You are on page 1of 69

 

 
The Chemistry of hydrothermal magnetite: A review
Patrick Nadoll, Thomas Angerer, Jeffrey L. Mauk, David French, John
Walshe
PII:
DOI:
Reference:

S0169-1368(13)00302-8
doi: 10.1016/j.oregeorev.2013.12.013
OREGEO 1151

To appear in:

Ore Geology Reviews

Received date:
Revised date:
Accepted date:

1 May 2013
20 December 2013
20 December 2013

Please cite this article as: Nadoll, Patrick, Angerer, Thomas, Mauk, Jerey L., French,
David, Walshe, John, The Chemistry of hydrothermal magnetite: A review, Ore Geology
Reviews (2014), doi: 10.1016/j.oregeorev.2013.12.013

This is a PDF le of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its nal form. Please note that during the production process
errors may be discovered which could aect the content, and all legal disclaimers that
apply to the journal pertain.

ACCEPTED MANUSCRIPT
The Chemistry of hydrothermal magnetite: A review

a,*

, Thomas Angerer , Jeffrey L. Mauk , David French , John Walshe

Patrick Nadoll

SC

RI
P

a
Commonwealth Scientific and Industrial Research Organisation (CSIRO), Australian Resources Research Centre, PO Box 1130, Bentley, WA 6102,
Australia
b
Centre for Exploration Targeting (CET), Department of Earth and Environmental Science, The University of Western Australia, M006, 35 Stirling
Highway, Crawley, WA 6009, Australia
c
US Geological Survey, Central Mineral and Environmental Resources Science Center, MS 973 Denver Federal Center, Denver, CO 80225, USA
d
Commonwealth Scientific and Industrial Research Organisation (CSIRO), Energy Technology, PO Box 52, North Ryde, NSW 1670, Australia

NU

Abstract

Magnetite (Fe3O4) is a well-recognized petrogenetic indicator and is a common accessory mineral in many ore deposits

MA

and their host rocks. Recent years have seen an increased interest in the use of hydrothermal magnetite for
provenance studies and as a pathfinder for mineral exploration. A number of studies have investigated how specific

ED

formation conditions are reflected in the composition of the respective magnetite. Two fundamental questions
underlie these efforts (i) How can the composition of igneous and, more importantly, hydrothermal magnetite be

PT

used to discriminate mineralized areas from barren host rocks, and (ii) how can this assist exploration geologists to
target ore deposits at greater and greater distances from the main mineralization? Similar to igneous magnetite, the

AC
CE

most important factors that govern compositional variations in hydrothermal magnetite are (A) temperature, (B) fluid
composition element availability, (C) oxygen and sulfur fugacity, (D) silicate and sulfide activity, (E) host rock
buffering, (F) re-equilibration processes, and (G) intrinsic crystallographic controls such as ionic radius and charge
balance. We discuss how specific formation conditions are reflected in the composition of magnetite and review
studies that investigate the chemistry of hydrothermal and igneous magnetite from various mineral deposits and their
host rocks. Furthermore, we discuss the redox-related alteration of magnetite (martitization and mushketovitization)
and mineral inclusions in magnetite and their effect on chemical analyses. Our database includes published and
previously unpublished magnetite minor and trace element data for magnetite from (1) banded iron formations (BIF)
and related high-grade iron ore deposits in Western Australia, India, and Brazil (2) Ag-Pb-Zn veins of the Coeur dAlene
district, United States (3) porphyry Cu-(Au)-(Mo) and associated (4) calcic and magnesian skarn deposits in the
southwestern United Sates and Indonesia, and (5) plutonic igneous rocks from the Henderson Climax-type Mo
deposit, United States and the un-mineralized Inner Zone Batholith granodiorite, Japan. These five settings represent
a diverse suite of geological settings and cover a wide range of formation conditions.
1

ACCEPTED MANUSCRIPT
The main discriminator elements for magnetite are Mg, Al, Ti, V, Cr, Mn, Co, Ni, Zn and Ga. These elements are
commonly present at detectable levels (10 to > 1000 ppm) and display systematic variations. We propose a
combination of Ni/(Cr+Mn) vs. Ti+V, Al+Mn vs. Ti+V, Ti/V and Sn/Ga discriminant plots and upper threshold

concentrations to discriminate hydrothermal from igneous magnetite and to fingerprint different hydrothermal ore

RI
P

deposits. The overall trends in upper threshold values for the different settings can be summarized as follows: (I) BIF
(hydrothermal) low Al, Ti, V, Cr, Mn, Co, Ni, Zn, Ga, Sn; (II) Ag-Pb-Zn veins (hydrothermal) high Mn, low Ga, Sn; (III)

SC

Mg-skarn (hydrothermal) high Mg, Mn, low Al, Ti, Cr, Co, Ni, Ga; (IV) skarn (hydrothermal) high Mg, Al, Cr, Mn, Co,
Ni, Zn, low Sn (V) porphyry (hydrothermal) high Ti, V, low Sn; (VI) porphyry (igneous) high Ti, V, Cr, low Mg; (VII)

NU

Climax-Mo (igneous) high Al, Ga, Sn, low Mg, Cr.

MA

Keywords: Magnetite, hydrothermal, mineral deposits, provenance studies, indicator mineral, petrogenetic

1 Introduction

ED

Magnetite (Fe3O4) is an important petrogenetic indicator and pathfinder mineral with a wide array of applications
including geophysical studies, igneous petrology, provenance studies and mineral exploration (e.g., Dupuis and

PT

Beaudoin, 2011, Ghiorso and Sack, 1991b, Grant, 1984/85a, Lindsley, 1976a, McClenaghan, 2005, Razjigaeva and

AC
CE

Naumova, 1992). The compositional variability of magnetite in response to varying formation conditions has been the
focus of many studies over the last few decades. Recent years have seen particular interest in hydrothermal magnetite
as a pathfinder mineral for exploration facilitated by the development and improvement of analytical techniques
such as laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) and trace mode electron
microprobe analysis (EMPA), which allow in-situ measurements with increasingly lower detection limits (e.g. Dupuis
and Beaudoin, 2011, Longerich et al., 1996, Nadoll and Koenig, 2011). We review compositional trends in magnetite
from a variety of hydrothermal ore deposits and their host rocks and discuss how these compositional variations can
be used as a geochemical fingerprint. We present previously published and unpublished minor and trace element data
for hydrothermal and igneous magnetite from (1) Banded iron formations (BIF) and related high-grade iron ore
deposits in Western Australia, India, and Brazil (2) Ag-Pb-Zn veins of the Coeur dAlene district, United States (3)
porphyry Cu-(Au)-(Mo) and associated (4) calcic and magnesian skarn deposits in the southwestern United Sates
(Chino, Cobre, Copper Flat, Safford, Morenci) and Indonesia (Ertsberg district), (5) plutonic igneous rocks from the
Henderson Climax-type Mo deposit, United States and the un-mineralized Inner Zone Batholith granodiorite, Japan.

ACCEPTED MANUSCRIPT
1.1 Mineralogy and Crystallography
Magnetite belongs to the space group Fd3m and has an inverse spinel structure with the general stoichiometry AB2O4
2+

(Bragg, 1915, Fleet, 1981), where A represents a divalent cation such as Mg, Fe , Ni, Mn, Co, or Zn, and B represents a
3+

trivalent cation such as Al, Fe , Cr, V, Mn, or Ga (Lindsley, 1976a, Wechsler et al., 1984). Titanium, with a 4+ charge,

RI
P

can also occupy the B site when substitution is coupled with a divalent cation (Wechsler et al., 1984). Octahedral sites
3+

2+

in the magnetite structure are randomly occupied by subequal numbers of ferric (Fe ) and ferrous (Fe ) iron atoms,
3+

2+

3+

SC

whereas tetrahedral sites are exclusively occupied by the smaller ferric iron atoms Fe [Fe Fe ]O4 (Lindsley, 1976a,

NU

Waychunas, 1991, Wechsler et al., 1984) (Fig. 1A).

Magnetite displays many phase transitions with other spinel group minerals, including but not limited to, spinel

MA

(MgAl2O4), ulvspinel (Fe2TiO4), chromite (FeCr2O4), galaxite (Mn,Mg)(Al,Fe)2O4, gahnite (ZnAl2O4), franklinite ((Zn, Fe,
Mn)(Fe, Mn)2O4), and jacobsite (Mn,Fe,Mg)(Fe,Mn)2O4. Above 600C a continuous solid solution exists between
magnetite and ulvspinel (titanomagnetitess - TixFe3-xO4), and their oxidation products (titanomaghemite), with
4+

3+

2+

3+

ED

coupled substitution of Ti for Fe in the octahedral sites and Fe for Fe in the tetrahedral sites (Buddington and

PT

Lindsley, 1964, Waychunas, 1991, White et al., 1994). Below 600C thermodynamic data are poorly constrained and
extensive miscibility gaps occur (Ghiorso and Sack, 1991b). A log fO2-T diagram with relevant buffers for the Fe-Si-O

AC
CE

system and a schematic phase diagram for the system Fe-O-S in fO2-fS2 space is shown are shown in Figure 2A, B
(redrawn after Frost, 1991a, Frost et al., 1988). The minimum oxygen fugacity for any given temperature at which
magnetite is stable is the iron-magnetite (IM) or magnetite-wstite (MW) buffer (Buddington and Lindsley, 1964,
Frost, 1991a). The fayalite-magnetite-quartz (FMQ) buffer marks the limit above which Fe is mostly incorporated into
magnetite. Below the FMQ buffer iron is predominantly present in silicates. The quartz-iron-fayalite (QIF) buffer marks
the limit below which Fe occurs in its native state. The upper limit for magnetite stability is defined by the hematitemagnetite (HM) buffer, beyond which hematite is the dominant oxide mineral (Buddington and Lindsley, 1964, Frost,
1991a, Grant, 1984/85a). Cation substitution is generally more likely to take place at lower oxygen fugacity
2+

(Buddington and Lindsley, 1964, Frost, 1991a, Lindsley, 1991). Magnesium, Mn, Zn and Ni may substitute Fe ,
whereas Fe

3+

can be replaced by Al, V and Cr (Fig. 1B, C) (e.g., Barnes and Roeder, 2001, Lindsley, 1991, Ramdohr,

1955, Righter et al., 2006a). Vanadium, which has a range of possible oxidation states depending on prevailing oxygen
3+

fugacity conditions, is mainly present as V in titanomagnetites but commonly shows variable minor amounts of V

4+

(Balan et al., 2006, Bordage et al., 2011, Toplis and Carroll, 1995). Vanadium is incompatible at high oxygen fugacity

ACCEPTED MANUSCRIPT
levels due to its 5+ oxidation state. The strong oxygen fugacity-dependence of V can be used to help model the
differentiation of magmatic intrusions (Toplis and Corgne, 2002). Similar applications are relevant for hydrothermal
conditions. For example, magnetite and coexisting hematite have been used to define redox-potentials in

hydrothermal fluids associated with ore deposits (Barnes, 1997, Otake et al., 2010).

RI
P

1.2 Magnetite as a petrogenetic indicator

SC

Magnetite is one of the most abundant oxide minerals in the continental crust and has been recognized as an
th

important indicator mineral for petrogenetic and geochemical studies since the early 20 century (Ramdohr, 1926).

NU

Indicator minerals such as magnetite are generally more resistant to weathering and transport than other coexisting
mineral phases (McClenaghan, 2005). Magnetite is a widespread and easily identifiable accessory mineral in igneous,

MA

sedimentary, and metamorphic host rocks of a wide range of compositions (e.g., Annersten, 1968, Buddington and
Lindsley, 1964, Prins, 1972, Ramdohr, 1955, Rumble, 1976b, Shcheka et al., 1980, Vincent and Phillips, 1954) and can
incorporate a large number of foreign cations (Bowles et al., 2011, Lindsley, 1976a, Wechsler et al., 1984) (Fig. 1). In

ED

combination with isotopic studies, compositional patterns in magnetite can provide important constraints on

PT

petrogenetic factors such as temperature, pressure, and oxygen/sulfur fugacity (Anderson et al., 2008, Chiba et al.,
1989, Frost, 1991b, Ghiorso and Sack, 1991b). Magnetite forms ideal pairs with silicates, carbonates, and other oxides

AC
CE

such as hematite or ilmenite for geothermometry and geobarometry due to its distinct oxygen isotope fractionation
factor (e.g., Buddington et al., 1955, Buddington and Lindsley, 1964, Chacko et al., 2001, Friedman and O'Neil, 1977,
Ghiorso and Sack, 1991a, Powell and Powell, 1977, Sauerzapf et al., 2008, Zheng and Simon, 1991). By virtue of its
magnetic properties, magnetite has also been the focus of many geophysical studies that investigated its use for
geophysical mapping and exploration (e.g., Akimoto, 1955, Behn et al., 2001, Clark, 2001, Grant, 1984/85b, Malmqvist
and Parasnis, 1972, McEnroe et al., 2001).

Minor and trace elements that have been reported in magnetite ranges from alkali and alkaline earth metals to
transition metals, including REEs, and metalloid and non-metals (e.g., Borisenko and Lapin, 1972, Dare et al., 2012,
Dupuis and Beaudoin, 2011, Klemm et al., 1985, Lindsley, 1991, McQueen and Cross, 1998, Nadoll et al., 2012, Nielsen
and Beard, 2000, Righter et al., 2006a, Singoyi et al., 2006). However, a suite of elements, namely Mg, Al, Ti, V, Co, Ni,
Zn, Cr, Mn, Ga and Sn are commonly present in magnetite of all origins at concentrations that can be detected by
electron microprobe analysis (10 to > 1000 ppm). We refer to these cations as magnetite elements.

ACCEPTED MANUSCRIPT
Concentrations of foreign cations in magnetite are responsive to a number of external factors that reflect formation
conditions of the respective host rock. Comprehensive overviews of physicochemical properties of magnetite can be
found in Bowles et al. (2011) and the Reviews in Mineralogy volumes on oxide minerals (Lindsley, 1991, Rumble,

1976a). Many classic studies that focused on igneous magnetite show extensive compositional variability in response

RI
P

to seven main controlling factors: (1) source rock or magma composition, (2) temperature, (3) pressure, (4) cooling
rate, (5) oxygen fugacity, fO2, (6) sulfur fugacity, fS2, and (7) silica and sulfide activity (Buddington and Lindsley, 1964,

SC

Frost and Lindsley, 1991, Ghiorso and Sack, 1991a, Haggerty, 1991a, Mollo et al., 2013, Whalen and Chappel, 1988).
Additionally, crystallographic factors such as ionic radius and overall charge balance are important aspects that put

NU

decisive constraints on possible substituting cations (Cornell and Schwertmann, 2003, Fleet, 1981, Goldschmidt, 1954,
Wechsler et al., 1984) (Fig. 1). Magnetite is also an important accessory mineral in metamorphic rocks. The

MA

composition of metamorphic magnetite changes in response to mainly two factors: temperature and oxygen fugacity
(Frost, 1991c). The partitioning behavior and distribution of trace elements in magnetite depends on the metamorphic
grade (Evans and Frost, 1975, Skublov and Drugova, 2003, Van Baalen, 1993). Low grade metamorphic magnetite is

ED

compositionally homogeneous and has very low trace element concentrations compared to most other types of

PT

hydrothermal magnetite (Frost, 1991c, Nadoll et al., 2012). This has been interpreted as a reflection of the low
formation temperatures (Nadoll et al., 2012). Although elements such as Ti are essentially immobile under low

AC
CE

temperatures, Van Baalen (1993) showed that Ti can be mobilized on a length scale of meters even in low-grade
metamorphic rocks. Verlaguet et al. (2006) came to similar conclusions for the comparably immobile element Al.
Furthermore, Al contents in spinel minerals from metamorphic rocks containing silicates are governed by pressure-,
temperature-, and water fugacity sensitive equilibria involving chlorite (Evans and Frost, 1975). Oxide-silicate, oxideoxide, and intra-oxide re-equilibration reactions are further important controls of the magnetite stability during
cooling of plutonic and metamorphic rocks (e.g., Frost, 1991a, b, c, Frost and Lindsley, 1991, Nadoll et al., 2012,
Wones, 1989).

Recent studies put further constraints on compositional trends of igneous magnetite and reinforced its petrogenetic
significance. For example, the significance of magnetite from massive sulfide NiCuPGE deposits as a sensitive gauge
for fractionation processes in sulfide melts has been investigated by Dare et al. (2012). They found that minor and
trace element concentrations in magnetite are governed by the successive depletion of lithophile elements in the
sulfide melt and the uptake of siderophile elements such as Co, Mo, Ni, Pb, Sn and Zn in co-crystallizing sulfides.
Jenner et al. (2010) recently emphasized the significant role that the onset of magnetite crystallization plays for the
5

ACCEPTED MANUSCRIPT
saturation of sulfides in a melt. Righter et al. (2006a) presented new values for Ni, Co, and V partition coefficients
between Cr-rich spinels and silicate melt. Igneous magnetite from carbonatites has been investigated by a number of
authors who found indicative minor and trace element signatures that can be attributed to controlling factors such as

oxygen fugacity and interdependencies among substituting cations (Bagdasarov, 1989, Bailey and Kearns, 2002, Reguir

RI
P

et al., 2008). Reguir et al. (2008) observed V and Mn concentrations to be directly controlled by oxygen fugacity.
Furthermore, they present data that Zn in igneous magnetite from carbonatites shows a covariation with Mn,

SC

suggesting a link between the two elements controlled by coupled substitution.

Ryabchikov and Kogarko (2006) constructed fO2-T diagrams for the Russian Khibina magmatic system using Fe-Ti-oxide

NU

equilibria. Varying Mg, Al, Mn, Ti, and Cr concentrations in magnetite have been linked with the different intrusive
stages at the Sokli complex in northeastern Finland by Lee et al. (2005). Laser ablation inductively coupled plasma

MA

mass spectrometry (LA-ICP-MS) trace element analyses of magnetite from Kiruna, Sweden showed that V and Mn are
the most abundant trace elements in igneous magnetite, whereas Cu, Zn, and Pb are not commonly incorporated

ED

(Mller et al., 2003).

The identification of characteristic minor and trace element signatures in magnetite of predominantly hydrothermal

PT

origin commenced with the pioneering work of Borisenko and Lapin (1971, 1972) and Shcheka et al. (1988), who were

AC
CE

also among the first to apply multivariate statistics to magnetite minor and trace element data in order to reveal
underlying compositional trends that could be used to discriminate sample populations. Recent studies have
presented data that, suggest that magnetite from magmatic-hydrothermal deposits displays systematic variations in
minor and trace element concentrations that can be used to fingerprint ore deposits (Beaudoin and Dupuis, 2009,
Dupuis and Beaudoin, 2011, Kamvong et al., 2007, Nadoll et al., 2012, Rusk et al., 2009, Singoyi et al., 2006). Carew et
al. (2006) and Rusk et al. (2010) presented evidence that Ti, V, and Mn concentrations in hydrothermal magnetite
from the Ernest Henry iron oxide copper gold (IOCG) deposit in Australias Cloncurry region can be used to
discriminate barren from mineralized host rocks. Singoyi et al. (2006) observed three distinct types of magnetite
(referred to as Type A, B, and C) from selected volcanic-hosted massive sulfide (VMS), skarn, IOCG, and Broken Hilltype clastic-dominated Pb-Zn deposits in Australia. Type A commonly and consistently incorporates Mg, Al, Ti, V, Mn,
Co, Ni, Zn, Ga and Sn at levels above LA-ICP-MS detection limits, whereas type B displays lesser concentrations and
more hetreogenous distribution across different samples for Cr, As, Zr, Nb, Mo, REE, Ta, W, Pb. Their type C magnetite
hosts no detectable concentrations of Cu, Ag, Se, Tl, Te, Bi, Au. Singoyi et al. (2006) further suggest that Sn/Ga and
Al/Co ratios can help to distinguish magnetite from VMS, skarn, IOCG, and Broken Hill-type clastic-dominated Pb-Zn
6

ACCEPTED MANUSCRIPT
deposits; an approach that was later adopted by Kamvong et al. (2007). McQueen and Cross (1998) demonstrated
how trace element variations in hydrothermal magnetite hosted in contact metasomatic skarn deposits can provide
characteristic signatures and consequently, help to target magnetite bearing ore deposits. Most recently, Duan et al.

(2012) documented systematic compositional variations among different types of hydrothermal magnetite, tracing the

RI
P

evolution of the magmatichydrothermal mineralization in the Washan porphyry-style Fe deposit, East China. Huang
et al. (2013) described the compositions of hydrothermal magnetite from the magmatic-hydrothermal Cihai Fe deposit

SC

in China. They compared their results with the magnetite classification scheme suggested by Dupuis and Beaudoin
(2011) and showed that magnetite from the Cihai deposit is depleted in V and Ti compared to magnetite from Fe-Ti-V

NU

deposits and has Mg, Al, Ti, V, Cr, Co, Ni, Mn, Zn, Ga, and Sn concentrations that exceed those commonly found in
magnetite from skarn deposits.

MA

1.3 Magnetite in exploration and provenance studies


Mining and exploration companies have recognized the great potential of employing the geochemistry of magnetite

ED

as an exploration tool for many years. There is a large body of unpublished work that investigated the use of

PT

magnetite as a pathfinder or indicator mineral. One of the main challenges in mineral exploration is to detect
geochemical signatures of ore deposits at greater and greater distances from the main mineralization. Magnetite,

AC
CE

among other detrital heavy minerals such as ilmenite, rutile, garnet, and zircon has been a valuable asset to mineral
exploration and provenance studies (e.g., Dupuis and Beaudoin, 2011, Nadoll et al., 2012, Nadoll et al., submitted,
Razjigaeva and Naumova, 1992, Shcheka et al., 1982). Hydrothermal and igneous magnetite can provide vectors to
mineralized areas where the primary mineralogical context is lost, deeply covered, or highly altered, similar to the use
of rutile as a resistate indicator mineral (Anand and Butt, 2010, Borisenko and Lapin, 1972, Buddington and Lindsley,
1964, Hutton, 1950, McQueen and Cross, 1998, McQueen and Whitbread, 2002, Rumble, 1976a and references
therein, Shcheka et al., 1982, Triebold et al., 2007, Zack et al., 2004). Razjigaeva and Naumova (1992) used variations
in Ti, Mn, Cr, V, Ni, Co, Zr, Sn, Zn, Pb, and Cu concentrations in detrital magnetite to trace the source rocks of
sediments, comparing them to a set of standard igneous magnetites first described by Shchecka et al. (1980).
Similarly, Grigsby (1990) used petrographic and chemical variations in detrital magnetite to discriminate among felsic,
intermediate, and mafic volcanic and plutonic source rocks.

1.4 Defining hydrothermal magnetite


7

ACCEPTED MANUSCRIPT
The term hydrothermal magnetite can be rather ambiguous especially when considering the complex geological and
mineralogical relationships found in many hydrothermal ore deposits (Fig. 3). Multiple vein generations, large to
small-scale overprints of multiple alteration stages, fluid/rock interaction, and secondary weathering processes

translate into a variety of different types of hydrothermal magnetite. These are commonly synonymously addressed in

RI
P

publications. However, it is important to consider the different types of hydrothermal magnetite when attributing
specific minor and trace element concentrations to a specific deposit or type of mineralization. One of the main

SC

challenges for researchers that investigate hydrothermal magnetite is to diligently record the large variety of minor
and trace element compositions in magnetite that reflect a diverse formation and alteration history, and at the same

NU

time invoke diagnostic trends that can be used to reliably discriminate different ore deposits, as well as barren from

MA

mineralized zones.

2.1 Electron microprobe

ED

2 Methods

PT

Electron microprobe analysis (EMPA) were carried out at North Ryde (Sydney) on a Cameca Camebax automated
wave length dispersive spectroscopy scanning electron microprobe with an acceleration voltage of 15 kV and a beam

AC
CE

current of 20 nA. The analytical conditions for each element are given in Table 1. Line overlap corrections were
performed for the overlap of Ti K on VK, V K on CrK, Cr K on MnK and Mn K on FeK. ZAF corrections were
applied for all elements. Microprobe results were accepted only if the compound weight percent total was between
98.5 and 101.5%. The ferric iron content of each analysis is calculated based on the assumption of stoichiometry and
an ideal AB2O4 formula. Tin and Ga were not measured.

