., Kumpulainen, R. & Larsson, D. (2012). Beskrivning till
berggrundskartorna 25K Harads NV, NO, SV & SO. Sveriges Geologiska Undersökning
[SGU], K406-409.

Larson. P. C. & Mooers. H. D. (2004). Glacial indicator dispersal processes: A conceptual


model. Boreas, 33, 238-249.

Levy. D., Giustetto. R. & Hoser. A. (2011). Structure of magnetite (Fe3O4) above the Curie
temperature: a cation ordering study. Physics and Chemistry of Minerals, 39(2), 169-176.

Liu. P-P., Zhou. M-F., Chen. W. T., Gao. J-F. & Huang. X-W. (2015). In-situ LA-ICP-MS
trace elemental analyses of magnetite: Fe-Ti-(V) oxide-bearing mafic-ultramafic layered
intrusions of the Emeishan large Igneous Province. SW China. Ore Geology Reviews, 65(4),
853-871.

Ljungberg. J. & Öhlander. B. (2001). The geochemical dynamics of oxidizing mine tailings at
Laver, northern Sweden. Journal of Geochemical Exploration, 74, 57-72.

Lundqvist. J., Lundqvist. T & Lindström. M. (2011). Sveriges geologi – Från urtid till nutid
(3ed Ed). Lund: Studentlitteratur.

McClenaghan. M. B. (2005). Indicator mineral methods in mineral exploration.


Geochemistry: Exploration, Environment, Analysis, 5(3), 233-245.

McClenaghan. M. B., Thorleifson. L. H. & DiLabio. R. N. W. (2000). Till geochemical and


indicator mineral methods in mineral exploration. Ore Geology Reviews, 16, 145-166.

44
Mellqvist. C., Öhlander. B., Skiöld. T. & Wikström. A. (1999). The Archaean-Proterozoic
palaeobounadary in the Luleå area, northern Sweden: field and isotope geochemical evidence
for sharp terrane boundary. Precambrian Research, 96, 225-243.

Nadoll. P., Angerer. T., Mauk. J. L., French. D. & Walshe. J. (2014a). The geochemistry of
hydrothermal magnetite: A review. Ore Geology Reviews, 61, 1-32.

Nadoll. P., Mauk. J. L., Leveille. R. A. & Koenig A. E. (2014b). Geochemistry of magnetite of
porphyry Cu and skarn deposits in the southwestern United States. Mineralium Deposita,
2014, 1-23.

Nadoll. P., Mauk. J. L., Hayes. T. S., Koenig. A. E. & Box. S. E. (2012). Geochemistry of
magnetite from hydrothermal ore deposits and host rocks of the Mesoproterozoic Belt
Supergroup. United States. Economic Geology, 107, 1275-1292.

Nadoll. P. (2011). Geochemistry of magnetite from hydrothermal ore deposits and host rocks
–Case studies from Proterozoic Belt Group. Cu-Mo-porphyry + skarn and Climax-Mo
deposits in the western United States (Doctoral thesis). University of Auckland, available only
at https://researchspace.auckland.ac.nz/bitstream/handle/2292/7512/whole.pdf?sequence=2,
retrieved 2015-05-14

Nadoll. P. & Koenig. A. E. (2011). LA-ICP-MS of magnetite: methods and reference


materials. Journal of Analytical Atomic Spectrometry, 26, 1872-1877.

Nesse. W. D. (2009). Introduction to mineralogy (International edition). New York: Oxford


University Press Inc.

Pisiak. L. K., Canil. D., Grondahl. C., Plouffe. A., Ferbey. T. & Anderson. R. G. (2015).
Magnetite as a porphyry copper indicator mineral in till: a test using the Mount Polley
porphyry copper-gold deposit. South-Central British Columbia (NTS 093A). Geoscience BC
Summary of Activites 2014, Geoscience BC, Report 2015-1, 141-150.

Perdhal. J-A. & Frietsch. R. (1993). Petrochemical and petrological characteristics of 1.9 Ga
old volcanics in northern Sweden. Precambrian Research, 64, 239-252.

Reed. S. J. B. (2005). Electron microprobe analysis and scanning electron microscopy in


geology. 2nd edition. United Kingdom: Cambridge University Press.

Plouffe. A. (1996). Detrital transport of metals by glaciers. an example from Pinchi Mine,
central British Columbia. Environmental Geology, 33, 183-196.

45
Rollinson, H. R. (1993). Using geochemical data: evaluation, presentation, interpretation.
University of Michigan: Longman Scientific & Technical.

Rudnick, R. L. & Gao, S. (2004). Composition of the continental crust in Holland, H. D. &
Turekian, K. K. ed., Treatise on Geochemistry, Elviser, Amsterdam. 3, 1-64.

Shilts. W. W. (1982). Glacial dispersal; principle and practical applications. Geoscience


Canada, 9, 42-47.

Sillitoe, R. H., (2010) Porphyry copper systems. Economic Geology, 105(1), 3-41.

Sillitoe. R. H. (1997). Characteristics and controls of the largest porphyry copper-gold and
epithermal gold deposits in the circum-Pacific region. Australian Journal of Earth Science,
44(3), 373-388.

Sinclair. W. D. (2007). Porphyry deposits in Goodfellow. W. D. ed., Mineral Deposits of


Canada: A synthesis of major deposit-types. District Metallogeny, the evolution of geological
provinces and exploration methods. Geological Association of Canada, Mineral Deposits
Divisions, Special Publication, 5, 223-243.