Back-scattered electron (BSE) and multi-element imaging with a field emission gun scanning electron microscope
(FEG-SEM) was undertaken at CSIRO Earth Sciences and Resource Engineering, Perth, with an acceleration voltage of
20 keV and a beam current of 2.9 nA.

2.2 Laser ablation inductively coupled plasma mass spectrometry


The growing number of studies that employ LA-ICP-MS as the preferred analytical technique to collect ppm to subppm trace element concentrations in magnetite have produced a body of evidence that a matrix matched external
standard is not required and that for most conditions and spot sizes reference materials such as GSE-1G, GRM278,
8

ACCEPTED MANUSCRIPT
BCR2-G, NIST610, and NIST612 are well suited and produce reliable results (Dare et al., 2012, Nadoll and Koenig, 2011,
Savard et al., 2010). Detection limits at ppm to ppb levels shed light on subtle but characteristic compositional
variations for elements of interest that have formerly been hampered by greater detection limits or limitations to the

number of elements measured.

RI
P

The LA-ICP-MS operation conditions for the different datasets are listed in Table 1. Iron concentrations determined by
electron microprobe analyses were used as an internal standard for calibration of LA-ICP-MS data. The ferric iron

SC

content of each analysis was calculated based on the assumption of stoichiometry and an ideal AB2O4 formula.

NU

Table 2 shows the average laser ablation ICP-MS reporting limits (USGS) and detection limits (UQAC, CODES - (Angerer
et al., 2013, Angerer et al., 2012)). Thirty-three elements were included in the analyses: Na23, Mg24, Al27, Si28, S34,

MA

Ca44, Ti47, V51, Cr52, Mn55, Fe57, Co59, Ni60, Cu63, Zn66, Ga71, Ge72, As75, Se82, Sr88, Y89, Nb93, Mo98, Ag107,
In115, Sn118, Sb121, Ba138, Ce140, W184, Au197, Tl205, Pb208, Th232. This list includes a number of non-spinel
elements, such as Si, Ca, and the REEs, which can be helpful for data screening and the identification of analyses that

ED

are affected by inclusions. Laser ablation data were processed by non-automatic spectra reduction to exclude spectra

AC
CE

2.3 Data treatment

PT

or parts of spectra that are affected by mineral inclusions.

Compiling data from different sources poses many challenges in regard to data quality and comparability. Different
analytical techniques such as EMPA and LA-ICP-MS, with a range of varying operation conditions, translate into a wide
range of detection limits. The lowest comparable element concentrations are unavoidably defined by the greatest
detection limit for the respective element when comparing minor and trace element concentrations across different
datasets. Several statistical methods can be used for data analysis; we show box and whisker plots that display lower
th

th

quartile, median, mean, upper quartile, and the 5 and 95 percentile values, as well as discriminate outliers. The
th

upper threshold values, representing the upper 95 percentile of the data, are usually not affected by the varying
detection limits and therefore represent a robust approach to compare data from different sources. To eliminate
detection limit related minor and trace element variations, element ratios have also been employed for data
comparison.

3 Geologic background
9

ACCEPTED MANUSCRIPT
This section provides a geologic overview of the investigated areas and deposits. Locality maps are shown in Fig. 4 and
Fig. 5. A complete list of samples including host rock and alteration types are shown in Table 3.

3.1 Banded iron formations and related iron ore deposits

RI
P

Banded iron formations (BIFs) are the most important sedimentary host rock for iron ore. BIF have been deposited
predominantly during Eoarchean to Paleoproterozoic times, and are preserved in two main tectonostratigraphic

SC

settings - volcanic rock dominated granite-greenstone terranes (Algoma-type BIF) or continental shelf environments
(Superior-type BIF). Despite their difference in stratigraphic successions, both Algoma- and Superior-type BIF are

NU

mineralogically similar, consisting of iron oxide, siderite to ankerite-dolomite, or iron silicate laminae interlayered with
chert/quartz laminae (Fig. 6A) (Klein, 2005, Trendall and Blockley, 1970). Both Algoma- and Superior-type BIF locally

MA

host high-grade iron ore bodies of significant economic size, in which iron underwent an enrichment process by postdepositional alteration processes. The processes and geological conditions for enrichment of iron in BIF can vary from
district to district, and several genetic models have been proposed (Angerer and Hagemann, 2010, Clout, 2003, Evans

ED

et al., 2013, Gruner, 1937, Lobato et al., 2008, Morris, 1985, Ramanaidou, 2009, Rosire et al., 2008, Taylor et al.,

PT

2001). Unaltered and unweathered BIF typically contains diagenetic or metamorphic magnetite and hematite, and a
dominance of one over the other reflects the metamorphic grade and oxygen fugacity in the rock (Klein, 2005).

AC
CE

Magnetite is only a minor constituent in the most economic zones of many BIF-hosted iron ore deposits, which is due
to pervasive oxidation and replacement by other minerals. However, magnetite occurs in deeper sections of ore
deposits, below the weathering front, where it commonly forms medium- to high grade proto-ore, and in these
zones complex mineral associations, including iron-rich carbonates and/or silicates are commonly observed (Angerer
et al., 2012, Duuring and Hagemann, 2012, Figueiredo e Silva et al., 2008, Gutzmer et al., 2005, Mukhopadhyay et al.,
2008, Taylor et al., 2001). Magnetite in BIF and BIF-hosted iron ore (proto-ore) has several forms: (1) laminated bands
made of amalgamated, anhedral to subhedral grains (Fig. 6A), (2) disseminated grains within, but obscuring BIF
laminae (Fig. 6B), and (3) layer-discordant veins, shear zones and cleavage filling. Magnetite grain sizes commonly
increase with metamorphic grade without modifying textural equilibrium relationships with neighboring quartz (Klein,
2005). The presented data set includes 158 LA-ICP-MS analyses of metamorphic and hydrothermally altered
magnetite from various major iron ore districts.

3.1.1

Meso-, and Neoarchean Algoma-type BIF

10

ACCEPTED MANUSCRIPT
The data set includes 57 LA-ICP-MS analyses representing magnetite from Mesoarchean Algoma-type BIF from the
Yilgarn craton, Western Australia: Koolyanobbing and Marda greenstone belts

(Angerer et al., 2013, Angerer et al.,

2012), and Weld Range (previously unpublished data, kindly provided by Paul Duuring, The University of Western

Australia) (Fig. 4A). The sample set consists of metamorphic and hydrothermally-altered magnetite. Peak burial

RI
P

metamorphic temperatures are recorded to range between 200 to 350C (Angerer et al., 2013, Angerer et al., 2012,
Duuring and Hagemann, 2013). , Temperatures of the various hydrothermal overprints are not well constraint, but

SC

isotopic equilibriums calculations (Angerer, unpublished data) suggest generally lower temperatures than peak
temperature (200 to 300C). Twenty-one analyses are from the Mesoarchean Iron Ore Group of JharkhandOrissa

NU

region, India (Bhattacharya et al., 2007) (Fig. 4C). Peak metamorphic overprint is reported as lower greenschist facies,
i.e., ~270-350C (Bhattacharya et al., 2007).

Neoarchean-Paleoproterozoic Superior-type BIF

MA

3.1.2

Neoarchean to Paleoproterozoic Superior-type BIF is represented by 39 magnetite analyses from the Marra Mamba

ED

and Mt Sylvia Formations in the Hamersley Province, Western Australia (Angerer et al, unpublished data), and three
unpublished analyses from the Cau Formation, Quadriltero Ferrfero (QF), Brazil (kindly provided by Ana-Sophie

PT

Hensler, The University of Western Australia). Peak burial metamorphic temperatures range between 200 and 300 C
in the Hamersley Province (Klein, 2005), and 250 to 350C in the QF (Rosire and Rios, 2004). Hydrothermally altered

AC
CE

magnetite from various alteration zones in the Tom Price high-grade iron ore deposit is represented by 38 previously
unpublished magnetite analyses (kindly provided by Warren Thorne, The University of Western Australia) (Fig. 4B).
The hydrothermal fluids producing magnetite in the Tom Price deposit were basinal brines (107142C, 25.5 wt %
equiv; NaCl-CaCl2rich) (Thorne et al., 2004).

3.2 Ag-Pb-Zn veins (Lucky Friday, Coeur dAlene district, Idaho)


The Lucky Friday mine is located in the Coeur dAlene mining district of northern Idaho, which is one of the major AgPb-Zn producers in the world. The Ag-Pb-Zn replacement veins formed at 250-350C and 1-3 kilobars in clastic burial
metamorphic rocks of the Middle Proterozoic Belt Supergroup (Beaudoin and Sangster, 1992, Landis and Hofstra,
1991, Leach et al., 1988). Vein formation temperatures correspond with peak metamorphic conditions in the area
(Doughty and Chamberlain, 1996, Nadoll et al., 2012, Sack and Lichtner, 2009). Siderite and calcite are the
predominant gangue minerals in mineralized veins. Hematite and magnetite are common accessory minerals in the
host rocks, but they can also occur as rare accessory minerals in hydrothermal veins (Fryklund, 1964, Nadoll et al.,
11

ACCEPTED MANUSCRIPT
2012). Our database includes thirty previously published LA-ICP-MS analyses of magnetite (Nadoll et al., 2012), from
the Gold Hunter calcite and siderite veins (Fig. 6C)of the Lucky Friday mine.

3.3 Porphyry Cu-(Au)-(Mo) and related skarn deposits

RI
P

The geology and magmatic-hydrothermal processes involved in the formation of porphyry and associated skarn
deposits have been described by many authors (e.g., Audetat et al., 2008, Einaudi, 1981b, Seedorff et al., 2005,

SC

Sillitoe, 2010, Titley, 1981, Titley and Beane, 1981). Multiple stages of mineralization, alteration and in most cases a
diverse suite of plutonic, volcanic and (meta)-sedimentary host rocks are common characteristics of porphyry and

NU

related skarn deposits (e.g., Beane and Titley, 1981, Einaudi, 1981b, Halsall, 1983, Lowell and Guilbert, 1970, Seedorff
et al., 2005, Sillitoe, 2010, Titley, 1981). Hydrothermal veins form at temperatures between 300 and 800C (e.g.,

MA

Beane and Titley, 1981, Einaudi, 1981a, Muntean and Einaudi, 2001, Rusk et al., 2008, Seedorff et al., 2005, Sillitoe,
2000, Sillitoe, 2010, Ulrich et al., 2001). Skarn deposits can form where porphyry systems intrude carbonate rocks,
(e.g., Beane and Titley, 1981, Einaudi, 1981b, Einaudi et al., 1981, Meinert et al., 2005). Magnetite of igneous and

ED

hydrothermal origin is a ubiquitous accessory mineral in porphyry Cu-(Au)-(Mo) and associated skarn deposits.

PT

Magnetite commonly forms prior to and during the early potassic alteration (Arancibia and Clark, 1996), but can also
be found in later stage magnetite-rich alteration as disseminated accessory minerals or in veins of varying

AC
CE

assemblages (e.g., Beane and Titley, 1981, Gustafson and Hunt, 1975, Sillitoe, 2000, Sillitoe, 2010, Titley and Beane,
1981). In skarn deposits, hydrothermal magnetite can be an abundant mineral phase, occurring as massive
replacement mineralization as a result of Fe-rich fluids (e.g., Einaudi, 1981b, McQueen and Cross, 1998, Meinert et al.,
2005). To assure that hydrothermal and igneous magnetite are properly identified in the complex porphyry and skarn
environment, a combination of field observations and detailed microscopy as well as careful mineral separation has
been undertaken.

3.3.1

American southwest (Arizona, New Mexico)

The Laramide Belt in the southwestern United States hosts porphyry and skarn deposits of different ages (Jurassic,
Cretaceous, and Tertiary), dimensions, and metal content (Cu, Mo, Ag, Au) (e.g., Ahmad and Rose, 1980, Enders, 2000,
Halsall, 1983, McLemore et al., 1999, Robinson and Cook, 1966, Rose and Baltosser, 1966, Seedorff et al., 2005,
Sillitoe, 2010, Titley, 1981, Titley and Beane, 1981). Formation temperatures range from 300 to 800C in porphyry
deposits and approximately from 300 to 500C in the associated skarn deposits (e.g., Ahmad and Rose, 1980, Audetat
and Pettke, 2006, Beane and Titley, 1981, Manske and Paul, 2002, Sillitoe, 2010). Previously published data (Nadoll et
12

ACCEPTED MANUSCRIPT
al., submitted) presented here are from five classic porphyry and related skarn deposits (Fig. 5A): (1) Morenci, which is
North Americas largest producer of copper and one of the largest open pit mines in the world; (2) the Safford district,
with its four porphyry Cu deposits; (3) Chino (Santa Rita), which was one of the worlds first low-grade open pit copper

mines, (4) the Cobre porphyry Cu and skarn deposit, and (5) the Copper Flat porphyry and Zn-skarn deposit. Our

RI
P

database includes previously published data for 100 analyses of igneous magnetite and 118 analyses of hydrothermal
magnetite from these porphyry deposits. Furthermore, 59 analyses of hydrothermal magnetite from calcic skarn, and

SC

17 analyses of hydrothermal magnetite from magnesian skarn are presented here. All of these data have been

3.3.2

Ertsberg district (Papua, Indonesia)

NU

previously published (Nadoll et al., submitted).

The Ertsberg district is among the worlds largest Cu-Au resources districts and contains a number of porphyry and

MA

skarn Cu-Au-(Mo) deposits. The Grasberg Igneous Complex is located within the Ertsberg district, in the central
highlands of New Guinea in the province of Papua (formerly Irian Jaya), Indonesia (Fig. 5D). Pliocene intrusions of

ED

dioritic to monzonitic compositions were emplaced in deeply folded siliciclastic and carbonate host rocks of Tertiary
age to produce two porphyry orebodies (ca. 3 Ma, (Mathur et al., 2000)) with disseminated and stockwork-type

PT

mineralization and sulfide-rich skarn at the margins of the igneous complex (Fig. 5E) (MacDonald and Arnold, 1994,
Pollard and Taylor, 2002, Pollard et al., 2005). Grasberg is one of the biggest Cu and Au mines in the world

AC
CE

(MacDonald and Arnold, 1994, Meinert et al., 1997, Pollard et al., 2005). The main Grasberg intrusion (ca. 3 Ma) is
characterized by successive stages of intense K-feldspar, biotite, and magnetite alteration (Pollard and Taylor, 2002).
The nearby Ertsberg East Skarn (ca. 2.6 Ma, (Pollard et al., 2005)) is a calcium-magnesium silicate skarn hosted by
carbonates of the Tertiary Waripi and Faumai Formations (Mertig et al., 1994). Magnetite of igneous and
hydrothermal origin is a widespread accessory to major mineral phase in the intrusive rocks and hydrothermal veins,
as well as in the associated skarn. A total of 892 previously unpublished EMPA analyses of hydrothermal magnetite
from the Ertsberg district are included in our database. Of these, 722 represent hydrothermal magnetite from the
Grasberg porphyry and associated intrusive rocks and 59 are from hydrothermal skarn magnetite. An overview of the
geology and sample locations is shown in Fig. 5 and Table 3.

3.4 Plutonic igneous rocks


3.4.1

Climax-type Mo deposit (Henderson, Red Mountain), Colorado

13

ACCEPTED MANUSCRIPT
A set of igneous magnetites (41 analyses) is from the Henderson Climax-type Mo deposit (Fig. 5B), which is the worlds
largest operating molybdenum mine (Carten et al., 1993), located on the east side of the Continental Divide fifty miles
west of Denver, Colorado. The stock, represented by a leucorhyolite/leucogranite intrusive complex with high-

temperature quartz-molybdenum veins and fractures, is potassically altered and enveloped by distal quartz-sericite-

RI
P

pyrite (QSP) alteration (Seedorff and Einaudi, 2004, Shannon et al., 2004). Magnetite is a common disseminated
accessory mineral in the igneous host rocks, and, although no data for hydrothermal magnetite from this deposit are

3.4.2

SC

presented here, it also occurs in hydrothermal veins.

Inner Zone Batholith (Japan)

NU

The unmineralized granodiorite of the Inner Zone Batholith in southwest Japan (Fig. 5C) is one of the classic
magnetite-series granitic rocks first described by Ishihara (1977). Median minor and trace element concentrations for

MA

igneous magnetite from the Inner Zone Batholith (Spong, 1998) have previously been used for comparison of
magnetite from porphyry and associated skarn deposits, and can be employed as a baseline that highlights relative

ED

enrichment or depletion of foreign cations in magnetite (Nadoll et al., 2012, Nadoll et al., submitted).

PT

4 Magnetite chemistry

AC
CE

A total of 1415 EMPA and LA-ICP-MS analyses, of which 1103 have not been previously published, embody the
database for our geochemical evaluation. Summary statistics are shown in Table 4 and the corresponding box and
whisker plots are shown in Fig. 7. Probability plots (Fig. 8) and an upper threshold value radar plots (Fig. 9) are
essentially derivates of the box and whisker plots, but help to identify compositional trends by displaying different
aspects of the data. Probability plots help to distinguish sub-populations within groups and radar plots concisely
illustrate relative variations for the suite of magnetite elements, namely Mg, Al, Ti, V, Cr, Mn, Co, Ni, Zn, Ga and Sn.
These elements have been selected as main discriminators for magnetite based on their common incorporation into
the magnetite structure at detectable limits and their data availability across EMPA and LA-ICP-MS datasets. Further
important elements, such as Si, Pb, W, and Nb, that can, in some cases, also be useful discriminators, are discussed
where applicable.

4.1 Deposit type comparison

14

ACCEPTED MANUSCRIPT
The descriptive illustration of compositional variations in magnetite from various geological settings, including
hydrothermal ore deposits, lays the foundation for a more abstract classification scheme that is used to discriminate
different types of mineral deposits and, hopefully will help exploration geologists to target mineralized areas.

Banded iron formation and related iron ore deposits (158 analyses)

4.1.1

RI
P

Magnetite in BIF and iron ore deposits commonly have only trace amounts of Fe-substituting cations. Seventy to 90
percent of the analyses show concentrations of less than 10 ppm for Ti, V, Cr, Co, Ni, and Zn and merely 40 percent of

SC

the data exceed 1 ppm for Ga. Only Mg, Al, Si (not shown), and Mn consistently have concentrations above 10 ppm in
magnetite from BIF and iron-ore deposits. In particular, Mg and Si show trends towards greater concentrations, with

NU

median values of 556 ppm and 8000 ppm respectively (Fig. 7, Fig. 8, Table 4). Higher values with positive correlation
among Si, Mg, Al, Ca, and/or Mn are common but may reflect contamination from ubiquitous micro to nanometer

MA

scale inclusions. Cations such as Cu, As, Pb, Mo and W occur in trace amounts and do not correlate with inclusion
elements such as Si, P, or S. The color-coded field emission gun SEM image in Fig. 10 illustrates that mineral inclusions

ED

can be an abundant feature in magnetite grains and have the potential to contribute and skew chemical analyses.

Microprobe data from several BIFs from North America (Dupuis and Beaudoin, 2011) show median concentrations of

PT

Si 1400, Zn 600, Al 350, Cu 340, Ni 310, Cr 250, Mn 180, V 160, Ti 150, Mg 130, and Ca 90 ppm, and thus yield enriched

AC
CE

concentrations compared to our LA-ICP-MS data (Fig. 12). Control analyses to compare EPMA (Dupuis and Beaudoin,
2011 set up) and ICPMS (Chicoutimi) data revealed that, at ICPMS values <50 ppm, EPMA data of Zn, Cu, Ni, V, and
partly Cr are at least one order of magnitude higher than ICPMS data. Thus, data for Zn, Cu, Ni, Cr, and V in BIF (Dupuis
and Beaudoin, 2011) may be overestimated as a result of insufficiently low microprobe detection limits.

Metamorphic magnetite in unaltered BIF rarely contains more than 10 ppm Co and 50ppm Ni and has Co/Ni ratios
ranging from 0.1 to 1 (this study, Bajwah et al., 1987). No significant differences between Algoma type and Superior
type BIF have been observed. In comparison to metamorphic magnetite, hydrothermally-altered magnetite in low
grade metamorphic hematite-chert BIF from the Carajas Serra Norte deposits, Brazil, is characterized by average Co
and Ni concentrations that exceed 500 and 350 ppm, respectively, and Co/Ni ratios greater than 1 (Figueiredo e Silva
et al., 2009).

4.1.2

Ag-Pb-Zn veins (30 analyses)

Median minor and trace element concentrations in hydrothermal magnetite from the Lucky Friday Ag-Pb-Zn veins are
comparatively low, with the lowest median concentration for Mg observed in the data (Fig. 7, Table 4). Median
15

ACCEPTED MANUSCRIPT
concentrations for Al, Cr, Mn, Co, Zn, and Ga are also characteristically low compared to most other magnetite
occurrences; only magnetite from BIF and iron ore deposits has more depleted Al, Zn, and Ga concentrations.

The minor and trace element concentrations in hydrothermal magnetite from the Lucky Friday Ag-Pb-Zn veins range

widely. In particular, Mg, Mn, and Ni have minimum and maximum concentrations that cover four orders of

RI
P

magnitude. This compositional variation can partly be attributed to the different vein types that occur in the Coeur
dAlene mining district. Hydrothermal magnetite from siderite veins displays higher Mg and Mn concentrations

SC

compared to magnetite in calcite veins as described in detail by Nadoll et al. (2012). This distinguishing feature is

0.38), shows two distinct sub-groups for V (split at 0.5).

Calcic and magnesian skarn (229 and 17 analyses)

MA

4.1.3

NU

illustrated in the probability plots in Figure 8, which, in addition to Mn (split at 0.5), Mg (split at 0.4), and Ni (split at

Hydrothermal magnetite from calcic and magnesian skarn has the highest recorded median Mg concentrations.
Gallium concentrations are comparable in both types. Calcic and magnesian skarn magnetite display distinct variations

ED

in their Mn, Ti, V, Co, and Zn concentrations. These elements have concentrations that are one to two orders of
magnitude greater in calcic skarn magnetite than in magnesian skarn magnetite. Furthermore, median Mn and Zn

PT

concentration in magnetite from calcic skarn are an order of magnitude greater than in any other type of magnetite,

AC
CE

with 60% exceeding 10000 ppm for Mn and 80% of the measured Zn concentrations exceeding 1000 ppm. This trend is
also reflected in the upper threshold values for calcic skarn magnetite, which significantly exceed the values for Mg,
Mn, and Zn measured in any other type of magnetite. Magnesian skarn magnetite has the lowest overall upper
threshold values for Ti and Co in the dataset (Table 4, Fig. 7) and Ni and Cr concentrations lie below the detection
limit. Upper threshold values for Mn are lower in magnesian skarn magnetite than in most other magnetite types.
Magnesian skarn magnetite incorporates characteristically high Sn concentrations with a median of 82.5 ppm and an
upper threshold value of 99.1 ppm (Table 4, Fig. 7). Igneous magnetite from the Henderson Climax-type Mo deposit is
the only other magnetite occurrence with comparable Sn concentrations. The probability plot in Figure 8 shows that
magnetite hosted in magnesian skarn has two distinct sub-populations with characteristically different Mg
concentrations illustrated by the split at 0.48 in the probability plot. Nadoll et al. (submitted) have linked the two
populations to magnetite from outside and inside the chalcocite enrichment zone of the Chino (Santa Rita, USA) skarn.
Similarly magnetite from calcic skarn can be divided into sub-populations based on characteristic variations in Mg and
Mn. Low-Mg (< 1 wt%) and low-Mn (90% of sub-population below 2000 ppm) skarn magnetite also shows a tendency
to incorporate less Al and Ti than the rest of the skarn magnetites. This trend could not yet be linked to a specific
16

ACCEPTED MANUSCRIPT
locality, host rock, or alteration zone but it has to be noted that most Ertsberg skarn magnetites fall outside the lowMg-Mn population.

4.1.4

Porphyry Cu-(Au)-(Mo) deposits hydrothermal and igneous (840 and 100 analyses)

Igneous and hydrothermal magnetite from the porphyry Cu-(Au)-(Mo) deposits display similar trends for several

RI
P

magnetite elements. Median and upper threshold Al, Mn, Zn, Ga and Sn concentrations are similar, and, although
those elements can help to classify local sample sub-populations, their use as overall discriminators seems to be

SC

limited. Distinguishing compositional features that help discriminate between magnetite of igneous and hydrothermal
origin are described in the following section. In comparison to magnetite from other settings, igneous porphyry

NU

magnetite has the highest median V and Cr concentrations and the highest upper threshold value for Ti (7.63 wt%, Fig.
7). Hydrothermal porphyry magnetite has characteristically high Co and Ni concentrations that are comparable only to

MA

those found in hydrothermal skarn magnetite. Both igneous and hydrothermal porphyry magnetite have median V
concentrations that exceed those found in other types of magnetite by an order of magnitude. Furthermore,

ED

characteristically high Ga contents are a feature for both types, and are only exceeded by igneous magnetite from

PT

Henderson Climax-Mo deposit.

4.1.4.1 Hydrothermal vs. igneous porphyry magnetite

AC
CE

When comparing hydrothermal and igneous magnetite it is important to acknowledge that many important spinel
elements such as Ti, V and Cr can be highly variable among different types of igneous host rocks and reflecting a wide
range of whole rock compositions and cooling histories. Magnetite from mafic and ultramafic igneous rocks often
have higher concentrations of these elements magnetite crystallizing from felsic to intermediate magmas (e.g. Dare et
al., 2012, Grigsby, 1990, Lindsley, 1991). Magnesium, Ti, V, Cr, and Co are useful discriminator elements that can
differentiate porphyry magnetite of igneous and hydrothermal origin. Titanium is one of the best discriminators: only
ten percent of hydrothermal porphyry magnetites contain more than 1 wt% Ti, whereas thirty percent of igneous
magnetite from porphyry deposits contain more than 1 wt% Ti. Similarly, lesser V concentrations are more indicative
of hydrothermal porphyry magnetite than for igneous porphyry magnetite. Approximately fifty percent of
hydrothermal porphyry magnetites have V concentrations below 1000 ppm, whereas only twenty percent of igneous
porphyry magnetites have V concentrations below that level (Fig. 8). Chromium concentrations are greater in igneous
porphyry magnetite than in hydrothermal occurrences. Hydrothermal porphyry magnetite has median Cr
concentrations of 370 ppm, whereas igneous porphyry magnetite has a median Cr content of 544 ppm. The difference
17

ACCEPTED MANUSCRIPT
is more evident in their upper threshold values: 3290 ppm for igneous and 426 ppm for hydrothermal magnetite (Fig.
9, Table 4). Cobalt concentrations display the opposite trend, with greater concentrations in hydrothermal porphyry
magnetite. Approximately sixty percent of hydrothermal porphyry magnetites have Co concentrations greater than

100 ppm whereas Co concentrations in excess of 100 ppm could only be recorded in two analyses (Fig. 8). Magnesium

RI
P

concentrations can also help to distinguish hydrothermal from igneous porphyry magnetite. Hydrothermal
occurrences display a much larger range of Mg concentrations, with concentrations above 1000 ppm and below 100

SC

ppm being indicative for hydrothermal occurrences. Igneous porphyry magnetite shows less variability in Mg
concentrations (100-1000 ppm) (Table 4, Fig. 7).

Plutonic igneous rocks

NU

4.1.5

MA

4.1.5.1 Climax-type Mo (Henderson, Red Mountain) (41 analyses)


Igneous magnetite from the Henderson Climax-type Mo deposit has one of the most distinct compositional signatures.

ED

Gallium concentrations are an order of magnitude higher than in any other type of magnetite and Sn is only present at
comparable concentrations in hydrothermal magnetite from magnesian skarn. Median Sn concentrations reach 88.2

PT

ppm and the upper threshold value is 175 ppm. Tin is an element that is not commonly present at concentrations
exceeding trace levels (10 ppm) and therefore the high concentrations in igneous magnetite from Henderson are

AC
CE

very characteristic and can serve as a unique compositional fingerprint. Molybdenum could not be detected. Median
Ti concentrations are the highest in the dataset, and Al and Mn concentrations are greater than in most other settings
and are only exceeded by hydrothermal magnetite from skarn deposits. Chromium could not be detected in magnetite
from the Henderson Climax-type Mo deposit. Overall, the minor and trace element signature of igneous magnetite
from Henderson is distinctive and can be easily discriminated from other magnetite occurrences. However, due to a
very limited number of samples it is an open question if this distinctive signature in igneous magnetite is also reflected
in hydrothermal magnetite from this deposit. If so that would be a clear indicator for a strong host rock control.

4.1.5.2 Inner Zone Batholith (12 analyses)


Minor and trace element concentrations in igneous magnetite from the Inner Zone Batholith (IZB) show similar trends
to other igneous magnetites (Fig. 7, Fig. 8, Table 4). Concentrations of elements such as Mg, Al, Co, Ni, Zn and Ga are
comparable to those found in igneous porphyry magnetite. Igneous magnetite from the Inner Zone Batholith is
characteristically high in V, with the highest reported median concentration of 3600 ppm. Titanium and Cr

18

ACCEPTED MANUSCRIPT
concentrations on the other hand are significantly lower compared to porphyry magnetite. Median concentrations for
these elements lie at 168 ppm and 68.2 ppm, respectively.

4.1.6

Magnetite from iron oxide copper-gold deposits

Iron oxide copper-gold (IOCG) deposits have recently undergone intense study (e.g. Barton and Johnson, 2004, Sillitoe,

RI
P

2003, Williams et al., 2005). Magnetite is an abundant mineral in many IOCG orebodies, and, although we present no
data for magnetite from IOCG deposits, several studies have reported chemical trends in magnetite from these

SC

deposits. The IOCG field defined by Dupuis and Beaudoin (2011) in their Ti+V vs. Al+Mn diagram plots at slightly lower
Ti+V and Al+Mn concentrations than those commonly found in hydrothermal magnetite from porphyry deposits (Fig.

NU

12). Carew (2004) investigated magnetite from the Cloncurry region and the Ernest Henry IOCG deposit. He found
characteristically high Sn and Mn concentrations and low V, Ti, Mg, Si, Cr, and Zn contents compared to barren

MA

magnetite occurrences. Rusk et al. (2009) and Rusk et al. (2010) reported a common incorporation of Mg, Al, Ti, V, Cr,
Mn, Co, Ni, Zn and Ga at detectable concentrations and confirmed the high Mn concentrations in hydrothermal

ED

magnetite from the Ernest Henry deposit. However, their results showed that hydrothermal magnetite associated
with mineralization is not relatively depleted in V compared to barren hydrothermal occurrences. Rusk et al. (2010)

PT

show that Ti/Mn ratios in hydrothermal magnetite present the best discriminator between barren and mineralized
systems, and between different deposits. However, Zhang et al. (2009) point out that no systematic compositional

AC
CE

trends exist in magnetite from IOCG deposits that directly link minor and trace element signatures of magnetite with
paragenesis, rock type, or mineralization.

4.2 Classification of orebody magnetite


A number of researchers have suggested diagrams to discriminate magnetite from different types of mineral deposits
(Beaudoin and Dupuis, 2009, Dupuis and Beaudoin, 2011, Kamvong et al., 2007, Nadoll et al., 2012, Nadoll et al.,
submitted, Singoyi and Zaw, 2001). Nadoll et al. (2012) present data for low temperature hydrothermal and burial
metamorphic magnetite from the Belt Supergroup. They show that variations in minor and trace elements between
these two types of magnetite are subtle but can be resolved using multi-element radar plots and factor score plots.
They established five factors that can discriminate between igneous, hydrothermal, and burial metamorphic
magnetite, namely (F1) Mg-Mn, (F2) Ga-Zn-Cr, (F3) Co-Ni-V, (F4) Al, and (F5) Ti. Their data for hydrothermal magnetite
from Ag-Pb-Zn veins, which is included in this study, suggest that hydrothermal magnetite from mineralized areas is
enriched in Mg and Mn in comparison to the host rock magnetite. Singoyi et al. (2006) and Kamvong (2007) proposed
19

ACCEPTED MANUSCRIPT
that Sn/Ga can be used to discriminate different types of hydrothermal magnetite. Our data show a distinct grouping
of the various hydrothermal magnetite populations (Fig. 11), however, Sn concentrations commonly lie below
detectable limits or, when the analytic resolution yields data, concentrations lie below 10 ppm. Dupuis and Beaudoin

(2011) recently proposed that Ni+Cr vs. Si+Mg, Al/(Zn+Ca) vs. Cu/(Si+Ca), Ni/(Cr+Mn) vs. Ti+V, and Ca+Al+Mn vs. Ti+V

RI
P

diagrams can be used to discriminate among a wide variety of mineral deposits, including iron oxide-copper-gold
(IOCG), banded iron formation (BIF), porphyry Cu, Fe-Cu skarn, Ni-Cu-PGE, Cu-Zn-Pb volcanogenic massive sulfide

SC

(VMS), and Archean porphyry Cu-Au. To test these suggested deposit fields we plotted our data in their Ni/(Cr+Mn) vs.
Ti+V and (Ca)+Al+Mn vs. Ti+V diagrams (Fig. 12). We decided to omit Ca because it contributes negligible amounts to

NU

the overall cation sum - typically one to two orders of magnitude lesser concentrations than the Al+Mn cation sum.
Furthermore, LA-ICP-MS detection limits for Ca are often high and due to the relatively large LA beam size Ca is

MA

commonly attributed to inclusions where present at greater concentrations (Nadoll and Koenig, 2011, Nadoll et al.,
2012). However, other workers have used Ca concentrations in magnetite determined by trace mode EMPA to
discriminate skarn and porphyry deposits (e.g. Dupuis and Beaudoin, 2011, Ray and Webster, 2007). Median values for

ED

the individual types of hydrothermal magnetite roughly confirm the porphyry and skarn fields laid out by Dupuis and

PT

Beaudoin (2011). However, in particular hydrothermal porphyry magnetite extends into the neighboring IOCG, skarn,
Kiruna, and Fe-Ti, V fields. Hydrothermal magnetite from BIF-Iron ore plots at significantly lesser Ti+V values and

AC
CE

extents past their proposed BIF field on the Al+Mn axis by several orders of magnitude. We suggest extending the BIF
field toward lower Ti+V and Al+Mn values (Fig. 12A). The graph in Fig. 12A also shows the skarn and porphyry fields
suggested by Nadoll et al. (submitted). Although hydrothermal magnetites from skarn are neatly enclosed by their
proposed field boundaries, the hydrothermal porphyry magnetite extends into the skarn field. The corresponding data
that these deposit-type fields are based on, are also part of our database (Table 3).

Although element ratio diagrams can help to discriminate magnetite from different mineral deposits, their field
boundaries still need refinement and stringent verification. In Fig. 13A-D probability plots and boxplots show Ti/V and
(Ti+V)/(Al+Mn) ratios for magnetite from the variety of investigated mineral deposits and geological settings. The plots
illustrate a distinct trend for two igneous magnetite occurrences. The highest Ti/V ratio (> 5) is present in igneous
magnetite from Henderson, whereas the lowest Ti/V ratio (<0.1) can be found in igneous magnetite from the barren
Inner Zone Batholith. Igneous and hydrothermal magnetites from porphyry deposits show a similar Ti/V ratio with
over 90% of the data plotting below 10, and median ratios of 0.9 and 2.1 respectively. Hydrothermal magnetite from
BIF (median 3.5) and Ag-Pb-Zn veins (median 5.9) also show similar Ti/V ratios. Ti/V ratios for hydrothermal skarn
20

ACCEPTED MANUSCRIPT
magnetite are the third highest in the dataset with a median value of 4.3. Apart from hydrothermal porphyry
magnetite which plots at similar Ti+V/Al+Mn ratios to igneous magnetite from barren and mineralized host rocks (~1),
other hydrothermal magnetite, from calcic and magnesian skarn, BIF and Ag-Pb-Zn veins, has distinctly lower ratios -

between 0.01 to 0.5) (Fig. 13C-D).

RI
P

Upper threshold values are a robust approach when comparing different types of hydrothermal and igneous
magnetite analyzed by different labs and under different operating conditions. They are usually not affected by

SC

varying detection or reporting limits across a combined dataset. Upper threshold values for the presented data span
two to four orders of magnitude and display several compositional trends among the different types of magnetite (Fig.

NU

9). There is, for example, a distinctive difference between igneous and hydrothermal magnetite in regard to their
upper threshold concentrations for Mg. Igneous magnetite has values in the lower 1000 ppm to upper 100 ppm

MA

region, whereas hydrothermal magnetite across all occurrences lies at or above 1 wt%. The highest upper threshold
values for Mg as well as for Mn, Co, Ni, and Zn can be found in hydrothermal magnetite from skarn. Upper threshold

ED

values for Ti, V, Cr, Co, Ni, and Ga are lowest in hydrothermal BIF-Iron ore magnetite. Igneous magnetite from
Henderson has the lowest upper threshold value for Mg and the highest for Ga among all occurrences. Igneous

PT

porphyry magnetite can be discriminated from the other types of magnetite by its high upper threshold values for Ti
and Cr. Vanadium, Mn, Ni, Zn, and Ga values in igneous porphyry magnetite are comparable to upper threshold values

AC
CE

observed in hydrothermal porphyry magnetite. The overall trends in upper threshold values for the different settings
can be summarized as follows: (I) BIF (hydrothermal) low Al, Ti, V, Cr, Mn, Co, Ni, Zn, Ga, Sn; (II) Ag-Pb-Zn veins
(hydrothermal) high Mn, low Ga, Sn; (III) Mg-skarn (hydrothermal) high Mg, Mn, low Al, Ti, Cr, Co, Ni, Ga; (IV) skarn
(hydrothermal) high Mg, Al, Cr, Mn, Co, Ni, Zn, low Sn (V) porphyry (hydrothermal) high Ti, V, low Sn; (VI) porphyry
(igneous) high Ti, V, Cr, low Mg; (VII) Climax-Mo (igneous) high Al, Ga, Sn, low Mg, Cr. We propose a combination
of the discussed discriminant plots and upper threshold concentrations to discriminate hydrothermal and igneous
magnetite from different deposit types.

One of the primary objectives of studies investigating the chemistry of magnetite from an exploration perspective is to
reliably discriminate mineralized from barren areas. As discussed, a number of authors have put forward
compositional trends to differentiate among ore deposits (e.g. Angerer et al., 2012, Dare et al., 2012, Dupuis and
Beaudoin, 2011, Nadoll et al., 2012, Nadoll et al., submitted). However, further research is required to address that
most fundamental question How can the chemistry of magnetite help to clearly distinguish prospective areas from

21

ACCEPTED MANUSCRIPT
barren background? Hereby, a profound understanding of the factors that control the chemistry of magnetite is
required.

4.3 Controlling factors and processes for the composition of magnetite

RI
P

In the following section we review and discuss the various factors and processes that can directly or indirectly affect
the composition of igneous and hydrothermal magnetite. It has not yet been experimentally determined whether the

SC

same factors that control the composition of high temperature igneous magnetite apply to cooler temperature
hydrothermal magnetite. However, first-order principles suggest that matrix (melt/fluid) composition, temperature,

NU

cooling rate, pressure, oxygen fugacity, sulfur fugacity, and silica activity, which apply to igneous conditions (6001200C), also apply equally to cooler temperature hydrothermal and metamorphic conditions (200-650C). Relative

MA

variations among elements at igneous temperatures should also be seen in magnetites formed under cooler
temperature hydrothermal conditions. Elements with larger partition coefficients for magnetite-melt pairs than other

4.3.1

Crystallographic factors

ED

magnetite elements are likely to display comparable trends under hydrothermal conditions.

PT

Crystallographic factors such as ionic radius and the overall charge balance place further constraints on possible
substituting cations (Cornell and Schwertmann, 2003, Fleet, 1981, Goldschmidt, 1954, Wechsler et al., 1984) (Fig. 1). In
2+

3+

AC
CE

general, siderophile and lithophile elements that have similar ionic radii to Fe and Fe are more readily incorporated
into magnetite than chalcophile elements or cations of mismatching ionic radii when other controlling factors such as
temperature, matrix/melt/fluid composition, and oxygen fugacity are favorable. Dare et al. (2012) illustrate how the
partition coefficients for relevant magnetite elements vary with changing ionic radii.

4.3.2

Igneous processes

Petrogenetic implications of compositional variations in magnetite from igneous rocks of all compositions have been
the centre of many geochemical studies for more than fifty years (Anderson et al., 2008, Bagdasarov, 1989, Borisenko
and Lapin, 1972, Buddington et al., 1955, Buddington and Lindsley, 1964, Cornell and Schwertmann, 2003, Frost and
Lindsley, 1991, Ghiorso and Sack, 1991a, Grant, 1984/85a, Klemm et al., 1985, Klemme et al., 2006, Lee et al., 2005,
Lindsley, 1976b, Nielsen and Beard, 2000, Nielsen et al., 1994, Powell and Powell, 1977, Prins, 1972, Ramdohr, 1955,
Sauerzapf et al., 2008, Shcheka et al., 1980, Wones, 1989). Source rock or magma composition, temperature,
pressure, cooling rate, oxygen fugacity, sulfur fugacity, silica and sulfide activity govern the composition of different
22

ACCEPTED MANUSCRIPT
types of igneous magnetite (Buddington and Lindsley, 1964, Dare et al., 2012, Frost and Lindsley, 1991, Ghiorso and
Sack, 1991a, Haggerty, 1991a, Whalen and Chappel, 1988). Many studies have investigated magnetite-mineral
partition coefficients and magnetite-melt distribution coefficients are well established experimentally for igneous

magnetite for a range of conditions and melts/host rocks (EarthRef.org, 2013 and references therein, Ewart and

RI
P

Griffin, 1994, LaTourrette et al., 1991, Okamoto, 1979, Toplis and Corgne, 2002). These coefficients can vary across
three to five orders of magnitude for relevant magnetite elements depending mostly on temperature, host rock/melt

SC

composition, and oxygen/sulfur fugacity (Fig. 15, Table 5). The composition of the respective magnetite can also exert
a certain control on mineral/melt partition coefficients. For example, Ti rich magnetite has been described as a more

NU

likely host for Cu, Nb and Ta than Ti-poor magnetite (Nielsen and Beard, 2000, Simon et al., 2008). Most recently Dare
et al. (2012) presented a comprehensive compilation of magnetite/melt partition coefficients for igneous conditions.

MA

Magnetite partition coefficients for divalent cations such as Mg, Mn, Ni and Co have been shown to be largely
independent of oxygen fugacity (Toplis and Corgne, 2002). Instead their dependency is coupled with ionic radius, site

ED

occupancies, and composition of the coexisting melt. Silver, Au, and Pb have large ionic radii compared to Fe and
require coupled substitution. The incorporation of these cations in magnetite involves structural alterations of the

PT

cubic spinel cell and is therefore less likely (Fleet, 1981, Wechsler et al., 1984).

AC
CE

Concentrations of elements such as Cr, Ni, and Ti in igneous magnetite can significantly exceed those found in
hydrothermal magnetite. However, igneous magnetite can also have low to negligible concentrations of these
elements for a number of reasons other than the availability of the respective element in the melt. For example, subsolidus re-equilibration during cooling of plutonic rocks leads to low-Ti magnetite that is also depleted in other minor
and trace elements (Frost and Lindsley, 1991, Wones, 1989). This is particularly apparent in felsic plutonic rocks that
are commonly depleted in minor and trace elements (Czamanske et al., 1977, Grigsby, 1990). Furthermore, foreign
cations will partition strongly into coexisting mineral phases, such as sulfides and silicates, under specific T-fO2-fS2
conditions, resulting in depleted minor and trace element concentrations in the corresponding magnetite. Although
Cr, Ni, and Ti are good pointers for igneous magnetite it is important to consider the evolution of the host rock and its
mineral assemblage. For example, ilmenite exsolutions in igneous magnetite will render the surrounding magnetite Tipoor. Dare et al. (2012) have shown that large laser ablation beam sizes that incorporate the host magnetite as well as
any occurring ilmenite exsolutions can provide an approximation of the original magmatic composition of these
magnetites.

23

ACCEPTED MANUSCRIPT
Measured partition coefficients for igneous Fe-(Ti)-oxide minerals indicate that elements such as REE, Y, Sr, U, Th, Mo,
Sb, and W are highly incompatible (Klemme et al., 2006). Indeed there is little to no evidence in our data that these
elements are incorporated into magnetite at levels exceeding common LA-ICP-MS detection limits. However, other

authors have reported the incorporation of Mo and W into Fe-oxide phases, such as magnetite, at significant levels

RI
P

(Candela, 1997). Mahood and Hildreth (1983) published partition coefficients for Mo (D~2 to D~20) in igneous
magnetite that are comparable to those for other transition metals like Co and Mn (Table 5). The interdependencies

SC

among elements can have further affects on the incorporation of foreign cations in magnetite. For example, Simon et
al. (2008) reported that the addition of Ti to end-member magnetite can increase the incorporation of Au and Cu.

Hydrothermal processes

NU

4.3.3

It is often difficult to attribute specific element enrichments or depletions to one singular physicochemical factor. For

MA

example, skarn magnetite has the highest median concentrations for Mg, Al, Mn, Co, Ni, and Zn among all types of
magnetite, which might reflect one or all of the following: (1) elevated temperatures compared to other types of

ED

hydrothermal magnetite (Fig. 12B), (2) high concentrations of these elements in the original hydrothermal fluids,
and/or (3) strong fluid-host rock interaction. Experimental data that measure the relative significance of these factors

PT

are not yet available for hydrothermal magnetite. However, (1) formation temperature and (2) the composition of the
original hydrothermal fluid are less likely to be the defining physicochemical control for Fe-substitution in magnetite

AC
CE

because they match the temperatures and fluid compositions responsible for porphyry mineralization in the adjacent
igneous rocks. Instead, (3) significant fluid-host rock interaction is a more likely explanation for the enrichment of a
number of elements in skarn magnetite (Einaudi et al., 1981, Meinert, 1987, Nadoll et al., submitted). This case
exemplifies the complex controls on the geochemistry of magnetite which might be further complicated by multiple
successive stages of alteration and alteration zonation (Nadoll et al., 2010, submitted).

In spite of these complexities, compositional trends can be put into a process-oriented context and increasing
amounts of experimental and fluid inclusion data can be consulted to support these observations. For example,
several authors have shown that even low Au concentrations (sub-ppm) in magnetite can be significant because of the
relatively large grain size and abundance of Fe-Ti oxides in many hydrothermal ore deposits (e.g. Larocque et al.,
2002). Most recently Tauson et al. (2012) showed experimental evidence that the partition coefficient for Au between
magnetite and hydrothermal solution is 1.0 0.3 [450C, 1 kbar (100 MPa)], indicating that Au is not an incompatible
element in magnetite. In comparison the distribution coefficient of Au in magnetite is much lower than in pyrite and

24

ACCEPTED MANUSCRIPT
arsenopyrite (Tauson et al., 2011). Tauson et al. (2011) also observed that Au is mostly enriched in the outermost layer
of magnetite crystals (~500 nm).

4.3.3.1 Fluid composition element availability

Element availability in hydrothermal fluids provides a first order control for minor and trace element concentrations in

RI
P

hydrothermal magnetite. Data for spinel element concentrations in hydrothermal fluids, especially fluids related to
porphyry and skarn deposits, have become increasingly available. Rusk et al. (2009) and Baker et al. (2004) list

SC

concentrations of relevant elements that range from <100 to over 100 ppm for Cu and 100 to several 100 ppm for Mn,
Zn, Pb, and As. Ilton and Eugster (1989) suggested that even low concentrations of these elements in hydrothermal

NU

magnetite can be an indication for a strong enrichment of these elements in the corresponding hydrothermal fluid.
Element availability in hydrothermal fluids is a first order control for minor and trace element concentrations in

MA

hydrothermal magnetite. Elements such as Mg, Mn, Co, and Zn that are otherwise commonly incorporated into
magnetite at significant levels (100s-1000s ppm) under hydrothermal conditions can be used as indicators for the

ED

composition of a hydrothermal fluid if not offset by secondary effects such as host rock buffering. If a specific
hydrothermal magnetite occurrence is depleted in these elements it is likely that the fluid was equally depleted, if

AC
CE

the magnetite structure.