Van Achterbergh. E., Ryan. C. G., Griffin. W. L. (No date). GLITTER! User’s manual –
Online interactive data reduction for the LA-ICPMS microprobe. ARC National Key Centre
for Geochemical Evolution and Metallogeny of Continents (GEMOC), available only at
http://www3.nd.edu/~asimonet/ENGV60500/Glitter_V4.4_manual.pdf

Van Baalen. M. R. (1993). Titanium mobility in metamorphic systems: a review. Chemical


Geology, 110, 233-249.

Wholfarth. B., Björck. S., Funder. S., Houmark-Nilsen. M., Ingólfsson. Ó., Lunkka. J. P.,
Mangerud. J., Saarnisto. M. & Vorren, T. (2008). Quaternary of Norden. Episodes, 31(1).

Öhlander. B., Mellqvist. C. & Skiöld. T. (1999). Sm-Nd isotope evidence of collisional event
in the Precambrian of northern Sweden. Precambrian Research, 93, 105-117.

Ödman. O. H. (1945). A nickel-cobalt-silver mineralization in the Laver copper mine.


Northern Sweden. Sveriges Geologiska Undersökning [SGU], C470, Uppsala, 1-10.

Ödman. O. H. (1943). Geology of the copper deposit at Laver, Northern Sweden. Sveriges
Geologiska Undersökning [SGU], C452, Uppsala, 1-3

46
Appendix
Field notes
Name Location Description *Notes
M1 N7304741 Sampling depth approx. 80 cm, sample size roughly 4 Sample site moved next to
kg. Sandy till with fine material, brown-grey colour. an uprooted tree.
E1702144 Few boulders are noted on the surface.
M2 N7304357 Depth ~90 cm and sample size estimated to 3.6 kg. The
till is sandy and finer material such clay & silt, brow-
E1702217 grey colour and the area are bouldery at the surface.
M3 N7303882 Depth ~80 cm, sample size about 5.3 kg. The till is Collection site moved due
sandy with a light (brown) grey colour. No boulders in to anthropogenic impact
E1703465 the imminent area. (road construction).
M4 N7303957 Depth ~80 cm, sample size estimated to 4.5 kg. Sample Collection site moved
collected close to a road. The sediment is sandy till with again due to anthropogenic
E1703104 small amounts of silt, the colour is light (brown) grey. impact
M5 N7303751 Depth ~56 cm, sample size roughly 5 kg. Sandy till with Site moved due to road
small amounts of silt and clay present. Boulders are construction
E1703006 observed on the surface.
M6 N7302950 Depth ~80 cm, sample size about 5 kg. The till is (fine) Road construction nearby
sand and some silt and the till seems to be sorted. The
E1703723 colour is grey-brown and few boulders are seen.
M7 N7303055 Depth ~80 cm, sample size estimated to 5 kg. The till is Moved due to a bog on the
sandy (fine to medium grain size), light brown-grey original collection site
1703830 colour and boulders are observed at the surface.
M8 N7303732 Depth ~1 m, sample size about 4 kg. Sandy till with silt,
colour is light brown-grey. Big boulders on the surface
E1704073 and the ground is wet.
M9 Not available for sampling due to artificial lake

M10 N7303504 Depth ~60 cm, sample size estimated 4 kg. The till is Collected in old tractor
sandy with some silt and clay, the colour is light brown- tracks.
E1704186 grey. The surface area contain boulders.
M11 N7301485 Depth ~60 cm, sample size roughly 4 kg. The till is Moved up a slope to avoid
sandy and contains some silt, colour is light grey-brown. a bog
E1706129 Boulders are seen at the surface.
M12 N7301934 Depth ~70 cm, sample size approx. 5 kg. The till is
sandy with silt and a light grey-brown colour. The
N1706189 surface area are boulrey (+ one gigantic boulder)
M13 N7301184 Depth ~80 cm, sample size about 4.5 kg. Sandy till with
some silt, colour is light grey-brown. Observations on
1706163 the surface reveals boulders.
M14 N7300806 Depth ~80 cm, sample size roughly 4 kg. Sandy till with
silt, colour is brown. Seems to be sorted based on grain
E1706041 size. Few boulders observed on the surface.
M15 N7300314 Depth ~80 cm, sample size approx. 4.5 kg. The till is Collection site moved due
consisting of fine sand and silt. The colour is light grey- to bog
N1705833 brown. Boulders are observed in the surface area.
M16 N7298348 Depth ~80 cm, sample size estimated at 5 kg. Sandy till
with some silt and clay, light brown-grey colour.
E1711260 Boulders are seen on the surface in the area.
M17 N7296690 Depth ~70 cm, sample size approx. 2.5 kg. The till is
sandy-bouldery, little fine material present, the colour is
E1711742 brown. The surface are full of big boulders.
M18 N7296584 Depth ~80 cm, sample size about 4 kg. Sandy till with Original collection site
some boulders, colour is brown. No boulders are moved due to not finding
E1710147 observed on the surface. any till
M19 N7303240 Depth ~1.2 m, sample size roughly 5 kg. The till is Original collection site
sandy with a grey colour. The areas contain few moved due to not finding
E1699843 boulders on the surface. any till
Old Laver N7304552 Depth ~80 cm, sample estimated to 4 kg. The till consist
of fine sand and silt. The colour is light grey-brown.
mine E739685 Metavolcanic bedrock is present in the area.

47

Das könnte Ihnen auch gefallen