PT

other conditions such as oxygen and sulfur fugacity were favorable for the incorporation of these foreign cations into

Hydrothermal fluids can contain spinel elements such as Mn, Fe, Cu, Zn, As, Mo, Ag, Au, and Pb (Kesler, 2005, Meinert
et al., 2005, Roedder and Bodnar, 1997, Yardley, 2005). These elements are commonly highly enriched in fluids
compared to their concentration in igneous host rocks (Audetat et al., 2000, Audetat et al., 2008, Hedenquist and
Lowenstern, 1994, Rusk et al., 2004, Ulrich et al., 1999). Fluids associated with skarn deposits are commonly enriched
in K, Ca, Mn, Fe, Cl, Br, Cu, Pb, Zn, and As (e.g., Meinert et al., 2005). Manganese and Zn are commonly incorporated
into hydrothermal magnetite at levels above 100 ppm, whereas the chalcophile elements Cu, As, Mo, Ag, Au and Pb
typically have concentrations that are less than their detection limits.

Tin is rarely present at concentrations above 100 ppm and, if detectable, is incorporated at levels at or below 10 ppm
in igneous or hydrothermal magnetite of all origins. However, igneous magnetite from the Henderson Climax-type Mo
deposits and hydrothermal magnetite from magnesian skarn display consistently higher Sn concentrations, exceeding
100 ppm (Table 4, Fig. 11), which is most likely controlled by the composition of the respective fluid or fluid/host rock
interaction. Sodium, Si, S, Ca, Cu, Ge, As, Se, Sr, Y, Mo, Ag, In, Sb, Ba, Ce, Au, Tl, Pb, and Th- are not commonly present
25

ACCEPTED MANUSCRIPT
(tens to hundreds of ppm) in magnetite from porphyry and associated skarn deposits in spite of their availability
(Nadoll et al., submitted). Other authors have reported Cu, As, Mo, Ag, Au and Pb to be incorporated in magnetite
from hydrothermal ore deposits (Angerer and Hagemann, 2010, Carew et al., 2006, Lee et al., 2005, Lindsley, 1991,

Simon et al., 2008), but reported concentrations are typically at or below the 100 ppm limit for Cu, As, Mo, and Pb and

RI
P

even lower for Au, and Ag. Those concentrations still lie at or below the quantification limit for many analytical
techniques. Elements such as Na, Sr, Y, Ba, Ce, Tl, and Th are highly incompatible in magnetite (e.g., Nielsen et al.,

SC

1994) and even though elements such as Na and Sr may have enriched concentrations in hydrothermal fluids, they are
not likely to be incorporated into magnetite (Fig. 1).

NU

Seawater is the mineralizing fluid for primary and metamorphic (pre-hydrothermal alteration) iron oxides in BIF. In
terms of absolute element concentrations there is no strictly coherent discrimination possible between Algoma- and

MA

Superior-type BIF. The Algoma dataset (Koolyanobbing, Marda, and Weld Range greenstone belts) tends to have
greater Ga and lesser Ti and Mn than the Superior dataset (Hamersley Province and QF), but there is considerable

ED

overlap. A consistently lower Ti/V ratio in Algoma-type BIF magnetite (Fig. 13, Fig. 14A) may reflect changing
availability of weathered material, from V-enriched ultramafic to Ti-enriched mafic, or changed oxidation states of

PT

seawater that led to variation of Ti and V solubility. Similarly, high concentrations of Mn (3700 ppm), Ti (800 ppm), Ni
(580 ppm), Cr (39 ppm), and V (17 ppm) in magnetite from the Garumahishani silicate facies iron-formation from the

AC
CE

Archean Iron Ore Group of JharkhandOrissa region, India, is attributed to the mixing of mafic-ultramafic debris
derived from weathering of oceanic crust, and/or post-sedimentary influx of high-temperature metalliferous brines
produced from fluid-rock interaction in the ultramafic to mafic oceanic crust (Bhattacharya et al., 2007). Pecoits at al.
(2009) analyzed magnetite in BIF from the Brockman Iron Formation by LA-ICP-MS revealing that Ni/Cr ratios show
close relationship to underlying mafic igneous rocks, providing evidence of a mafic-dominated source. Similarly, an
ultramafic metal source has been inferred for the Paleoarchean Isua BIF that shows high Ni/Cr ratios. The Ti/V ratio is
partially disturbed in hydrothermally altered magnetite from BIF-hosted iron ore deposits (Fig. 14B), In the Hamersley
and India dataset hydrothermally altered magnetite is comparatively low in Ti and high in V.

4.3.3.2 Temperature
Temperatures recorded for hydrothermal alteration and vein formation generally lie below magmatic temperatures.
They are approximately 100 to 300C for BIF-hosted high-grade ore (Thorne et al., 2008), 250 to 350C in the Coeur
dAlene district (e.g., Doughty and Chamberlain, 1996, Hayes, 1990, Landis and Hofstra, 1991, Leach et al., 1988,

26

ACCEPTED MANUSCRIPT
Nadoll et al., 2012), 300 to 800C in porphyry deposits, and 300 to 500C in the associated skarn deposits (e.g., Ahmad
and Rose, 1980, Audetat and Pettke, 2006, Beane and Titley, 1981, Manske and Paul, 2002, Sillitoe, 2010) (Fig. 12B).
Aluminum and Ti concentrations in magnetite are directly linked with temperature in igneous systems (Nielsen et al.,

1994, Toplis and Carroll, 1995), with greater concentrations indicating hotter temperatures. Even though Ti, and to

RI
P

some degree Al, seem to follow this trend for hydrothermal magnetite from BIF-related ore and Ag-Pb-Zn veins,
concentrations of these elements in hydrothermal porphyry and skarn magnetite are comparable to those found in

SC

igneous magnetite. This suggests that temperatures in these settings were not significantly different or that the
concentrations observed are not solely controlled by temperature. On a deposit scale Ti and V concentrations can be

NU

two to three orders of magnitude lower in hydrothermal magnetite from calcic and magnesian skarn than in most
igneous magnetites (Nadoll et al., submitted, Ray and Webster, 2007). Elements such as Al and Ti are considered to be

MA

immobile at sub-magmatic temperatures (Van Baalen, 1993, Verlaguet et al., 2006), and are typically not detected in
hydrothermal fluids. Their incorporation in magnetite is largely temperature controlled (Nielsen et al., 1994, Toplis
and Carroll, 1995) and therefore more likely under high temperature igneous conditions. Our data confirm this

ED

assumption and show that Al and Ti have on average considerably greater concentrations in igneous magnetite than

PT

for hydrothermal occurrences (Fig. 12B). For example, Ti concentrations are two to three orders of magnitude lower in
hydrothermal magnetite from calcic and magnesian skarn than in most igneous magnetites. Similar trends have been

AC
CE

observed in magnetite from the Heff Cu-Au skarn of British Columbia (Ray and Webster, 2007). The affiliation of
elements such as Ti with hotter-temperature formation conditions can have implications for elements such as Au, Ag,
and Cu that are more readily incorporated into ulvspinel than magnetite due to its greater Ti contents (Simon et al.,
2008). Magnetite has been described as an important host for Au in igneous rocks and hydrothermal ore deposits
(Gammons and Williams-Jones, 1997, Larocque et al., 2002, Togashi and Terashima, 1997). The trend toward higher Ti
concentrations with hotter formation temperatures can have repercussions on the incorporation of elements that are
directly linked with Ti.

Partition coefficients are greatly temperature dependent (McIntire, 1963). The data presented here show a trend that
cooler temperature hydrothermal magnetite generally has lesser trace element concentrations. High-temperature
hydrothermal porphyry and skarn magnetite shows relatively higher minor and trace element concentrations that for
many elements are comparable with those found in igneous magnetite, whereas hydrothermal-metamorphic
magnetite from BIF represents the low end of the temperature spectrum, and has the lowest overall minor and trace

27

ACCEPTED MANUSCRIPT
element concentrations among all investigated magnetites. Therefore, temperature must be considered a major
governing factor for the composition of hydrothermal magnetite.

Magnetite can accommodate measurable trace element contents at temperatures below 100C (Sidhu et al., 1978).

Re-equilibration of hydrothermal magnetite can also have significant repercussions on minor and trace element

RI
P

concentrations. Foreign cations in magnetite can be continuously expelled by this process. This is analogous to subsolidus re-equilibration of plutonic magnetite during cooling (Frost, 1991c).

SC

Although only semi-quantitative, there is a visible temperature trend in the data. Fig. 12B shows a Ti+V vs. Al+Mn plot

NU

where data points have been attributed with approximate temperatures from low-temperature BIF and related iron
ore (Thorne et al., 2008), medium temperature Ag-Pb-Zn veins (e.g., Doughty and Chamberlain, 1996, Hayes, 1990,

MA

Leach et al., 1988), medium to high temperature hydrothermal skarn and porphyry magnetite, to high temperature
igneous magnetite from porphyry and Climax-Mo deposits (e.g., Ahmad and Rose, 1980, Audetat and Pettke, 2006,
Beane and Titley, 1981, Manske and Paul, 2002, Sillitoe, 2010). Decreasing formation temperatures are reflected in

ED

the chemistry of the corresponding magnetite with high temperature igneous magnetite plotting at high Ti+V and high
Al+Mn and low temperature hydrothermal-metamorphic BIF magnetite plotting at low Ti+V and Al+Mn values.

PT

Gallium and Sn concentrations seem to follow this trend with the highest concentrations in high temperature igneous

AC
CE

magnetite from the Henderson Climax-type Mo deposits and the lowest concentrations in BIF magnetite (Fig. 11, Fig.
12). The temperature dependence of Ti, V, Al, Mn, and Ga can distinguish between low and high T mineral systems.
However, the other controlling factors discussed here can also exert a significant control on cation substitution and in
some cases mask other process-related compositional trends. For example, seawater composition and/or seawater
oxygen fugacity for magnetite formed in BIF, or re-equilibration processes that can expel cations over time.

4.3.3.3 Oxygen and sulfur fugacity


Oxygen fugacity within a melt is the major control on element partitioning between oxides, and between oxides and
silicates at magmatic temperatures (e.g., Frost, 1991a). Elements, such as Cr and V, can occur in various valence states
and therefore their behavior is strongly linked to oxygen fugacity (Klemme et al., 2006, Nielsen et al., 1994, Righter et
al., 2006b, Ryabchikov and Kogarko, 2006) (Fig. 1). Minor and trace element compositions of oxides such as
titanomagnetite are highly dependent on the prevailing oxygen fugacity (Andersen and Lindsley, 1988, Buddington
and Lindsley, 1964, Ghiorso and Sack, 1991b, Toplis and Carroll, 1995). However, the effect at cooler temperature
hydrothermal conditions is less well constrained. Several studies reported that the incorporation of foreign cations is
28

ACCEPTED MANUSCRIPT
more likely at lower oxygen fugacity (Buddington and Lindsley, 1964, Frost, 1991a, Lindsley, 1991). Vanadium is
usually preferentially incorporated into the magnetite structure at low oxygen fugacity and becomes incompatible at
high oxygen fugacity due to its 5+ oxidation state (Bordage et al., 2011, Toplis and Corgne, 2002, Toplis and Carroll,

1995). Mallmann and ONeill (2009) demonstrate that the partitioning behavior of homovalent cations such as Ga

RI
P

however, is not dependent on the prevailing oxygen fugacity conditions.

In sulfide dominated environments sulfur fugacity can be the primary control for the partitioning behavior of relevant

SC

elements (Dare et al., 2012) and it can affect Fe/Mg ratios in associated minerals (Heiligmann et al., 2008). Coexisting
sulfide minerals have considerably higher partition coefficients than magnetite for chalcophile elements such as Cu,

NU

As, Ag, and Pb (e.g., Cygan and Candela, 1995, Gaetani and Grove, 1997, Lemarchand et al., 1987, Simon et al., 2008,
Stimac and Hickmott, 1994). The data presented in the present study show no evidence that Cu, As, Ag, and Pb are

MA

widely incorporated at detectable levels into magnetite under the conditions found in the examined geological
settings. The chalcophile element Zn seems to be an exception, as it is commonly present at detectable levels in

ED

hydrothermal magnetite (several 10-100 ppm), and in some geological settings, it can also form Zn spinel, gahnite
(ZnAl2O4). This trend seems to be in agreement with published Zn partition coefficients for magnetite-matrix pairs (Fig.

PT

15, Table 5) that are considerable higher than those for Cu or As (e.g. Ewart and Griffin, 1994). Calculations based on
fluid inclusion data support the argument for low concentrations of chalcophile elements such as Cu in hydrothermal

AC
CE

magnetite (Ilton and Eugster, 1989, Sawkins and Scherkenbach, 1981) with a theoretical maximum of 10 ppm under
the temperature and oxygen fugacity conditions commonly associated with porphyry Cu deposits (690 C, oxygen
fugacity: NNO buffer). This might be an explanation why Cu is usually not reported to have significant concentrations
in magnetite.

Siderophile elements such as Ni and in particular Mo are restricted to very low oxygen and sulfur fugacity conditions
because they are otherwise more likely to form pentlandite or molybdenite respectively (Drabek, 1982). Similarly,
platinum group elements (PGEs) are more likely to partition into sulfide minerals than into magnetite (e.g., Barnes et
al., 2004).

4.3.3.4 Silica and sulfide activity


Interdependencies among coexisting mineral phases can be crucial for trace elements concentrations in magnetite.
Analogous to oxygen and sulfur fugacity, coexisting silicates and sulfides exert a strong control on the composition of
hydrothermal magnetite (e.g., Cygan and Candela, 1995, Frost et al., 1988, Gaetani and Grove, 1997, Lemarchand et
29

ACCEPTED MANUSCRIPT
al., 1987, Schilling et al., 2011, Simon et al., 2008, Stimac and Hickmott, 1994). For example, sulfide minerals will
preferentially incorporate chalcophile elements (e.g. Cygan and Candela, 1995, Simon et al., 2008) whereas lithophile
elements such as Mg, Al and Ti have a strong preference to partition into silicates if conditions are favorable (Frost,

1991b, Simon et al., 2008, Toplis and Corgne, 2002). Silicates also commonly have considerably higher partition

RI
P

coefficients for Pb than magnetite (Bea et al., 1994, Stimac and Hickmott, 1994). Measured partition coefficients for
ilmenite and ulvspinel indicate that elements such as Mo are highly incompatible, suggesting that their incorporation

SC

in magnetite is equally unlikely (Klemme et al., 2006). However, a number of studies have shown that Mo can be a
compatible element with partition coefficients as high as 29 (Table 5).

NU

4.3.3.5 Host rock buffering

Host rock buffering is another important control for compositional variations in hydrothermal magnetite, particularly

MA

where large scale metasomatic processes are involved. It has been shown that elements like Mg and Mn can
successively be enriched in hydrothermal fluids by extensive fluid/rock interactions (Einaudi et al., 1981, Meinert,

ED

1987). This is commonly reflected in high Mg and Mn concentrations in hydrothermal magnetite occurring in skarn. In
particular, magnetite from magnesian skarn shows this trend and points towards host rock buffering as a major

PT

control for hydrothermal magnetite in metasomatic environments such as skarn deposits. Ilton and Eugster (1989)
suggest that the formation of massive magnetite in skarn can lead to significant enrichment of Mn relative to Fe in the

AC
CE

remaining hydrothermal fluid.

Compared to igneous magnetite, skarn contains magnetite with higher Mg, Mn and Zn concentrations (McQueen and
Cross, 1998, Nadoll et al., submitted, e.g., Rusk et al., 2009). This may reflect the primary fluid composition or host
rock buffering via significant fluid/host rock interaction (Baker et al., 2004, e.g., Einaudi et al., 1981, Meinert, 1987,
Nadoll et al., submitted).

In BIF-hosted high-grade iron ore deposits, magnetite is commonly chemically and texturally altered by hydrothermal
fluid-rock interaction. Hence, magnetite can be used as an indicator mineral for hydrothermal alteration (Angerer et
al., 2012). Hydrothermal alteration of metamorphic magnetite has been discussed for the distinct magnetite chemistry
in the Neo Dala BIF in the Archean Iron Ore Group by Bhattacharya et al. (2007). Here, elevated W, Pb, As, Mo and Sn
concentrations were attributed to hydrothermal metasomatism related to the granitic host rock. In the iron ore
deposits of the Koolyanobbing and Marda greenstone belts of the Yilgarn craton, the Co/Ni ratios and W
concentrations in hydrothermally modified magnetite in carbonate- and silicate-altered BIF are significantly greater
30

ACCEPTED MANUSCRIPT
than those in magnetite from associated least-altered BIF (up to 7 in altered versus < 1 in least-altered BIF) (Angerer et
al., 2013, Angerer et al., 2012). Because the Yilgarn craton deposits are hosted in granite-greenstone terranes, the
magnetite chemistry through the evolution of the iron ore deposits is attributed to changes from a more mafic-

ultramafic (Ni) to dominantly granitic (Co, W) host rock chemistry (Angerer et al., 2013, Angerer et al., 2012,

RI
P

Figueiredo e Silva et al., 2009).

SC

4.4 Pseudomorphic magnetite alteration

Magnetite grains commonly show textures, crystal habits, and oxidation-exsolution features, such as ilmenite lamellae

NU

and martitization along spinel planes, that reflect physicochemical formation conditions (e.g., Haggerty, 1991b,
McQueen and Cross, 1998, Mcke and Raphael Cabral, 2005). Ilmenite exsolutions (in igneous magnetite) and

MA

oxidation to hematite are common features in magnetite. The occurrence and abundance of these features must be
assessed by careful petrography prior to EPMA and LA-ICP-MS analyses. Grains that show only slight to moderate
martitization may be easily analyzed with the small EPMA beam size (<2 m). Laser ablation ICP-MS on the other

ED

hand, with considerably larger analytical spot sizes (commonly >10 m), can be adversely affected by martite and

PT

ilmenite in magnetite. However, magnetite with a low degree of martitization (< 10% of the grain surface) will not
negatively affect LA-ICP-MS analyses (Nadoll and Koenig, 2011, Nadoll et al., 2012).

Martitization and mushketovitization

AC
CE

4.4.1

If hydrothermal alteration affecting magnetite is caused by an oxidizing or acidic fluid, then martitization (hematite
after magnetite) may occur (Ohmoto, 2003). The reversed alteration, magnetite after hematite under reducing fluidrock reaction, is called mushketovitization.

Martitization and mushketovitization are commonly pseudomorphic processes but can, depending on the governing
oxygen fugacity and pH conditions, represent considerable volumetric changes in the rock (Mcke and Raphael Cabral,
2005). Two theoretically reversible chemical processes have been proposed for martitization and mushketovitisation:
one is a redox reaction (equation 1), and the other is a non-redox transformation (equation 2) under acidic conditions
(Ohmoto, 2003).

equation 1:

2Fe3O4 + 0.5O2aq = 3Fe2O3

equation 2:

Fe3O4 + 2H aq= Fe2O3 + Fe

2+
aq

+ H 2O

31

ACCEPTED MANUSCRIPT
The lack of significant cation exchange during hydrothermal martitization supports the oxidation reaction because
2+

there is little mobilization of Fe and associated cations.

Magnetite is commonly subject to martitization in weathering zones and supergene-modified ore deposits but also in

oxidative hydrothermal zones. It is one of the major minerals in BIF and iron ore deposits. The understanding of cation

RI
P

exchange between magnetite and oxidizing fluids during martitization is important. It has been suggested that most
spinel elements, with the exception of Mg, are retained during martitization under moderate-temperature

SC

hydrothermal oxidation (Angerer et al., 2012, Sidhu et al., 1981). Other workers suggest that the martitization process
can expel divalent cations because of their incompatible valency and ionic radii (Cornell and Schwertmann, 2003).

NU

Martite in regolith environments contains lesser Fe and Ni and greater Si, Al, Ti, Cr than parent magnetite, but Zn, V
and Mn remain relatively unchanged. Additionally Ti and Cr may substitute for Fe in the lattice, but secondary phases

MA

such as kaolinite, gibbsite and halloysite are the most likely hosts for Si and Al in supergene martite (Anand and Gilkes,

ED

1984).

5 Inclusions

PT

Mineral inclusions are common features in igneous and hydrothermal magnetite; they reflect magnetite formation

AC
CE

conditions and give a direct insight into the evolution of the corresponding host rock. Mineral inclusions are of
particular interest when geological and mineralogical context have been removed or significantly altered, which is the
case for stream sediments or regolith cover. Some inclusions, such as Cu sulfide minerals, can be a direct pointer
towards mineralized areas. Wang et al. (2012) recently reported that magnetite with cassiterite inclusions is diagnostic
of oxidized tin granites and hence magnetite from stream sediments, may serve as a strategic tracer for tin
exploration in those environments.

Despite their virtues for discriminating sample populations, mineral inclusions need to be thoroughly assessed in
because of their impact on EMPA and LA-ICP-MS analyses (Nadoll and Koenig, 2011). Critical assessment of whether
elements are in genuine solid solution or are contributed by micro- and nano-scale mineral inclusions is essential to
rigorously define the chemistry of magnetite. Magnetite commonly contains sulfide, phosphate, quartz, silicate, or
carbonate inclusions in many geological settings including burial metamorphic, BIF, igneous, and hydrothermal
environments (Fig. 6, Fig. 10). Laser ablation analyses of microdrilled (bulk) magnetite mesobands in the Neoarchean
Temagami BIF, Canada show large ranges of Fe (36 to 57 wt%), Si (17 to 34 wt%), Al, Mg, and Ca (up to 1 wt%), and P
32

ACCEPTED MANUSCRIPT
(660 to 2230 ppm), and therefore clearly represent mineral mixtures (Bau and Alexander, 2009). Inclusions such as
quartz, silicates (typically talc, stilpnomelane or amphiboles), and carbonates (siderite, magnesite, and ankerite) will
contribute Mg, Al, and Ca. Phosphates and sulfides contribute P, Ca, REE, S, and transition metals. In some cases, it is

difficult to ascertain whether inclusions are responsible for elevated concentrations of non-typical spinel elements.

RI
P

For example, magnetite from Henderson has significant concentrations of the high field strength (HFS) element Nb,
with median values of 19 ppm. It is debatable whether these Nb contents represent genuine solid solution, because

SC

Nb is not usually incorporated in magnetite at high concentrations - up to 248 ppm in individual samples from
Henderson. The Nb contents may be attributed to mineral inclusions of pyrochlore (approximate formula)

NU

(Na,Ca)2Nb2O6(OH,F) or columbite [(Fe, Mn)(Nb, Ta)2O6]. Rutile, which contains Nb impurities (Nb2O5) and other HFS
elements of up to several weight percent (Ghent, 1975, Waychunas, 1991), is also an abundant and widespread

MA

mineral at Henderson (Seedorff and Einaudi, 2004), and is a potential candidate for mineral inclusions in magnetite.
However, the relatively low Ti concentrations in magnetite from Henderson suggest that Sn and Nb concentrations are
not attributed to rutile inclusions, and careful petrography showed no rutile inclusions. Furthermore, Na and Ca

ED

concentrations could not be detected in samples that contain high Nb concentrations. These considerations suggest

PT

that Sn and Nb are incorporated in the magnetite structure at Henderson, rather than residing in mineral inclusions.
Although Si is typically not significantly incorporated into magnetite, Huberty et al. (2012) reported metamorphic

AC
CE

silician magnetite (1 to 3 wt% SiO2 in solid solution, determined by micro-XRD and transmission electron microscopy)
overgrowth on Si-poor magnetite in Dales Gorge BIF, Hamersley Province, and other BIF worldwide.

6 Concluding remarks
One of the main challenges for mineral exploration is the recognition of mineralized areas at increasingly greater
distance from primary mineralization. Minor and trace element data for magnetite from a variety of mineral deposits
are becoming more available, and help to further elucidate trends that can be used as geochemical signatures for
provenance studies and mineral exploration. Published and previously unpublished magnetite minor and trace
element data presented here give evidence for a number of systematic variations among hydrothermal magnetite
from different deposit types and igneous magnetite from mineralized and barren host rocks. The most important
factors that govern compositional variations in hydrothermal magnetite are (A) temperature, (B) fluid composition
element availability, (C) oxygen and sulfur fugacity, (D) silicate and sulfide activity, (E) host rock buffering, (F) reequilibration processes, and (G) intrinsic crystallographic controls such as ionic radius and charge balance.
33

ACCEPTED MANUSCRIPT
The main discriminator elements are Mg, Al, Ti, V, Cr, Mn, Co, Ni, Zn, Ga and Sn. Their common incorporation and
characteristic variations can be used to fingerprint magnetite from different host rocks and deposit types. The
classification scheme proposed by Dupuis and Beaudoin (2011) offers a good baseline for the discrimination of

magnetite from various types of mineral deposits and ongoing research will provide increasingly more data to refine

RI
P

their suggested deposit fields. We propose a combination of the available discriminant plots such as the Ti+v vs.
Al+Mn plot and box and whisker plots for the suite of magnetite elements can distinguish among magnetite from BIF,

SC

Ag-Pb-Zn veins, porphyry and associated skarn deposits, Climax-Mo deposits, and barren igneous host rocks. Low
levels of Al, Ti, V, Cr, Co, Zn, and Ga are particularly diagnostic for magnetite from BIF, Mg-skarn, and Ag-Pb-Zn veins.

NU

Igneous and hydrothermal magnetite from porphyry deposits have remarkably similar compositional patterns but can
be discriminated based on Cr, Co, Ni, and Ga concentrations. Gallium concentrations are consistently greater in

MA

igneous magnetite than in hydrothermal occurrences. Based on the established upper threshold concentrations for
the suite of magnetite elements the different deposit types display the following characteristics: (I) BIF (hydrothermal)
low Al, Ti, V, Cr, Mn, Co, Ni, Zn, Ga, Sn; (II) Ag-Pb-Zn veins (hydrothermal) high Mn, low Ga, Sn; (III) Mg-skarn

ED

(hydrothermal) high Mg, Mn, low Al, Ti, Cr, Co, Ni, Ga; (IV) skarn (hydrothermal) high Mg, Al, Cr, Mn, Co, Ni, Zn, low

PT

Sn (V) porphyry (hydrothermal) high Ti, V, low Sn; (VI) porphyry (igneous) high Ti, V, Cr, low Mg; (VII) Climax-Mo
(igneous) high Al, Ga, Sn, low Mg, Cr.

AC
CE

In combination with petrographic studies, the chemistry of hydrothermal magnetite can assist exploration geologists
to target prospective areas in regolith- and glacigenic environments or stream sediments distal to mineralization. The
use of mineral chemistry to characterize host rocks and fingerprint hydrothermal ore deposits, strongly benefits from
a sound petrographic description of magnetite, especially in terms of paragenetic sequence and foreign mineral
inclusions characterization. Further refinement of magnetite classification schemes and experimental data are needed
to gain a better understanding of how physicochemical factors, hydrothermal processes, and secondary enrichment
processes are reflected in the composition of the corresponding magnetite.

7 Acknowledgements
We thank the iron ore research group at UWA (CET) - Paul Duuring, Ana-Sophie Hensler, and Warren Thorne -for
kindly providing data for this project, and Steffen Hagemann for helpful feedback. Thomas Angerer thanks the
LabMaTer group at UQAC, Sarah-Jane Barnes, Sarah Dare, Dany Savard, Sadia Mehdi, and the CODES laser ablation
team at UTAS, Sarah Gilbert and Leonid Danyushevsky, for granting access to their laser systems and helping with
34

ACCEPTED MANUSCRIPT
analyses and data reduction. Further acknowledgments go to industry partners, Cliffs Natural Resources, Rio Tinto Iron
Ore, and BHP Billiton Iron Ore for granting access to samples. We also thank Freeport McMoRan, and in particular
Richard Leveille and Clyde Leys, for granting access to sites and providing drill hole documentation for the Ertsberg

RI
P

district. For helping us with the visual content of this paper we thank Travis Naughton.

SC

8 References

AC
CE

PT

ED

MA

NU

Ahmad SN, Rose AW. Fluid inclusions in porphyry and skarn ore at Santa Rita, New Mexico. Econ. Geol. 1980;75:229-50.
Akimoto S. Magnetic properties of ferromagnetic minerals contained in igneous rocks. Japanese Journal of Geophysics. 1955;1:1-31.
Anand RR, Butt CRM. A guide for mineral exploration through the regolith in the Yilgarn Craton, Western Australia. Aust. J. Earth Sci. 2010;57:1015114.
Anand RR, Gilkes RJ. Mineralogical and chemical properties of weathered magnetite grains from lateritic saprolite. J. Soil Sci. 1984;35:559-67.
Andersen DJ, Lindsley DH. Internally consistent solution models for Fe-Mg-Mn-Ti oxides; Fe-Ti oxides. Am. Mineral. 1988;73:714-26.
Anderson JL, Barth AP, Wooden JL, Mazdab F. Thermometers and Thermobarometers in Granitic Systems. Rev. Mineral. Geochem. 2008;69:121-42.
Angerer T, Hagemann S, Danyushevsky L. High-grade iron ore at Windarling, Yilgarn Craton: a product of syn-orogenic deformation, hypogene
hydrothermal alteration and supergene modification in an Archean BIF-basalt lithostratigraphy. Mineral. Deposita. 2013;48:697-728.
Angerer T, Hagemann SG. The BIF-Hosted High-Grade Iron Ore Deposits in the Archean Koolyanobbing Greenstone Belt, Western Australia:
Structural Control on Synorogenic- and Weathering-Related Magnetite-, Hematite-, and Goethite-rich Iron Ore. Econ. Geol. 2010;105:917-45.
Angerer T, Hagemann SG, Danyushevsky LV. Geochemical evolution of the banded iron formation-hosted high-grade iron ore system in the
Koolyanobbing greenstone belt, Western Australia. Econ. Geol. 2012;107:599-644.
Annersten H. A mineral chemical study of a metamorphosed iron formation in northern Sweden. Lithos. 1968;1:374-97.
Arancibia ON, Clark AH. Early magnetite-amphibole-plagioclase alteration-mineralization in the Island copper porphyry copper-gold-molybdenum
deposit, British Columbia. Econ. Geol. 1996;91:402-38.
Audetat A, Guenther D, Heinrich CA. Magmatic-hydrothermal evolution in a fractionating granite: a microchemical study of the Sn-W-F-mineralized
mole granite (Australia). Geochim. Cosmochim. Acta. 2000;64:3373-93.
Audetat A, Pettke T. Evolution of a porphyry-Cu mineralized magma system at Santa Rita, New Mexico (USA). J. Petrol. 2006;47:2021-46.
Audetat A, Pettke T, Heinrich CA, Bodnar RJ. Special Paper: The Composition of Magmatic-Hydrothermal Fluids in Barren and Mineralized Intrusions.
Econ. Geol. 2008;103:877-908.
Bacon CR, Druitt TH. Compositional Evolution of the Zoned Calcalkaline Magma Chamber of Mount-Mazama, Crater Lake, Oregon. Contrib. Mineral.
Petrol. 1988;98:224-56.
Bagdasarov YA. Composition of magnetites from carbonatite complexes and depth facies of carbonatites. In: Bagdasarov YA, Efimov AF, editors.
Forecasting and Evaluating Carbonatites. Moscow, USSR: IMGRE; 1989. p. 124-56.
Bailey DK, Kearns S. High-Ti magnetite in some fine-grained carbonatites and the magmatic implications. Mineral. Mag. 2002;66:379-84.
Bajwah ZU, Seccombe PK, Offler R. Trace element distribution, Co:Ni ratios and genesis of the Big Cadia iron-copper deposit, New South Wales,
Australia. Mineral. Deposita. 1987;22:292-300.
Baker T, Van Achterberg E, Ryan CG, Lang JR. Composition and evolution of ore fluids in a magmatic-hydrothermal skarn deposit. Geology.
2004;32:117-20.
Balan E, De Villiers JPR, Eeckhout SG, Glatzel P, Toplis MJ, Fritsch E, et al. The oxidation state of vanadium in titanomagnetite from layered basic
intrusions. Am. Mineral. 2006;91:953-6.
Barnes HL. Geochemistry of Hydrothermal Ore Deposits (third ed.). New York: John Wiley & Sons Inc.; 1997.
Barnes S-J, Maier WD, Ashwal LD. Platinum-group element distribution in the Main Zone and Upper Zone of the Bushveld Complex, South Africa.
Chem. Geol. 2004;208:293-317.
Barnes SJ, Roeder PL. The range of spinel composition in terrestrial mafic and ultramafic rocks. J. Petrol. 2001;42:2279-302.
Barton M, Johnson D. Footprints of Fe-oxide(-Cu-Au) systems. SEG 2004: Predictive Mineral Discovery Under Cover. Centre for Global Metallogeny,
Spec. Pub. 33, The University of Western Australia. 2004:112-6.
Bau M, Alexander BW. Distribution of high field strength elements (Y, Zr, REE, Hf, Ta, Th, U) in adjacent magnetite and chert bands and in reference
standards FeR-3 and FeR-4 from the Temagami iron-formation, Canada, and the redox level of the Neoarchean ocean. Precambrian Res.
2009;174:337-46.
Bea F, Pereira MD, Stroh A. Mineral/leucosome trace-element partitioning in a peraluminous migmatite (a laser ablation-ICP-MS study). Chem.
Geol. 1994;117:291-312.
Beane RE, Titley SR. Porphyry copper deposits, part II. Hydrothermal alteration and mineralization. Econ. Geol. 1981;75th Anniversary:235-69.
Beaudoin G, Dupuis C. Iron-oxide trace element fingerprinting of mineral deposit types. In: Corriveau L, Mumin H, editors. Exploring for Iron Oxide
Copper-Gold Deposits: Canada and Global Analogues, Short Course Volume, Geological Association of Canada Annual Meeting2009. p. 107-21.
Beaudoin G, Sangster DF. A descriptive model for silver-lead-zinc veins in clastic metasedimentary terranes. Econ. Geol. 1992;87:1005-21.
Behn G, Camus F, Carrasco P, Ware H. Aeromagnetic signature of porphyry copper systems in northern Chile and its geologic implications. Econ.
Geol. 2001;96:239-48.
Bhattacharya HN, Chakrabortym I, Ghosh KK. Geochemistry of some banded iron-formations of the Archean Supracrustals, JharkhandOrissa region,
India. J. Earth Syst. Sci. 2007;116:24559.
Bordage A, Balan E, Villiers JR, Cromarty R, Juhin A, Carvallo C, et al. V oxidation state in FeTi oxides by high-energy resolution fluorescencedetected X-ray absorption spectroscopy. Phys. Chem. Minerals. 2011;38:449-58.
Borisenko LF, Lapin AV. O kontsentratsiyakh elementov-primesey v titanomagnetite i magnetite endogennykh mestorozhdeniy razlichnykh tipov.
Trace-element concentrations in titanomagnetite and magnetite from endogene deposits of various types. Dokl. Akad. Nauk SSSR. 1971;196:1441-4.

35

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

SC

RI
P

Borisenko LF, Lapin AV. Trace-element concentrations in titanomagnetite and magnetite from magmatic, contact-metasomatic and hydrothermal
deposits of different types. Doklady. Earth Science Sections. 1972;196:217-20.
Bowles JFW, Howie RA, Vaughan DJ, Zussman J. Rock-forming minerals - Non-silicates: Oxides, hydroxides and sulphides. 2nd ed. London: Geological
Society; 2011.
Bragg WH. The structure of the spinel group of crystals. Phil. Mag. 1915;30:305-15.
Buddington AF, Fahey J, Vlisdis A. Thermometric and petrogenetic significance of titaniferous magnetite. Am. J. Sci. 1955;253:497-532.
Buddington AF, Lindsley DH. Iron-titanium oxide minerals and synthetic equivalents. J. Petrol. 1964;5:310-57.
Candela PA. A Review of Shallow, Ore-related Granites: Textures, Volatiles, and Ore Metals. J. Petrol. 1997;38:1619-33.
Carew MJ. Controls on Cu-Au mineralisation and Fe-oxide metasomatism in the Eastern Fold Belt, NW Queensland, Australia [PhD]. Townsville:
James Cook University; 2004.
Carew MJ, Mark G, Oliver NHS, Pearson N. Trace element geochemistry of magnetite and pyrite in Fe oxide (+/-Cu-Au) mineralised systems: Insights
into the geochemistry of ore-forming fluids. Geochim. Cosmochim. Acta. 2006;70:A83.
Carten RB, White WH, Stein HJ. High-grade granite-related molybdenum systems: Classification and origin. Geological Association of Canada Special
Paper. 1993;40:521-54.
Chacko T, Cole DR, Horita J. Equilibrium oxygen, hydrogen and carbon isotope fractionation factors applicable to geologic systems. In: Valley J. W.,
Cole DR, editors. Stable Isotope Geochemistry. Washington: Mineralogical Society of America; 2001. p. 1-81.
Chiba H, Chacko T, Clayton RN, Goldsmith JR. Oxygen isotope fractionations involving diopside, forsterite, magnetite, and calcite: Application to
geothermometry. Geochim. Cosmochim. Acta. 1989;53:2985-95.
Clark DA. A review of geological factors that control magnetic signatures of porphyry and epithermal deposits; towards predictive magnetic
exploration models. Vietnam 2001; IAGA-IASPEI first joint scientific assembly; abstracts. 2001.
Clout JMF. Upgrading processes in BIF-derived iron ore deposits: implications for ore genesis and downstream mineral processing. Transactions of
the Institution of Mining and Metallurgy Section B-Applied Earth Science. 2003;112:B89-B95.
Cornell RM, Schwertmann U. The iron oxides: Structure, properties, reactions, occurrences, and uses. 2 ed. Weinheim: Wiley-VCH; 2003.
Cygan GL, Candela PA. Preliminary study of gold partitioning among pyrrhotite, pyrite, magnetite, and chalcopyrite in gold-saturated chloride
solutions at 600 to 7008C, 140 MPa, 1400 bars. In: Thompson JFH, editor. Magmas, Fluids, and Ore Deposits, Short Course: Mineralogical
Association of Canada; 1995. p. 129-37.
Czamanske GK, Wones DR, Eichelberger J. Mineralogy and petrology of the intrusive complex of the Pliny Range, New Hampshire. Am. J. Sci.
1977;277:1073-123.
D'Orazio M, Armienti P, Cerretini S. Phenocryst/matrix trace-element partition coefficients for hawaiite-trachyte lavas from the Ellittico volcanic
sequence (Mt. Etna, Sicily, Italy). Mineral. Petrol. 1998;64:65-88.
Dare SAS, Barnes S-J, Beaudoin G. Variation in trace element content of magnetite crystallized from a fractionating sulfide liquid, Sudbury, Canada:
Implications for provenance discrimination. Geochim. Cosmochim. Acta. 2012;88:27-50.
Doughty PT, Chamberlain KR. Salmon River arch revisited: New evidence for 1370 Ma rifting near the end of deposition in the Middle Proterozoic
belt basin. Canadian Journal of Earth Science. 1996;33:1037-52.
Drabek M. The system Fe-Mo-S-O and its geologic application. Econ. Geol. 1982;77:1053-6.
Duan C, Li YH, Yuan SD, Hu MY, Zhao LH, Chen XD, et al. Geochemical characteristics of magnetite from Washan iron deposit in Ningwu ore district
and its constraints on ore-forming. Acta Petrol. Sin. 2012;28:243-57.
Dudas MJ, Harward ME, Schmitt RA. Identification of dacitic tephra by activation analysis of their primary mineral phenocrysts. Quat. Res.
1973;3:307-15.
Dupuis C, Beaudoin G. Discriminant diagrams for iron oxide trace element fingerprinting of mineral deposit types. Mineral. Deposita. 2011;46:31935.
Duuring P, Hagemann S. Leaching of silica bands and concentration of magnetite in Archean BIF by hypogene fluids: Beebyn Fe ore deposit, Yilgarn
Craton, Western Australia. Mineral. Deposita. 2013;48:341-70.
Duuring P, Hagemann SG. Leaching of silica bands and concentration of magnetite in Archean BIF by hypogene fluids: Beebyn Fe ore deposit, Yilgarn
Craton, Western Australia. Mineral. Deposita. 2012.
EarthRef.org. Geochemical Earth Reference Model (GERM) Partition Coefficient (Kd) Database. EarthRef.org; 2013.
Einaudi MT. Description of skarns associated with porphyry copper plutons. In: Titley SR, editor. Advances in geology of the porphyry copper
deposits. Tucson: The University of Arizona Press; 1981a. p. 139-83.
Einaudi MT. General features and origin of skarns associated with porphyry copper plutons. In: Titley SR, editor. Advances in geology of the
porphyry copper deposits. Tucson: The University of Arizona Press; 1981b. p. 185-209.
Einaudi MT, Meinert LD, Newberry RJ. Skarn deposits. Economic Geology 75th Anniversary Volume. 1981:317-91.
Enders MS. The evolution of supergene enrichment in the Morenci porphyry copper deposit, Greenlee County, Arizona. unpub. PhD dissert.,
University of Arizona. 2000:252.
Esperanca S, Carlson RW, Shirey SB, Smith D. Dating crust-mantle separation: Re-Os isotopic study of mafic xenoliths from central Arizona. Geology.
1997;25:651-4.
Evans BW, Frost BR. Chrome-spinel in progressive metamorphism--a preliminary analysis. Geochim. Cosmochim. Acta. 1975;39:959-72.
Evans KA, McCuaig TC, Leach D, Angerer T, Hagemann SG. Banded iron formation to iron ore: A record of the evolution of Earth environments?
Geology. 2013;41:99-102.
Ewart A, Bryan WB, Gill JB. Mineralogy and geochemistry of the younger volcanic islands of Tonga, S.W. Pacific. J. Petrol. 1973;14:429-65.
Ewart A, Griffin WL. Application of proton-microprobe data to trace-element partitioning in volcanic rocks. Chem. Geol. 1994;117:251-84.
Figueiredo e Silva RC, Lobato LM, Hagemann S, Danyushevsky L. Laser-ablation ICP-MS analyses on oxides of hypogene iron ore from the giant Serra
Norte jaspilite-hosted iron ore deposits, Carajs Mineral Province, Brazil. In: Williams PJ, editor. Proceedings of the 10th Biennial SGA Meeting of
The Society for Geology Applied to Mineral Deposits Townsville Australia 17th - 20th August 2009. Townsville2009. p. 570-2.
Figueiredo e Silva RC, Lobato LM, Rosiere CA. A hydrothermal origin for the jaspilite-hosted giant Sierra Norte deposits in the Cajajas Mineral
Province, Para State, Brazil. In: Hagemann S, Rosire CA, Gutzmer J, Beukes NJ, editors. Banded Iron Formation-related High-grade Iron Ore: Society
of Economic Geologists; 2008. p. 255-90.
Fleet M. The structure of magnetite. Acta Crystallographica Section B. 1981;37:917-20.
Friedman I, O'Neil JR. Compilation of stable isotope fractionation factors of geochemical interest. U.S. Geological Survey Professional Paper 440-KK.
1977:12.
Frost BR. Introduction to oxygen fugacity and its petrologic importance. In: Lindsley Donald H, editor. Oxide minerals: Petrologic and magnetic
significance: Mineralogical Society of America; 1991a. p. 1-9.
Frost BR. Magnetic petrology: Factors that control the occurence of magnetite in crustal rocks. In: Lindsley DH, editor. Oxide Minerals: Petrologic
and Magnetic Significance: Reviews in Mineralogy, Mineralogical Society of America; 1991b. p. 489-509.

36

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

SC

RI
P

Frost BR. Stability of oxide minerals in metamorphic rocks. In: Lindsley DH, editor. Oxide Minerals: Petrologic and Magnetic Significance: Reviews in
Minerology, Mineralogical Society of America; 1991c. p. 469-87.
Frost BR, Lindsley DH. Occurrence of iron-titanium oxides in igneous rocks In: Lindsley DH, editor. Oxide Minerals: Petrologic and Magnetic
Significance: Reviews in Minerology, Mineralogical Society of America; 1991. p. 489-509.
Frost BR, Lindsley Donald H, Andersen DJ. Fe-Ti oxide-silicate equilibria: Assemblages with fayalitic olivine. Am. Mineral. 1988;73:727-40.
Fryklund VC. Ore deposits of the Coeur d'Alene district, Shoshone County, Idaho. U.S. Geological Survey Professional Paper 445. 1964:103.
Gaetani GA, Grove TL. Partitioning of moderately siderophile elements among olivine, silicate melt, and sulfide melt: Constraints on core formation
in the Earth and Mars. Geochim. Cosmochim. Acta. 1997;61:1,829-1,46.
Gammons CH, Williams-Jones AE. Chemical mobility of gold in the porphyry-epithermal environment. Econ. Geol. 1997;92:45-59.
Garrity CP, and Soller, D.R., 2009. Database of the Geologic Map of North America; adapted from the map by J.C. Reed, Jr. and others (2005), U.S.
Geological Survey Data Series 424. 2009.
Ghent E. Temperature, pressure, and mixed-volatile equilibria attending metamorphism of staurolite-kyanitebearing assemblages, Esplanade
Range, British Columbia. Geol. Soc. Am. Bull. 1975;86:1654-60.
Ghiorso MS, Sack O. Fe-Ti oxide geothermometry: thermodynamic formulation and the estimation of intensive variables in silicic magmas. Contrib.
Mineral. Petrol. 1991a;108:485-510.
Ghiorso MS, Sack O. Thermochemistry of the oxide minerals. In: Lindsley DH, editor. Oxide Minerals: Petrologic and Magnetic Significance:
Mineralogical Society of America; 1991b. p. 221-64.
Goldschmidt VM. Geochemistry. Oxford: Oxford University Press; 1954.
Grant FS. Aeromagnetics, geology and ore environments, I. Magnetite in igneous, sedimentary and metamorphic rocks: An overview.
Geoexploration. 1984/85a;23:303-33.
Grant FS. Aeromagnetics, geology and ore environments, II. Magnetite and ore environments. Geoexploration. 1984/85b;23:335-62.
Grigsby JD. Detrital magnetite as a provenance indicator. J. Sediment. Res. 1990;60:940-51.
Gruner JW. Hydrothermal leaching of iron ores of the Lake Superior type - a modified theory. Econ. Geol. 1937;32:121-30.
Gustafson LB, Hunt JP. The porphyry copper deposit at El Salvador, Chile. Econ. Geol. 1975;70:857-912.
Gutzmer J, Beukes NJ, de Kock MO, Netshiozwi ST. Origin of high-grade iron ores at the Thabazimbi Deposit, South Africa. In: AusIMM, editor.
Proceedings Iron Ore Conference 2005, 19 - 21 September 2005, Freemantle, Western Australia2005. p. 57-65.
Haggerty SE. Oxide mineralogy of the upper mantle. In: Lindsley Donald H, editor. Oxide minerals: Petrologic and magnetic significance:
Mineralogical Society of America; 1991a. p. 355-416.
Haggerty SE. Oxide Textures - A Mini-Atlas. In: Lindsley Donald H, editor. Oxide Minerals: Petrologic and Magnetic Significance: Reviews in
Mineralogy, Mineralogical Society of America; 1991b. p. 129-219.
Hall AJ. Pyrite-pyrrhotine redox reactions in nature. Mineral. Mag. 1986;50:223-9.
Halsall TJ. Advances in geology of the porphyry copper deposits: Southwestern North America : Edited by Spencer R. Titley. Pp. 560. University of
Arizona Press, Tucson.1983.
Haskin LA, Frey FA, Schmitt RA, Smith RH. Meteoritic, solar and terrestrial rare-earth distributions. Physical and Chemical Earth. 1966;7:167-321.
Hayes TS. A preliminary study of thermometry and metal sources of the Spar Lake stratabound copper-silver deposit, Belt Supergroup, Montana.
U.S. Geological Survey Open-File Report 90-484. 1990:30.
Hedenquist JW, Lowenstern JB. The role of magmas in the formation of hydrothermal ore deposits. Nature. 1994;370:519-27.
Heiligmann M, Williams-Jones AE, Clark JR. The Role of Sulfate-Sulfide-Oxide-Silicate Equilibria in the Metamorphism of Hydrothermal Alteration at
the Hemlo Gold Deposit, Ontario. Econ. Geol. 2008;103:335-51.
Horn I, Foley SF, Jackson SE, Jenner GA. Experimentally determined partitioning of high field strength- and selected transition elements between
spinel and basaltic melt. Chem. Geol. 1994;117:193-218.
Huang X-W, Zhou M-F, Qi L, Gao J-F, Wang Y-W. ReOs isotopic ages of pyrite and chemical composition of magnetite from the Cihai magmatic
hydrothermal Fe deposit, NW China. Mineral. Deposita. 2013:1-22.
Huberty JM, Konishi H, Heck PR, Fournelle JH, Valley JW, Xu H. Silician magnetite from the Dales Gorge Member of the Brockman Iron Formation,
Hamersley Group, Western Australia. Am. Mineral. 2012;97:26-37.
Hutton CO. Studies of heavy detrital minerals. Geol. Soc. Am. Bull. 1950;61:635-710.
Ilton ES, Eugster HP. Base metal exchange between magnetite and a chloride-rich hydrothermal fluid. Geochim. Cosmochim. Acta. 1989;53:291-301.
Ishihara S. Magnetite-series and ilmenite-series granitic rocks. Mining Geology. 1977;27:293-305.
Jenner FE, ONeill HSC, Arculus RJ, Mavrogenes JA. The Magnetite Crisis in the Evolution of Arc-related Magmas and the Initial Concentration of Au,
Ag and Cu. J. Petrol. 2010;51:2445-64.
Kamvong T, Zaw K, Siegele R. PIXE/PIGE microanalysis of trace elements in hydrothermal magnetite and exploration significance: a pilot study.
15th Australian Conference on Nuclear and Complementary Techniques of Analysis and 9th Vacuum Society of Australia Congress. Melbourne,
Australia: University of Melbourne; 2007.
Kesler SE. Ore-Forming Fluids. Elements. 2005;1:13-8.
Klein C. Some Precambrian banded iron-formations (BIFs) from around the world: Their age, geologic setting, mineralogy, metamorphism,
geochemistry, and origin. Am. Mineral. 2005;90:1473-99.
Klemm DD, Henckel J, Dehm RM, Von Gruenewaldt G. The geochemistry of titanomagnetite in magnetite layers and their host rocks of the eastern
Bushveld Complex. Econ. Geol. 1985;80:1075-88.
Klemme S, Gnther D, Hametner K, Prowatke S, Zack T. The partitioning of trace elements between ilmenite, ulvospinel, armalcolite and silicate
melts with implications for the early differentiation of the moon. Chem. Geol. 2006;234:251-63.
Kretz R, Campbell JL, Hoffman EL, Hartree R, Teesdale WJ. Approaches to equilibrium in the distribution of trace elements among the principal
minerals in a high-grade metamorphic terrane. J. Metamorph. Geol. 1999;17:41-59.
Landis GP, Hofstra AH. Fluid inclusion gas chemistry as a potential minerals exploration tool: Case studies from Creede, CO, Jerritt Canyon, NV,
Coeur d'Alene district, ID and MT, southern Alaska mesothermal veins, and mid-continent MVT's. J. Geochem. Explor. 1991;42:25-59.
Larocque ACL, Stimac JA, McMahon G, Jackman JA, Chartrand VP, Hickmott D, et al. Ion-Microprobe analysis of FeTi Oxides: Optimization for the
determination of invisible gold. Econ. Geol. 2002;97:159-64.
LaTourrette TZ, Burnett DS, Bacon CR. Uranium and minor-element partitioning in Fe-Ti oxides and zircon from partially melted granodiorite, Crater
Lake, Oregon. Geochim. Cosmochim. Acta. 1991;55:457-69.
Leach DL, Landis GP, Hofstra AH. Metamorphic origin of the Coeur d'Alene base- and precious-metal veins in the Belt Basin, Idaho and Montana.
Geology. 1988;16:122-5.
Lee MJ, Lee JI, Moutte J. Compositional variation of Fe-Ti oxides from the Sokli complex, north-eastern Finland. Geosci. J. 2005;9:1-13.
Leeman WP. Experimental determination of partitioning of divalent cations between olivine and basaltic liquid.: University of Oregon; 1974.

37

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

SC

RI
P

Leeman WP, Ma MS, Murali AV, Schmitt RA. Empirical estimation of magnetite/liquid distribution coefficients for some transition elements. Contrib.
Mineral. Petrol. 1978;65:269-72.
Lemarchand F, Villemant B, Calas G. Trace element distribution coefficients in alkaline series. Geochim. Cosmochim. Acta. 1987;51:1071-81.
Lindsley DH. The crystal chemistry and structure of oxide minerals as exemplified by the Fe-Ti oxides. In: Rumble III D, editor. Oxide Minerals:
Reviews in Mineralogy, Mineralogical Society of America; 1976a. p. L1-L60.
Lindsley DH. Experimental studies of oxide minerals. In: Rumble III D, editor. Oxide Minerals: Reviews in Mineralogy, Mineralogical Society of
America; 1976b. p. L61-L88.
Lindsley DH. Oxide Minerals: Petrologic and Magnetic Significance: Reviews in Mineralogy, Mineralogical Society of America; 1991.
Lindstrom DJ. Experimental study of the partitioning of the transition metals between clinopyroxene and coexisting silicate liquids: University of
Oregon; 1976.
Lobato LM, Figueiredo e Silva RC, Hagemann SG, Thorne WS, Zuchetti M. Hypogene alteration associated with high-grade banded iron formationrelated iron ore. In: Hagemann SG, Rosire CA, Gutzmer J, Beukes NJ, editors. Banded iron formation-related high-grade iron ore. Reviews in
Economic Geology2008. p. 107-28.
Longerich HP, Jackson SE, Gnther D. Laser ablation-inductively coupled plasma-mass spectrometric transient signal data acquisition and analyte
concentration calculation. J. Anal. At. Spectrom. 1996;11:899-904.
Lowell JD, Guilbert JM. Lateral and vertical alteration-mineralization zoning in porphyry ore deposits. Econ. Geol. 1970;65:373-408.
Luhr JF, Carmichael ISE. The Colima Volcanic complex, Mexico. Contrib. Mineral. Petrol. 1980;71:343-72.
Luhr JF, Carmichael ISE, Varekamp JC. The 1982 eruptions of El Chichon volcano, Chiapas, Mexico: mineralogy and petrology of the anhydritebearing pumices. J. Volcanol. Geotherm. Res. 1984;23:69-108.
MacDonald GD, Arnold LC. Geological and geochemical zoning of the Grasberg Igneous Complex, Irian Jaya, Indonesia. J. Geochem. Explor.
1994;50:143-78.
Mahood G, Hildreth W. Large partition coefficients for trace elements in high-silica rhyolites. Geochim. Cosmochim. Acta. 1983;47:11-30.
Mahood GA, Stimac JA. Trace-element partitioning in pantellerites and trachytes. Geochim. Cosmochim. Acta. 1990;54:2257-76.
Mallmann G, O'Neill HSC. The crystal/melt partitioning of V during mantle melting as a function of oxygen fugacity compared with some other
elements (Al, P, Ca, Sc, Ti, Cr, Fe, Ga, Y, Zr and Nb). J. Petrol. 2009;50:1765-94.
Malmqvist D, Parasnis DS. Aitik: geophysical documentation of a third-generation copper deposit in North Sweden. Geoexploration. 1972;10:14960.
Manske SL, Paul AH. Geology of a major new porphyry copper center in the Superior (Pioneer) district, Arizona. Econ. Geol. 2002;97:197-220.
Mathur R, Ruiz J, Titley S, Gibbins S, Margotomo W. Different crustal sources for Au-rich and Au-poor ores of the Grasberg Cu-Au porphyry deposit.
Earth Planet. Sci. Lett. 2000;183:7-14.
McClenaghan MB. Indicator mineral methods in mineral exploration. Geochem. Explor. Environ. Anal. 2005;5:233-45.
McEnroe SA, Robinson P, Panish PT. Aeromagnetic anomalies, magnetic petrology, and rock magnetism of hemo-ilmenite- and magnetite-rich
cumulate rocks from the Sokndal Region, South Rogaland, Norway. Am. Mineral. 2001;86:1447-68.
McIntire WL. Trace element partition coefficients--a review of theory and applications to geology. Geochim. Cosmochim. Acta. 1963;27:1209-64.
McLemore VT, Munroe EA, Heizler MT, McKee C. Geochemistry of the Copper Flat porphyry and associated deposits in the Hillsboro mining district,
Sierra County, New Mexico, USA. J. Geochem. Explor. 1999;67:167-89.
McQueen KG, Cross AJ. Magnetite as a geochemical sampling medium: application to skarn deposits. In: Eggleton RA, editor. The State of the
Regolith. Brisbane: Geological Society of Australia; 1998. p. 194-9.
McQueen KG, Whitbread MA. Mineralogical exploration: using element-host mineral associations in the search for ore. In: Roach IC, editor. Regolith
and Landscapes in Eastern Australia: CRC LEME; 2002. p. 96-9.
Meinert LD. Skarn zonation and fluid evolution in the Groundhog Mine, Central mining district, New Mexico. Econ. Geol. 1987;82:523-45.
Meinert LD, Dipple GM, Nicolescu S. World skarn deposits. Economic Geology 100th Anniversary Volume. 2005:299-336.
Meinert LD, Hefton KK, Mayes D, Tasiran I. Geology, zonation, and fluid evolution of the Big Gossan Cu-Au skarn deposit, Ertsberg District, Irian Jaya.
Econ. Geol. 1997;92:509-34.
Mertig HJ, Rubin JN, Kyle JR. Skarn Cu Au orebodies of the Gunung Bijih (Ertsberg) district, Irian Jaya, Indonesia. J. Geochem. Explor.
1994;50:179-202.
Mollo S, Putirka K, Iezzi G, Scarlato P. The control of cooling rate on titanomagnetite composition: implications for a geospeedometry model
applicable to alkaline rocks from Mt. Etna volcano. Contrib. Mineral. Petrol. 2013;165:457-75.
Morris RC. Genesis of iron ore in banded iron-formation by supergene and supergene-metamorphic processes - a conceptual model. In: Wolf KH,
editor. Handbook of strata-bound and stratiform ore deposits. Amsterdam: Elsevier; 1985. p. 73-235.
Mcke A, Raphael Cabral A. Redox and nonredox reactions of magnetite and hematite in rocks. Chem. Erde Geochem. 2005;65:271-8.
Mukhopadhyay J, Gutzmer J, Beukes NJ, Hayashi KI. Stratabound magnetite deposits from the eastern outcrop belt of Archaean Iron Ore Group,
Singhbhum craton, India. Applied Earth Science (Trans. Inst. Min. Metall. B). 2008;117:175-86.
Mller B, Axelsson MD, Ohlander B. Trace elements in magnetite from Kiruna, northern Sweden, as determined by LA-ICP-MS. Gff. 2003;125:1-5.
Muntean JL, Einaudi MT. Porphyry-Epithermal Transition: Maricunga Belt, Northern Chile. Econ. Geol. 2001;96:743-72.
Nadoll P, Koenig AE. LA-ICP-MS of magnetite: methods and reference materials. J. Anal. At. Spectrom. 2011;26:1872-7.
Nadoll P, Mauk JL. Wstite in a hydrothermal silver-lead-zinc vein, Lucky Friday mine, Coeur d'Alene mining district, U.S.A. Am. Mineral.
2011;96:261-7.
Nadoll P, Mauk JL, Hayes TS, Koenig AE, Box SE. Geochemistry of Magnetite from Hydrothermal Ore Deposits and Host Rocks of the
Mesoproterozoic Belt Supergroup, United States. Econ. Geol. 2012;107:1275-92.
Nadoll P, Mauk JL, Leveille R, Koenig AE. Geochemistry of magnetite from Cu-Mo porphyry + skarn, and Climax Mo deposits in the western United
States. In: Pl-Molnr E, editor. IMA 2010. Budapest: Department of Mineralogy, Geochemistry and Petrology, University of Szeged; 2010. p. 507.
Nadoll P, Mauk JL, LeVeille R, Koenig AE. Geochemistry of magnetite from porphyry Cu and skarn deposits in the southwestern United States.
submitted.
Nash WP, Crecraft HR. Partition coefficients for trace elements in silicic magmas. Geochim. Cosmochim. Acta. 1985;49:2309-22.
Nielsen RL. BIGD.FOR: A FORTRAN program to calculate trace-element partition coefficients for natural mafic and intermediate composition
magmas. Comput. Geosci. 1992;18:773-88.
Nielsen RL, Beard JS. Magnetite-melt HFSE partitioning. Chem. Geol. 2000;164:21-34.
Nielsen RL, Forsythe LM, Gallahan WE, Fisk MR. Major- and trace-element magnetite-melt equilibria. Chem. Geol. 1994;117:167-91.
Ohmoto H. Nonredox transformations of magnetite-hematite in hydrothermal systems. Econ. Geol. 2003;98:15761.
Okamoto K. Geochemical study on magmatic differentiation of Asama Volcano, central Japan. J. Geol. Soc. Jpn. 1979;85:525-35.
Otake T, Wesolowski DJ, Anovitz LM, Allard LF, Ohmoto H. Mechanisms of iron oxide transformations in hydrothermal systems. Geochim.
Cosmochim. Acta. 2010;74:6141-56.

38

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

SC

RI
P

Pecoits E, Gingras MK, Barley ME, Kappler A, Posth NR, Konhauser KO. Petrography and geochemistry of the Dales Gorge banded iron formation:
Paragenetic sequence, source and implications for palaeo-ocean chemistry. Precambrian Res. 2009;172:163-87.
Pollard P, Taylor R. Paragenesis of the Grasberg CuAu deposit, Irian Jaya, Indonesia: results from logging section 13. Mineral. Deposita.
2002;37:117-36.
Pollard PJ, Taylor RG, Peters L. Ages of intrusion, alteration, and mineralization at the Grasberg Cu-Au deposit, Papua, Indonesia. Econ. Geol.
2005;100:1005-20.
Powell R, Powell M. Geothermometry and oxygen barometry using coexisting iron-titanium oxides: a reappraisal. Mineral. Mag. 1977;41:257-63.
Prins P. Composition of magnetite from carbonatites. Lithos. 1972;5:227-40.
Ramanaidou ER. Genesis of lateritic iron ore from banded iron-formation in the Capanema mine (Minas Gerais, Brazil). Aust. J. Earth Sci.
2009;56:605-20.
Ramdohr P. Beobachtungen an Magnetit, Ilmenit, Eisenglanz, und Uberlegungen iiber das System FeO-Fe2O3-TiO2. Neues Jahrbuch fuer
Mineralogie. 1926;54A:320-79.
Ramdohr P. Die Erzminerale und ihre Verwachsungen. Berlin: Akademie-Verlag; 1955.
Ray GE, Webster ICL. Geology and chemistry of the low Ti magnetite-bearing Heff Cu-Au skarn and its associated plutonic rocks, Heffley Lake, SouthCentral British Columbia. Explor. Min. Geol. 2007;16:159-86.
Razjigaeva NG, Naumova VV. Trace element composition of detrital magnetite from coastal sediments of northwestern Japan Sea for provenance
study. J. Sediment. Petrol. 1992;62:802-9.
Reguir EP, Chakhmouradian AR, Halden NM, Yang P. Early magmatic and reaction-induced trends in magnetite from the carbonatites of Kerimasi,
Tanzania. The Canadian Mineralogist. 2008;46:879-900.
Reid F. Origin of the Rhyolitic Rocks of the Taupo Volcanic Zone, New-Zealand. J. Volcanol. Geotherm. Res. 1983;15:315-38.
Righter K, Leeman WP, Hervig RL. Partitioning of Ni, Co and V between spinel-structured oxides and silicate melts: Importance of spinel
composition. Chem. Geol. 2006a;227:1-25.
Righter K, Sutton SR, Newville M, Le L, Schwandt CS, Uchida H, et al. An experimental study of the oxidation state of vanadium in spinel and basaltic
melt with implications for the origin of planetary basalt. Am. Mineral. 2006b;91:1643-56.
Robinson RF, Cook A. The Safford copper deposit, Lone Star mining district, Graham County, Arizona. In: Titley SR, Hicks CL, editors. Geology of the
porphyry copper deposits, southwestern North America. Tucson: University of Arizona Press; 1966. p. 251-66.
Roedder E, Bodnar R. Fluid inclusion studies of hydrothermal ore deposits. In: Barnes H, editor. Geochemistry of hydrothermal ore deposits. New
York: John Wiley; 1997. p. 657-98.
Rose AW, Baltosser WW. The porphyry copper deposit at Santa Rita, New Mexico. In: Titley SR, Hicks CL, editors. Geology of the porphyry copper
deposits - Southwestern North America. Tucson: The University of Arizona Press; 1966. p. 205-20.
Rosire CA, Rios FJ. The origin of hematite in high-grade iron ores based on infrared microscopy and fluid inclusion studies: the example of the
Conceio mine, Quadriltero Ferrfero, Brazil. Econ. Geol. 2004;99:611-24.
Rosire CA, Spier CA, Rios FJ, Suckau VE. The itabirites of the Quadriltero Ferrfero and related high-grade iron ore deposits: An overview. In:
Hagemann SG, Rosire CA, Gutzmer J, Beukes NJ, editors. Banded Iron Formation-related High-grade Iron Ore2008. p. 223-54.
Rumble D. Oxide Minerals. Rev. Mineral.: Mineralogical Society of America, Short Course Notes; 1976a.
Rumble D. Oxide Minerals in Metamorphic Rocks. In: Rumble III D, editor. Oxide Minerals: Reviews in Mineralogy, Mineralogical Society of America;
1976b. p. R1-R20.
Rusk B, Oliver N, Cleverley J, Blenkinsop T, Zhang D, Williams P, et al. Physical and chemical characteristics of the Ernest Henry iron oxide copper
gold deposit, Australia; implications for IOGC genesis In: Porter TM, editor. Hydrothermal Iron Oxide Copper-Gold and Related Deposits: A Global
Perspective. Adelaide: PGC Publishing; 2010. p. 201-18.
Rusk BG, Oliver N, Brown A, Lilly R, Jungmann D. Barren magnetite breccias in the Cloncurry region, Australia; comparisons to IOCG deposits. In:
Williams PJ, editor. Proceedings of the 10th Biennial SGA Meeting of The Society for Geology Applied to Mineral Deposits. Townsville Australia: The
Society for Geology Applied to Mineral Deposits; 2009. p. 656-8.
Rusk BG, Reed MH, Dilles JH. Fluid inclusion evidence for magmatic-hydrothermal fluid evolution in the porphyry copper-molybdenum deposit at
Butte, Montana. Econ. Geol. 2008;103:307-34.
Rusk BG, Reed MH, Dilles JH, Klemm LM, Heinrich CA. Compositions of magmatic hydrothermal fluids determined by LA-ICP-MS of fluid inclusions
from the porphyry copper-molybdenum deposit at Butte, MT. Chem. Geol. 2004;210:173-99.
Ryabchikov ID, Kogarko LN. Magnetite compositions and oxygen fugacities of the Khibina magmatic system. Lithos. 2006;91:35-45.
Sack RO, Lichtner PC. Constraining compositions of hydrothermal fluids in equilibrium with polymetallic ore-forming sulfide assemblages. Econ.
Geol. 2009;104:1249-64.
Sauerzapf U, Lattard D, Burchard M, Engelmann R. The Titanomagnetite-Ilmenite Equilibrium: New Experimental Data and Thermo-oxybarometric
Application to the Crystallization of Basic to Intermediate Rocks. J. Petrology. 2008;49:1161-85.
Savard D, Barnes S-J, Sunder Raju PV. Accurate LA-ICP-MS calibration for magnetite analysis using multiple reference materials. Goldschmidt
Conference Abstracts. 2010:A914.
Sawkins FJ, Scherkenbach DA. High copper content of fluid inclusions in quartz from northern Sonora: Implications for ore genesis theory. Geology.
1981;9:37-40.
Schilling J, Frost BR, Marks MAW, Wenzel T, Markl G. Fe-Ti oxide-silicate (QUIlF-type) equilibria in feldspathoid-bearing systems. Am. Mineral.
2011;96:100-10.
Seedorff E, Dilles JH, Proffett Jr. JM, Einaudi MT, Zurcher L, Stavast WJA, et al. Porphyry Deposits: Characteristics and origin of hypogene features.
Econ. Geol. 2005;100th Anniversary Volume:25198.
Seedorff E, Einaudi MT. Henderson porphyry molybdenum system, Colorado: I. Sequence and abundance of hydrothermal mineral assemblages,
flow paths of evolving fluids, and evolutionary style. Econ. Geol. 2004;99:3-37.
Shannon JR, Nelson EP, Golden Jr. RJ. Surface and underground geology of the world-class Henderson molybdenum porphyry mine, Colorado.
Geological Society of America, Field Guide. 2004;5.
Shannon R. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallographica Section
A. 1976;32:751-67.
Shcheka SA, Naumova VV, Wrzosek AA. Trace-element paragenesis in hydrothermal magnetite as indicators of the origin and ore potential of
mineralization. Dokl. Akad. Nauk SSSR. 1988;294:958-62.
Shcheka SA, Platkov AV, Vrzhosek AA, Levashov GB, Oktyabrsky RA. The Trace element paragenesis of magnetite. Nauka. 1980:147.
Shcheka SA, Zabelin VV, Chubarov VM. Magnetites and ferric hydroxides in sediments of the Japan and Philippine seas and their genetic
information. Mar. Geol. 1982;45:M23-M9.
Sidhu PS, Gilkes RJ, Posner AM. The synthesis and some properties of Co, Ni, Zn, Cu, Mn and Cd substituted magnetites. Journal of Inorganic and
Nuclear Chemistry. 1978;40:429-35.

39

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

SC

RI
P

Sidhu PS, Gilkes RJ, Posner AM. Oxidation and Ejection of Nickel and Zinc from Natural and Synthetic Magnetites1. Soil Sci. Soc. Am. J. 1981;45:6414.
Sillitoe RH. Gold-rich porphyry deposits: Descriptive and genetic models and their role in exploration and discovery. Rev. Econ. Geol. 2000;13:31545.
Sillitoe RH. Iron oxide-copper-gold deposits: an Andean view. Mineral. Deposita. 2003;38:787-812.
Sillitoe RH. Porphyry Copper Systems. Econ. Geol. 2010;105:3-41.
Simon AC, Candela PA, Piccoli PM, Mengason M, Englander L. The effect of crystal-melt partitioning on the budgets of Cu, Au, and Ag. Am. Mineral.
2008;93:1437-48.
Singoyi B, Danyushevsky L, Davidson GJ, Large R, Zaw K. Determination of trace elements in magnetites from hydrothermal deposits using the LAICP-MS technique. Abstracts of oral and poster presentations from the SEG 2006 conference. Keystone, USA: Society of Economic Geologists; 2006.
p. 367-8.
Singoyi B, Zaw K. A petrological and fluid inclusion study of magnetite-scheelite skarn mineralization at Kara, Northwestern Tasmania: implications
for ore genesis. Chem. Geol. 2001;173:239-53.
Skublov S, Drugova G. Patterns of trace-element distribution in calcic amphiboles as afunction of metamorphic grade. Can. Mineral. 2003;41:383-92.
Spong PL. Geochemistry of magnetite from convergent-margin plutonic rocks of Australia, Japan and New Zealand [MSc]: University of Auckland;
1998.
Stimac J, Hickmott D. Trace-Element Partition-Coefficients for Ilmenite, Ortho-Pyroxene and Pyrrhotite in Rhyolite Determined by Micro-Pixe
Analysis. Chem. Geol. 1994;117:313-30.
Streck MJ, Grunder AL. Compositional gradients and gaps in high-silica rhyolites of the Rattlesnake Tuff, Oregon. J. Petrol. 1997;38:133-63.
Tacker RC, Candela PA. Partitioning of molybdenum between magnetite and melt; a preliminary experimental study of partitioning of ore metals
between silicic magmas and crystalline phases. Econ. Geol. 1987;82:1827-38.
Tauson VL, Babkin DN, Pastushkova TM, Akimov VV, Krasnoshchekova TS, Lipko SV, et al. Dualistic distribution coefficients of elements in the system
mineral-hydrothermal solution. II. Gold in magnetite. Geochem. Int. 2012;50:227-45.
Tauson VL, Babkin DN, Pastushkova TM, Krasnoshchekova TS, Lustenberg EE, Belozerova OY. Dualistic distribution coefficients of elements in the
system mineral-hydrothermal solution. I. Gold accumulation in pyrite. Geochem. Int. 2011;49:568-77.
Taylor D, Dalstra HJ, Harding AE, Broadbent GC, Barley ME. Genesis of high-grade hematite orebodies of the Hamersley province, Western Australia.
Econ. Geol. Bull. Soc. Econ. Geol. 2001;96:837-73.
Thorne W, Hagemann S, Barley M. Petrographic and geochemical evidence for hydrothermal evolution of the North Deposit, Mt Tom Price, Western
Australia. Mineral. Deposita. 2004;39:766-83.
Thorne WS, Hagemann SG, Webb A, Clout J. Banded iron formation-related iron ore deposits of the Hamersley Province, Western Australia. In:
Hagemann SG, Rosire CA, Gutzmer J, Beukes NJ, editors. Banded iron formation-related high-grade iron ore2008. p. 197-222.
Titley SR. Advances in geology of the porphyry copper deposits of southwestern North America. Tucson: The University of Arizona Press; 1981.
Titley SR, Beane RE. Porphyry copper deposits, part I. Geologic settings, petrology, and tectogenesis. Econ. Geol. 1981;75th Anniversary:214-35.
Togashi S, Terashima S. The behavior of gold in unaltered island arc tholeiitic rocks from Izu-Oshima, Fuji, and Osoremaya Volcanic Areas, Japan.
Geochim. Cosmochim. Acta. 1997;61:543-54.
Toplis M, Corgne A. An experimental study of element partitioning between magnetite, clinopyroxene and iron-bearing silicate liquids with
particular emphasis on vanadium. Contrib. Mineral. Petrol. 2002;144:22-37.
Toplis MJ, Carroll MR. An experimental study of the influence of oxygen fugacity on Fe-Ti oxide stability, phase relations, and mineral-melt equilibria
in ferro-basaltic systems. J. Petrol. 1995;36:1137-70.
Trendall AF, Blockley JG. The iron formations of the Precambrian Hamersley Group, Western Australia; with special reference to the crocidolite.
Geological Survey of Western Australia Bulletin 119; 1970. p. 366p.
Triebold S, von Eynatten H, Luvizotto GL, Zack T. Deducing source rock lithology from detrital rutile geochemistry: An example from the Erzgebirge,
Germany. Chem. Geol. 2007;244:421-36.
Ulrich T, Gunther D, Heinrich CA. Gold concentrations of magmatic brines and the metal budget of porphyry copper deposits. Nature. 1999;399:676.
Ulrich T, Gunther D, Heinrich CA. The evolution of a porphyry Cu-Au deposit, based on LA-ICP-MS analysis of fluid inclusions: Bajo de la Alumbrera,
Argentina. Econ. Geol. 2001;96:1743-74.
Van Baalen MR. Titanium mobility in metamorphic systems: a review. Chem. Geol. 1993;110:233-49.
Verlaguet A, Brunet F, Goff B, Murphy WM. Experimental study and modeling of fluid reaction paths in the quartz-kyanite muscovite-water
system at 0.7 GPa in the 350-550C range: Implications for Al selective transfer during metamorphism. Geochim. Cosmochim. Acta. 2006;70:177288.
Villemant B, Jaffrezic H, Joron J-L, Treuil M. Distribution coefficients of major and trace elements; fractional crystallization in the alkali basalt series
of Chane des Puys (Massif Central, France). Geochim. Cosmochim. Acta. 1981;45:1997-2016.
Vincent EA, Phillips R. Iron-titanium oxide minerals in layered gabbros of the Skaergaard intrusion, East Greenland: Part I. Chemistry and oremicroscopy. Geochim. Cosmochim. Acta. 1954;6:1-4, IN1-IN2, 5-26.
Wang R, Yu AP, Chen J, Xie L, Lu J-J, Zhu J-C. Cassiterite exsolution with ilmenite lamellae in magnetite from the Huashan metaluminous tin granite
in southern China. Mineral. Petrol. 2012;105:71-84.
Waychunas GA. Crystal chemistry of oxides and oxyhydroxides. In: Lindsley DH, editor. Oxide Minerals: Petrologic and magnetic significance:
Mineralogical Society of America; 1991. p. 11-61.
Wechsler BA, Lindsley DH, Prewitt CT. Crystal structure and cation distribution in titanomagnetites (Fe (3-x) Ti (x) O (4) ). Am. Mineral. 1984;69:75470.
Whalen JB, Chappel BW. Opaque mineralogy and mafic mineral chemistry of I- and S-type granites of the Lachlan fold belt, southeast Australia. Am.
Mineral. 1988;73:281-96.
White AF, Peterson ML, Hochella MF. Electrochemistry and dissolution kinetics of magnetite and ilmenite. Geochim. Cosmochim. Acta.
1994;58:1859-75.
Whitney DL, Evans BW. Abbreviations for names of rock-forming minerals. Am. Mineral. 2010;95:185-7.
Williams PJ, Barton MD, Johnson DA, Fontbot L. Iron oxide copper-gold deposits: geology, space-time distribution, and possible modes of origin.
Econ. Geol. Bull. Soc. Econ. Geol. 2005.
Wones DR. Significance of the assemblage titanite + magnetite + quartz in granitic rocks. Am. Mineral. 1989;74:744-9.
Yardley BWD. Metal concentrations in crustal fluids and their relationship to ore formation, 100th anniversary special paper. Econ. Geol.
2005;100:613-32.
Zack T, von Eynatten H, Kronz A. Rutile geochemistry and its potential use in quantitative provenance studies. Sediment. Geol. 2004;171:37-58.
Zhang D, Rusk B, Oliver N. Trace elements in sulfides and magnetite from the Ernest Henry Iron Oxide-Copper-Gold deposit, Australia. Geological
Society of America Annual Meeting, Portland, Oregon, 18-21 October 2009. 2009;41:85.

40

ACCEPTED MANUSCRIPT
Zheng Y-F, Simon K. Oxygen isotope fractionation in hematite and magnetite; a theoretical calculation and application to geothermometry of
metamorphic iron-formations. Eur. J. Mineral. 1991;3:877-86.

9 List of Tables

Geological Survey Denver (USGS), and CODES University of Tasmania.

RI
P

Table 1 Laser ablation ICP-MS operation conditions for the Universit du Qubec Chicoutimi (UQAC), United Sates

SC

Table 2 LA-ICP-MS reporting limits (10x standard deviation of the blank) and detection limits (3x standard deviation of
the blank) for the three LA-ICP-MS laboratories and the range of measured elements and different laser ablation spot

NU

sizes and EPMA detection limits.

and number (#) of analysis (LA-ICP-MS or EMPA).

MA

Table 3 List of samples including information about domain, magnetite type, host rock, alteration, location, and type

ED

Table 4 Summary statistics for selected hydrothermal and igneous magnetite occurrences based on EMPA and LA-ICPMS analyses. The corresponding graph is shown in Fig. 7.

PT

Table 5 Published partition coefficients for igneous magnetite from different rock/melt types for a selection of
elements. Values compiled from (EarthRef.org, 2013) and other sources - (Bacon and Druitt, 1988, D'Orazio et al.,

AC
CE

1998, Dudas et al., 1973, Esperanca et al., 1997, Ewart et al., 1973, Ewart and Griffin, 1994, Haskin et al., 1966, Horn et
al., 1994, Klemme et al., 2006, Kretz et al., 1999, LaTourrette et al., 1991, Leeman, 1974, Leeman et al., 1978,
Lemarchand et al., 1987, Lindstrom, 1976, Luhr and Carmichael, 1980, Luhr et al., 1984, Mahood and Hildreth, 1983,
Mahood and Stimac, 1990, Nash and Crecraft, 1985, Nielsen, 1992, Nielsen and Beard, 2000, Nielsen et al., 1994,
Okamoto, 1979, Reid, 1983, Righter et al., 2006a, Streck and Grunder, 1997, Tacker and Candela, 1987, Toplis and
Corgne, 2002, Villemant et al., 1981)

10 List of Figures
3+

Fig. 1 Cubic inverse spinel structure of magnetite and common cations that can substitute for Fe in the tetrahedral
2+

3+

sites and Fe /Fe in the octahedral sites. The illustrations show the cubic inverse spinel cell (A), and two diagrams
that show ionic radius in angstrom vs. cation charge for the tetrahedral site (B) and the octahedral site (C) data from
Shannon (1976). The incorporation of foreign cations is more likely when the substituting cation has a similar charge
and ionic radius (15-18% radius variations, after Goldschmidts rule - (Goldschmidt, 1954)). Coupled substitution is
41

ACCEPTED MANUSCRIPT
4+

required for the incorporation of 4+ cations such as Ti to maintain charge balance. Niobium, with a 5+ charge has
also been reported in magnetite (Nielsen et al., 1994). Other cations such as V, Cr, and Mn can occur in different
oxidation states and their incorporation is strongly dependent on the prevalent oxygen fugacity (Lindsley, 1991 and

references therein). For example, V becomes incompatible at high oxygen fugacity levels due to its 5+ oxidation state.

RI
P

High relative element abundance in the mineralizing fluid is a further controlling factor for substitution of largely
incompatible cations (e.g., Si, Al).

SC

Fig. 2 (A) Log fO2-T diagram showing relevant buffers for the Fe-Si-O system. HM: hematite-magnetite, FMQ: fayalitemagnetite-quartz, MW: magnetite-wstite, IW: iron-wstite, IM: iron-magnetite, and QIF: quartz-iron-fayalite

NU

(redrawn after Frost, 1991a, Frost et al., 1988). (B) Schematic phase diagram for the system Fe-O-S in fO2-fS2 space,
redrawn after Hall (1986) and Nadoll et al. (2011). The geometry of the diagram remains essentially the same even

MA

though fO2 and fS2 values vary significantly with temperature.

Fig. 3 The complexity of hydrothermal ore deposits translates into a variety of different types of hydrothermal

ED

magnetite. Different vein sets, types of alteration, fluid/rock interaction, and secondary processes may be reflected in
the composition of the respective type of hydrothermal magnetite. To gain a better understanding of the chemistry of

PT

magnetite in terms of process controlled compositional variability it is not only important to be able to discriminate

AC
CE

igneous and hydrothermal magnetite but also to address variations among different types of hydrothermal magnetite
from a vein-scale up to the deposit type-scale The figure shows a porphyry style deposit but a similar degree of
complexity can be found in other hydrothermal deposit types.

Fig. 4 Locations and schematic geology for (A) Mesoarchean Algoma-type BIF in the Yilgarn craton in Western
Australia, including the Koolyanobbing and Marda greenstone belts (Angerer et al., 2013, Angerer et al., 2012), and
Weld Range (B) Neoarchean to Paleoproterozoic Superior-type BIF in the Hamersley basin, Western Australia (Angerer
et al, unpublished data), (C) the Mesoarchean Iron Ore Group of JharkhandOrissa region, India (Bhattacharya et al.,
2007).

Fig. 5 Locations and schematic geology for (A) porphyry Cu (Mo) and associated skarn deposits in the southwestern
United States (Garrity, 2009, Nadoll et al., submitted), (B) the Henderson Climax-type Mo deposit in the central United
States (Garrity, 2009), (C) the Inner Zone Batholith (IZB) in the magnetite series igneous province in Japan (Ishihara,
1977), (D) the Ertsberg district, in the central highlands of New Guinea in the province of Papua, Indonesia, (E) the
Grasberg Igneous Complex located in the Ertsberg mining district.
42

ACCEPTED MANUSCRIPT
Fig. 6 (A) Typical BIF micro-banding from the Marra Mamba Iron Formation, Mesa Gap deposit, Hamersley Province,
Western Australia. (reflected light). (B) euhedral metamorphic magnetite partially replaced by hydrothermal hematite
(martite), Brockman Iron Formation, Hashimoto deposit, Hamersley Province. (C) Gold Hunter siderite vein with

abundant galena, sphalerite, calcite, and minor Fe-sulfides. Magnetite is a rare accessory mineral in this setting (Lucky

RI
P

Friday mine, Coeur dAlene mining district, Idaho, USA). (D) Massive replacement crystallization of magnetite, garnet,
and carbonate with crosscutting veins of Fe-(Cu)-sulfides in calcic skarn (Chino, Santa Rita deposit, New Mexico, USA).

SC

(E) Massive magnetite with abundant serpentinite in magnesian skarn from the Santa Rita porphyry Cu deposit, New
Mexico, USA. (F) Hydrothermal magnetite closely associated with chalcopyrite in a quartz-magnetite-chalcopyrite-

NU

epidote vein in andesitic host rock from the Safford-Dos Pobres porphyry deposit, Arizona, USA. (reflected light) (G)
Hydrothermal magnetite grain with micro to nano-scale chalcopyrite inclusions from a quartz-magnetite-chalcopyrite

MA

vein from the Safford-Lone Star porphyry deposit, Arizona, USA. (H) Igneous magnetite in quartz-sericite-pyrite (QSP)
altered leuco-granite from the Henderson Climax-type Mo deposit. Abbreviations (after Whitney and Evans, 2010) Cb: carbonate, Ccp: chalcopyrite, Fsp: feldspar, Gn: galena, Grt: garnet, Hem: hematite, Mag: magnetite, Py: pyrite,

ED

Qz: quartz, Sd: siderite, Ser: sericite, Sp: sphalerite, Srp: serpentine.

PT

Fig. 7 Box and whisker plots for important magnetite minor and trace elements in parts per million grouped by deposit
style and magnetite type (hydrothermal vs. igneous). The data are plotted on a logarithmic scale in order of increasing

AC
CE

median concentrations. The upper and lower margins of the box represent the upper and lower 50 percentile of the
data. The whiskers represent the upper and lower threshold values (95 percentile of the data). Median values are
shown as solid black lines and mean values as solid black circles. See text for further discussion.

Fig. 8 Probability plots for important magnetite minor and trace elements in parts per million for the examined
magnetite occurrences. Probability plots show the data value distribution in relation to a normal distribution
(cumulative frequency). Distinct breaks or steps in the distribution curve indicate different populations. See text for
further discussion.

Fig. 9 Radar plots of upper threshold values, in parts per million, for the suite of important magnetite elements. The
upper threshold value (95 percentile of the data) is a robust way to compare upper limits for minor and trace element
concentrations in magnetite from various geological settings. Gaps occur where insufficient data were available to
calculate the upper threshold value. See text for further discussion.

43

ACCEPTED MANUSCRIPT
Fig. 10 False-color field emission gun scanning electron microscope image of (A) hydrothermal-metamorphic
magnetite from Mesoarchean Algoma-type BIF in the Yilgarn craton, Western Australia, with abundant micro- and
nano-scale mineral inclusions. (B) hydrothermal magnetite from the Chino (Santa Rita) porphyry Cu deposit in New

Mexico, USA. The grain displays micro- to nano-scale mineral inclusions that reflect the mineral assemblage of the

RI
P

hydrothermally altered host rock.

Fig. 11 Gallium vs. Sn diagram for the different types of magnetite (values in parts per million). Igneous magnetite

SC

from the Henderson Climax-type Mo deposit and hydrothermal Mg-skarn magnetite from Santa Rita have the highest
Sn concentrations. Igneous magnetite from Henderson also has characteristically high Ga concentrations. Consistent

NU

detectable Sn levels were also found in magnetite from Ag-Pb-Zn veins and BIF-Iron ore deposits but do not exceed
trace amounts (<10 ppm). Overall this plot suggests that lower concentrations of Ga and Sn are linked with decreasing

MA

temperatures. Tin and particularly Ga concentrations are significant lower in cooler temperature magnetite i.e.
magnetite from low temperature BIF and iron ore deposits plots at lower Ga and Sn concentrations than higher

ED

temperature porphyry magnetite (see text and Fig. 13).

Fig. 12 (A) Ti+V vs. Al+Mn plot with deposit fields proposed by Dupuis and Beaudoin (2011) and Nadoll et al.

PT

(submitted) values in weight percent. While the data confirm the proposed skarn field, the porphyry and BIF data

AC
CE

plot outside their suggested boundaries. However, the median for hydrothermal porphyry magnetite would plot in the
center of Dupuis and Beaudoin (2011) porphyry field. The data indicates that there may be a decisive split between
magnesian skarn and calcic skarn which could divide the skarn field into two parts. However, there is not sufficient
data available for magnesian skarn magnetite to make this conclusion with great confidence. (B) The Ti+V vs. Al+Mn
plots displays are clear decreasing temperature trend when plotting the data attributed by their respective formation
temperatures. Note that those temperatures are estimates based on published values. See text for discussion.

Fig. 13 Ti/V and (Ti+V)/(Al+Mn) ratios for igneous and hydrothermal magnetite for the range of investigated geological
settings. Element ratios are independent of the relative concentration of elements in the respective types of
magnetite and therefore represent a robust approach for comparing multiple data sets. (A) and (B) show probability
plots whereas (C) and (D) show boxplots. Ti/V ratios clearly discriminate high V igneous magnetite from the barren IZB
and high Ti igneous magnetite from the Henderson Climax-type Mo deposit. The rest of the data are clustered
relatively closely, indicating that even with decreasing temperatures and overall decreasing Ti and V concentrations
(Fig. 7 and Fig. 8) the Ti/V ratio is relatively constant. The (Ti+V)/(Al+Mn)-ratio plot confirms the decreasing
44

ACCEPTED MANUSCRIPT
temperature trend invoked in Fig. 11, with hotter temperature igneous and porphyry magnetite displaying higher
(Ti+V)/(Al+Mn) ratios than cooler temperature hydrothermal magnetite.

Fig. 14 Ti/V versus Mn diagrams for magnetite from BIF allow the discrimination between magnetite from different

tectonometamorphic settings of metamorphic BIF. (A) unaltered BIF samples, colour-coded by deposition age (B)

RI
P

Hydrothermally altered magnetite shows a large variability among the various iron ore deposits. A decreasing Mn and
Ti/V ratio can be observed in hydrothermally altered magnetite from Hamersley Province deposits, compared to

SC

unaltered magnetite.

NU

Fig. 15 Range of published partition coefficients for igneous magnetite (A) for a variety of host rocks/melts (B). The
figures illustrates that partition coefficients for magnetite vary over orders of magnitudes for the same element

MA

depending on host rock or melt chemistry. A wide range of formation temperatures, pressures, oxygen and sulfur
fugacity are inherently associated with varying host rocks or melts. A complete list of published partition coefficients

AC
CE

PT

ED

and the corresponding references are shown in Table 5.

45

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

Figure 1

46

SC

RI
P

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

Figure 2

47

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

AC
CE

PT

Figure 3

48

Figure 4

AC
CE

PT

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

49

Figure 5

AC
CE

PT

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

50

AC
CE

PT

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

Figure 6
51

Figure 7

AC
CE

PT

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

52

AC
CE

PT

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

Figure 8

53

AC
CE

PT

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

Figure 9
54

SC

RI
P

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

Figure 10

55

AC
CE

PT

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

Figure 11

56

Figure 12

AC
CE

PT

ED

MA

NU

SC

RI
P

ACCEPTED MANUSCRIPT

57

SC

RI
P

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

Figure 13

58

SC

RI
P

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

Figure 14

59

RI
P

ACCEPTED MANUSCRIPT

AC
CE

PT

ED

MA

NU

SC

Figure 15

60

ACCEPTED MANUSCRIPT
TABLE 1

CODES

UQAC

USGS

Location

Universit Chicoutimi, Qubec, Canada

US Geological Survey, Denver, USA

Laser ablation system


ICP-MS
Pulse frequency
Scan type
Stage speed
Beam size
Energy density
Analysis line raster
Gas blank
Signal
Internal standard
Reference material

RESOlution M-50 (Excimer 193 nm)


Agilent 7700x Quadrapole ICP-MS
5 Hz
Line
4 m/s
15-33 m
2
5 J/cm
~240 m long
30 s
60 s
Fe (EMPA)
GSE-1G, BC-28

New Wave Research UP-193 FX (Excimer 193 nm)


PerkinElmer DRC-e ICP
4 Hz
Spot
n/a
15-100 m
2
~5 J/cm
n/a
45-60 s
120 s
Fe (EMPA)
GSE-1G, NIST-610

CR

US

MA
N

ED

ARC Centre of Excellence in Ore Deposits, University of


Tasmania, Australia
New Wave Research UP-213 (Excimer 213 nm)
Agilent HP4500 Quadrapole ICP-MS
5 Hz
Spot
n/a
15-49 m
2
~3.3 J/cm
n/a
30 s
60 s
Fe (EMPA)
STDGLb2, NIST-610

PT

Analysing crystal
TAP
TAP
PET
LiF
LiF
LiF
LiF
LiF
LiF
LiF

CE
AC

EMPA - CSIRO, North Ryde, Australia


Element (X-ray line)
Standard
Mg (K)
Spinel
Al (K)
Spinel
Ti (K)
Rutile
V (K)
V metal
Cr (K)
Cr metal
Mn (K)
Mn metal
Fe (K)
Hematite
Co (K)
Co metal
Ni (K)
Ni metal
Zn (K)
Zn metal

IP
T

LA-ICP-MS

61

ACCEPTED MANUSCRIPT
TABLE 2

0.51
2
0.08
1
1
(10x standard deviation)

5
3

19
9

394
452
1240
47800
873
145000
330
170
2460
70
50
330
70
370
32
130
320
505
19
17
15
57
50
11
51
46
18
19
63
53
35
22

CODES
47m

25m

IP
T

10m
148
166
561
25400
342
64400
160
70
920
29
20
133
30
190
15
56
106
1720
9
7
7
22
23
4
23
19
8
6
107
56

35m

22m

15m

0.04
0.18
2000

0.06
0.29
3590

0.1
0.5

0.1
0.6

0.2
0.7

0.3
0.9

7.26
0.04
0.03
0.63
0.56
0.01
0.03
0.02
0.1
0.04

13.6
0.08
0.05
1.09
0.95
0.01
0.05
0.03
0.17
0.07

398
0.2
0.2
3.53
0.81
0.07
0.41
1.17
1.31
0.12

472
0.2
0.2
3.08
0.96
0.08
0.5
1.09
2.51
0.19

1210
0.4
0.4
7.22
1.24
0.09
1.07
10.63
10.31
0.07

927
0.5
0.5
8.02
1.54
0.2
0.4
3.69
8.15
0.10

0.09

0.16

1.50

4.88

11.97

15.97

0.01
0.07
n/a
0.01

0.10
0.06
n/a
0.02

0.04
0.01
0.02
0.2

0.04
0.02
0.05
0.22

0.12
0.09
0.1
0.49

0.11
0.07
0.2
0.65

0.03

0.05

0.44
0.06

0.84
0.11

0.09
0.01

0.21
0.17

0.54
0.06

0.74
0.06

0.17
0.3
0.01
0.01
0.06
0.1
2
(3x standard deviation)

0.03
0.02
0.02

0.08
0.04
0.06

0.21
0.06
0.04

0.2
0.19
0.06

CR

15m
48
48
219
8550
234
26900
75
35
120
10
8
47
11
60
6
15
34
224
2
2
2
8
8
1
7
6
3
2
5
5

US

25m
14
10
60
2790
215
10100
15
14
36
2
2
12
3
14
1
4
10
108
1
1
0.5
2
2
0.4
2
2
1
0.5
1
1

UQAC
33m

MA
N

563

50m
2.98
1.97
12.1
933
247
4060
9.13
1.51
7.56
1.34
0.38
1.48
0.95
3.37
0.43
<1
1.97
21.2
0.13
<1
0.08
0.24
0.41
0.09
1.21
0.28
0.07
0.05
0.29
0.1

ED

330
406
269
336
350
300

PT

218
494

USGS
100m

CE

North Ryde
<2m

AC

Laboratory
spot size
Na
Mg
Al
Si
S
Ca
Ti
V
Cr
Mn
Co
Ni
Cu
Zn
Ga
Ge
As
Se
Sr
Y
Nb
Mo
Ag
In
Sn
Sb
Ba
Ce
Au
Tl
W
Pb
Th

62

ACCEPTED MANUSCRIPT

Sample

Setting

Location

Domain

Type

Host Rock

6499220

BIF-Iron Ore

Tom Price deposit

Hamersley Province

metamorphic-hydrothermal

6499240

BIF-Iron Ore

Tom Price deposit

Hamersley Province

metamorphic-hydrothermal

NTD9113

BIF-Iron Ore

Tom Price deposit

Hamersley Province

metamorphic-hydrothermal

6499320

BIF-Iron Ore

Tom Price deposit

Hamersley Province

metamorphic-hydrothermal

NTD1173

11

BIF-Iron Ore

Tom Price deposit

Hamersley Province

metamorphic-hydrothermal

BIF-Iron Ore

Hashimoto deposit

Hamersley Province

metamorphic

BIF-Iron Ore

Mt Whaleback deposit

Hamersley Province

metamorphic

WHB1113-307.2

BIF-Iron Ore

Mt Whaleback deposit

Hamersley Province

metamorphic

Easting

Northin
g

Depth/Elevatio
n

n/a

n/a

n/a

n/a

n/a

n/a

n/a

n/a

siderite-silicate magnetite

n/a

n/a

n/a

n/a

siderite-silicate magnetite

n/a

n/a

n/a

n/a

siderite-silicate magnetite

n/a

n/a

n/a

n/a

Alteration
carbonate-talc-chlorite
magnetite
carbonate-talc-chlorite
magnetite

CR

Superior type BIF, Brockman Fm


Superior type BIF, Brockman Fm

US

WHB1113-232.7

Superior type BIF, Brockman Fm

Superior type BIF, Brockman Fm


Superior type BIF, Mt Sylvia Fm

DH Level

Mine

n/a

n/a

7027013

111

580196

7026154

152.7

data source
previously unpublished (W.
UWA)
previously unpublished (W.
UWA)
previously unpublished (W.
UWA)
previously unpublished (W.
UWA)
previously unpublished (W.
UWA)
previously unpublished (T.
UWA)
previously unpublished (T.
UWA)
previously unpublished (T.
UWA)
previously unpublished (T.
UWA)
previously unpublished (T.
UWA)
previosuly
unpublished
University)
previosuly unpublished (A.-S.
UWA)
previosuly unpublished (P.
UWA)
previosuly unpublished (P.
UWA)

n/a

n/a

70

Angerer et al. 2012

n/a

n/a

182.75

Angerer et al. 2012

n/a

n/a

270

Angerer et al. 2012

n/a

n/a

270

Angerer et al. 2012

n/a

n/a

100

Angerer et al. 2012

least-altered

n/a

n/a

n/a

n/a

magnetite-martite

n/a

n/a

n/a

n/a

magnetite-martite

n/a

n/a

n/a

n/a

least-altered

n/a

n/a

n/a

183.6

least-altered

n/a

n/a

n/a

181

least-altered

n/a

n/a

n/a

n/a

MA
N

HH188-208.8

Superior type BIF, Brockman Fm

IP
T

TABLE 3

Superior type BIF, Mt Sylvia Fm

Superior type BIF, Mt Sylvia Fm


Superior type BIF, Marra Mamba
Fm
2
Superior type BIF, Marra Mamba
Fm
2
Superior type BIF, Marra Mamba
Fm

MG368-183.6

BIF-Iron Ore

Mesa Gap deposit

Hamersley Province

metamorphic

MG368-181

BIF-Iron Ore

Mesa Gap deposit

Hamersley Province

metamorphic

77D153a

BIF-Iron Ore

B deposit

Hamersley Province

metamorphic

bn-w9-101_k1

BIF-Iron Ore

BIF-Iron Ore

BIF-Iron Ore

Cerro do Curral
Weld Range,
deposit
Weld Range,
deposit

Iron Quadrangle

metamorphic (supergene-altered)

hydrothermal-supergene

n/a

Yilgarn Craton

metamorphic

dolomite-altered magnetite

WRRC0253d

Yilgarn Craton

metamorphic

least-altered

WRRC1076d
surface
sample

Beebyn
Beebyn

PT

FSD 122-137.85
bn-w10-13_K17b

ED

Superior type BIF, Caue Fm


Algoma type BIF
Algoma type BIF

n/a
582437

BIF-Iron Ore

Koolyanobbing A deposit

Yilgarn Craton

metamorphic-hydrothermal

siderite-altered

BIF-Iron Ore

Koolyanobbing K deposit

Yilgarn Craton

metamorphic-hydrothermal

talc-altered magnetite

K901

BIF-Iron Ore

Koolyanobbing K deposit

Yilgarn Craton

metamorphic-hydrothermal

talc-altered magnetite

K910

BIF-Iron Ore

Koolyanobbing K deposit

Yilgarn Craton

talc-altered magnetite

KP-50

BIF-Iron Ore

Koolyanobbing K deposit

Yilgarn Craton

metamorphic-hydrothermal
metamorphic-hydrothermalsupergene

hydrothermal-supergene

KPDDH017
surface
sample
surface
sample
surface
sample

KK17-31

BIF-Iron Ore

Koolyanobbing K deposit

Yilgarn Craton

metamorphic

least-altered

KPDDH017

n/a

n/a

182.75

Angerer et al. 2012

metamorphic

least-altered

n/a

n/a

165.3

Angerer et al. 2012

metamorphic-contactmetamorphic

magnetite-hematite ore

KPDDH017
surface
sample

n/a

n/a

270

Angerer et al. 2012

metamorphic-hydrothermal

dolomite-altered magnetite

W2DDH007

n/a

n/a

331.7

Angerer et al. 2013

AC

CE

A-1
KK17-31

Algoma type BIF


Algoma type BIF
Algoma type BIF
Algoma type BIF
Algoma type BIF
Algoma type BIF

KK17-17/18

BIF-Iron Ore

Koolyanobbing K deposit

Yilgarn Craton

K110

10

BIF-Iron Ore

Koolyanobbing K deposit

Yilgarn Craton

W2-331.7

BIF-Iron Ore

Windarling W2 deposit

Yilgarn Craton

W2-343.4

BIF-Iron Ore

Windarling W2 deposit

Yilgarn Craton

metamorphic-hydrothermal

dolomite-altered magnetite

W2DDH008

n/a

n/a

343.4

Angerer et al. 2013

W2-377.1

BIF-Iron Ore

Windarling W2 deposit

Yilgarn Craton

metamorphic-hydrothermal

hematite-magnetite ore

W2DDH009

n/a

n/a

377.1

Angerer et al. 2013

W2-380.8

BIF-Iron Ore

Windarling W2 deposit

Yilgarn Craton

metamorphic-hydrothermal

magnetite-hematite ore

W2DDH010

n/a

n/a

380.8

Angerer et al. 2013

metamorphic

least-altered

W2DDH011

n/a

n/a

345.5

Angerer et al. 2013

metamorphic

least-altered

n/a

n/a

n/a

n/a

Battacharya et al. 2004


Battacharya et al. 2005

W2-345.5
H/1

10
5

BIF-Iron Ore
BIF-Iron Ore

Windarling W2 deposit

Yilgarn Craton

Gandhamardan

Algoma type BIF


Algoma type BIF
Algoma type BIF
Algoma type BIF
Algoma type BIF
Algoma type BIF
Algoma type BIF
Algoma type BIF

H/2

BIF-Iron Ore

Deo Nala

metamorphic-hydrothermal

altered by granitic fluids?

n/a

n/a

n/a

n/a

H/4

BIF-Iron Ore

Garumahishani

metamorphic

least-altered

n/a

n/a

n/a

n/a

Battacharya et al. 2006

H/5

Garumahishani

metamorphic

least-altered

n/a

n/a

n/a

n/a

Battacharya et al. 2007

Algoma type BIF


Algoma type BIF

563 GM6#1G3

12

BIF-Iron Ore
IZB
granodiorite

Algoma type BIF (silicate facies)

Japan

Magnetite series granitoid

igneous

granodiorite

least-altered

n/a

n/a

n/a

n/a

Spong 1998, Nadoll et al. 2012

AU58431

Ag-Pb-Zn

Lucky Friday

Gold Hunter

hydrothermal

calcite vein

siderite

4900 Level

591793

-1493

Nadoll et al. 2012

AU58451

Ag-Pb-Zn

Lucky Friday

Gold Hunter

hydrothermal

siderite vein

siderite

5900 Level

18060.79

5258197
24482.0
8

AU58484

Ag-Pb-Zn

Lucky Friday

Gold Hunter

hydrothermal

siderite vein

siderite

61-05

AU58493

Ag-Pb-Zn

Lucky Friday

Gold Hunter

AU60865

Mg-Skarn

Chino (Santa Rita)

hydrothermal

siderite/siltite

siderite

61-05

hydrothermal

magnesian skarn

skarn

1991

1701*

606*

-2464.87

Nadoll et al. 2012

29

Nadoll et al. 2012

421.5

Nadoll et al. 2012

1330

Nadoll et al. submitted

Thorne,
Thorne,
Thorne,
Thorne,
Thorne,
Angerer,
Angerer,
Angerer,
Angerer,
Angerer,
(Curtin
Hensler
Duuring,
Duuring,

63

ACCEPTED MANUSCRIPT
AU60869

Mg-Skarn

Chino (Santa Rita)

hydrothermal

magnesian skarn

skarn

1991

AU60908

Porphyry

Chino (Santa Rita)

hydrothermal

granodiorite

potassic-sericitic

D2245
2305

AU60916

Porphyry

Chino (Santa Rita)

hydrothermal

quartzite

calcic sodic

AU60895

Porphyry

Cobre

hydrothermal

hornfels

propylitic

1701*

606*

2478.1*

867.4*

772378

3637481

1454

Nadoll et al. submitted

30

Nadoll et al. submitted

2148

Nadoll et al. submitted


Nadoll et al. submitted

Porphyry

Morenci

hydrothermal

diabase

argillic

3898

-8000*

9800*

Porphyry

Morenci

hydrothermal

diabase

argillic

3898

-8000*

9800*

sericitic

3925

-17170*

11750*

240

Nadoll et al. submitted

sericitic

3944

-16200*

21600*

1502

Nadoll et al. submitted

sericitic

3944

-16200*

21600*

1555

Nadoll et al. submitted

propylitic

RL-1

20437*

4050*

1534

Nadoll et al. submitted

Porphyry

Morenci-Producer

hydrothermal

diorite

Porphyry

Morenci-Sun Ridge

hydrothermal

granite

AU61015

Porphyry

Morenci-Sun Ridge

hydrothermal

granite

AU60967

Porphyry

Safford-Dos Pobres

hydrothermal

andesite

AU60962

Porphyry

Safford-Dos Pobres

hydrothermal

andesite

AU60976

Porphyry

Safford-Dos Pobres

hydrothermal

granodiorite

AU60940

Porphyry

Safford-Lone Star

hydrothermal

andesite

AU60942

Porphyry

Safford-Lone Star

hydrothermal

andesite

AU60952

Porphyry

Safford-Lone Star

hydrothermal

andesite

AU60947

Porphyry

Safford-Lone Star

hydrothermal

andesite
andesite

Porphyry

Safford-San Juan

hydrothermal

Porphyry

Safford-San Juan

hydrothermal

AU60859

Skarn

Chino (Santa Rita)

hydrothermal

AU60875

Skarn

Chino (Santa Rita)

hydrothermal

AU60910

Skarn

Chino (Santa Rita)

hydrothermal

AU60930

Skarn

Chino (Santa Rita)

hydrothermal

AU60898

Skarn

Cobre

hydrothermal

AU60903

Skarn

Cobre

hydrothermal

AU60884

Skarn

Copper Flat

hydrothermal

Skarn

Morenci-Shannon

hydrothermal

Porphyry

Chino (Santa Rita)

igneous

AU60914

Porphyry

Chino (Santa Rita)

igneous

Porphyry

Chino (Santa Rita)

igneous

Porphyry

Chino (Santa Rita)

igneous

AU60893

Porphyry

Cobre

igneous

CE

AU60919
AU60921

PT

AU61024
AU60857

Nadoll et al. submitted


Nadoll et al. submitted

propylitic

RL-1

20437*

4050*

1464

Nadoll et al. submitted

propylitic

RL-1

20437*

4050*

2027

Nadoll et al. submitted

potassic-argillic

LS07-0153

1489

Nadoll et al. submitted

potassic-argillic

LS06-0083

Nadoll et al. submitted

LS06-0083

8799.35
*
8799.35
*

2173

potassic-sericitic oxidized

41201.81
*
41201.81
*

propylitic

LS07-0154

1411.5

Nadoll et al. submitted

1932

Nadoll et al. submitted

calcic sodic

SJP-64

27600*

4109*

1085

Nadoll et al. submitted

calcic sodic

SJP-64

27600*

4109*

1141

Nadoll et al. submitted

skarn

skarn

2209

5810*

1310*

1430

Nadoll et al. submitted

skarn

skarn

2166

5400*

1000*

2152

Nadoll et al. submitted

skarn

skarn

D2245

76.7

Nadoll et al. submitted

skarn

skarn

2330

-3832.9*

673.5*

skarn

skarn

772341

3637425

skarn

skarn

772232

3637506

skarn

skarn

769403

3633542

skarn

skarn

granodiorite

sericitic

granodiorite

potassic

granodiorite breccia

potassic

monzonite

potassic

granite

potassic

1004
2003

22266*

5300 onramp

1310*

1398

774830

3632319

1689-x

-115.07*

108.76*

783.5

1689-x

-115.07*

108.76*

773

772352

3637691

5180*

7405*

Climax-Mo

Henderson

igneous

leuco- rhyolite/granite

hypogene

NEHW

Climax-Mo

Henderson

igneous

leuco- rhyolite/granite

hypogene

662

AU61032

Climax-Mo

Henderson

igneous

leuco- rhyolite/granite

hypogene

P87

5188*

AU61030

Climax-Mo

Henderson

igneous

leuco- rhyolite/granite

hypogene

U87

AU61028

Climax-Mo

Henderson

igneous

leuco- rhyolite/granite

hypogene

U93

AU61029

Climax-Mo

Henderson

igneous

leuco- rhyolite/granite

hypogene

AU61017

Porphyry

Morenci

igneous

diabase

AU61018

Porphyry

Morenci

igneous

Nadoll et al. submitted


Nadoll et al. submitted

-9959*

AU61033

Nadoll et al. submitted


Nadoll et al. submitted

5810*

2209

AU61031

AC

1044

andesite

MA
N

8
10

ED

AU60990
AU60991

US

AU61022
AU61011

CR

IP
T

AU61017
AU61018

Nadoll et al. submitted


Nadoll et al. submitted
Nadoll et al. submitted
Nadoll et al. submitted
Nadoll et al. submitted
Nadoll et al. submitted

7150 level

previously unpublished

957.4

previously unpublished

7442*

7210 level

previously unpublished

5030*

7164*

7270 level

previously unpublished

4834*

7216*

7270 level

previously unpublished

U93-1

4790*

7100*

7270 level

previously unpublished

argillic

3898

-8000*

9800*

1044

diabase

argillic

3898

-16300*

21400*

Nadoll et al. submitted


Nadoll et al. submitted

AU61027

Porphyry

Morenci-Garfield

igneous

diabase

n/a

-10365*

27015*

5150 onramp

AU61022

Porphyry

Morenci-Producer

igneous

diorite

sericitic

3925

-17170*

11750*

240

Nadoll et al. submitted

AU61009

Porphyry

Morenci-Sun Ridge

igneous

granite

sericitic-propylitic

3944

-16200*

21600*

1462

Nadoll et al. submitted

AU60999

Porphyry

Safford-Dos Pobres

igneous

granodiorite

potassic

RL-26

Porphyry

Safford-Lone Star

igneous

andesite

potassic-oxidized

LS06-0083

4049*
8799.35
*

Nadoll et al. submitted

20813*
41201.81
*

1406

AU60957

1434

Nadoll et al. submitted

AU61007

Porphyry

Safford-Lone Star

igneous

andesite

potassic-oxidized

LS07-0140

1506

Nadoll et al. submitted

AU60934

Porphyry

Safford-Lone Star

igneous

andesite

propylitic

LS07-0154

1919

Nadoll et al. submitted

AU60936

Porphyry

Safford-Lone Star

igneous

andesite

propylitic

LS07-0151

1757.5

Nadoll et al. submitted

AU60980

Porphyry

Safford-San Juan

igneous

granodiorite

potassic-sericitic

SJP-70

26742*

4066*

520

Nadoll et al. submitted

AU60982

Nadoll et al. submitted

Porphyry

Safford-San Juan

igneous

granodiorite

potassic

SJP-70

26742*

4066*

720.7

Nadoll et al. submitted

126335

11**

Porphyry

Ertsberg district

Karume Intrusion

hydrothermal

monzodiorite

propylitic

CTS-1

735386

9550242

162.5

previously unpublished

126336

18**

Porphyry

Ertsberg district

Karume Intrusion

hydrothermal

monzodiorite

phyllic

CTS-1

735386

9550242

177.2

previously unpublished

126337

16**

Porphyry

Ertsberg district

Karume Intrusion

hydrothermal

monzodiorite

propylitic

CTS-1

735386

9550242

376.9

previously unpublished

126340

5**

Porphyry

Ertsberg district

Karume Intrusion

hydrothermal

monzodiorite

potassic

CTS-1

735386

9550242

698.5

previously unpublished

126312

7**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

736966

9548885

17.05

previously unpublished

64

ACCEPTED MANUSCRIPT
14**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

736972

9548891

26.46

previously unpublished

126314

5**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

736976

9548896

30.65

previously unpublished

126317

15**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

736984

9548904

44.38

previously unpublished

126318

17**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

737002

9548923

70

previously unpublished

126320

10**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

737037

9548959

121.74

previously unpublished

126323

19**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

737070

9548993

168.46

previously unpublished

126324

20**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

737076

9549000

179.2

previously unpublished

126325

12**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

737080

9549004

183

previously unpublished

126326

12**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

737088

9549012

197.35

previously unpublished

126327

11**

Porphyry

Ertsberg district

Ertsberg Diorite

hydrothermal

diorite

phyllic

FAS3-1

737094

9549019

204.38

previously unpublished

126328

17**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

737098

9549023

212.8

previously unpublished

126329

16**

Porphyry

Ertsberg district

Ertsberg Diorite

hydrothermal

diorite

phyllic

FAS3-1

737101

9549026

215.1

previously unpublished

126330

17**

Porphyry

Ertsberg district

Ertsberg Diorite

hydrothermal

diorite

phyllic

FAS3-1

737104

9549029

221

previously unpublished

126331

10**

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

skarn

FAS3-1

737106

9549032

222.3

previously unpublished

126341

11**

Porphyry

Ertsberg district

North Grasberg Intrusion

hydrothermal

diorite

propylitic

FNG-1

733862

9553477

126.35

previously unpublished

126342

12**

Porphyry

Ertsberg district

North Grasberg Intrusion

hydrothermal

diorite

phyllic

FNG-1

733857

9553435

265.4

previously unpublished

126343

12**

Porphyry

Ertsberg district

North Grasberg Intrusion


Main
Grasberg
Intrusion
(Early)
Main
Grasberg
Intrusion
(Early)
Main
Grasberg
Intrusion
(Middle)
Main
Grasberg
Intrusion
(Middle)
Main
Grasberg
Intrusion
(Middle)

hydrothermal

diorite

potassic

FNG-1

733852

9553399

379.3

previously unpublished

hydrothermal

monzodiorite

n/a

GRS24

734427

9551267

0.3

previously unpublished

monzodiorite

n/a

GRS24

734427

9551267

159.9

previously unpublished

monzodiorite

n/a

GRS32

734513

9551321

2.7

previously unpublished

monzodiorite

n/a

GRS32

734513

9551321

149.75

previously unpublished

monzodiorite

n/a

GRS32

734513

9551321

400

previously unpublished

monzodiorite

potassic

GRS32

734513

9551321

610.05

previously unpublished

monzodiorite

potassic

GRS32

734513

9551321

690

previously unpublished

Porphyry

Ertsberg district

Porphyry

Ertsberg district

126380

16**

Porphyry

Ertsberg district

126385

15**

Porphyry

Ertsberg district

126393

10**

Porphyry

Ertsberg district

126400

21**

Porphyry

Ertsberg district

South Kali

hydrothermal

126403

4**

Porphyry

Ertsberg district

South Kali

hydrothermal

hydrothermal
hydrothermal
hydrothermal

1**

Porphyry

Ertsberg district

Dalam Diorite

hydrothermal

1**

Porphyry

Ertsberg district

Dalam Diorite

hydrothermal

126456B

17**

Porphyry

Ertsberg district

Dalam Diorite

126459

16**

Porphyry

Ertsberg district

Dalam Diorite

126472

22**

Porphyry

Ertsberg district

126476

18**

Porphyry

Ertsberg district

South Kali
Main
Grasberg
(Early)

126479

15**

Porphyry

Ertsberg district

126482

20**

Porphyry

Ertsberg district

126486

16**

Porphyry

CR

diorite

potassic

GRS37-120

734289

9551082

416.9

previously unpublished

diorite

potassic

GRS37-120

734223

9550997

582.45

previously unpublished

hydrothermal

diorite

potassic

GRS37-120

734223

9550997

582.45

previously unpublished

hydrothermal

diorite

phyllic-potassic

GRS37-120

734152

9550908

763.15

previously unpublished

PT

126452

ED

hydrothermal

126456A

US

18**
16**

MA
N

126369
126375

IP
T

126313

monzodiorite

phyllic

GRS37-217

734536

9551394

230.5

previously unpublished

hydrothermal

hydrothermal

monzodiorite

n/a

GRS37-217

734590

9551461

400.5

previously unpublished

Dalam Diorite

hydrothermal

diorite

potassic

GRS37-217

734621

9551499

500.35

previously unpublished

Dalam Diorite

hydrothermal

diorite

potassic

GRS37-217

734662

9551549

630.25

previously unpublished

Ertsberg district

Border Phase

hydrothermal

diorite

massive sulphide

GRS37-217

734692

9551587

730.4

previously unpublished

AC

CE

Intrusion

126489

1**

Porphyry

Ertsberg district

Border Phase

hydrothermal

limestone

massive sulphide

GRS37-217

734726

9551629

841.3

previously unpublished

126495

19**

Porphyry

Ertsberg district

Main Grasberg Intrusion (Late)

hydrothermal

monzodiorite

phyllic

GRS37-219

734489

9551334

198.7

previously unpublished

126503A

9**

Porphyry

Ertsberg district

South Kali

hydrothermal

monzodiorite

phyllic

GRS37-219

734536

9551378

500.1

previously unpublished

126503Bi

5**

Porphyry

Ertsberg district

South Kali

hydrothermal

monzodiorite

phyllic

GRS37-219

734536

9551378

500.1

previously unpublished

126503Bii

6**

Porphyry

Ertsberg district

South Kali
Main
Grasberg
(Middle)

hydrothermal

monzodiorite

phyllic

GRS37-219

734536

9551378

500.1

previously unpublished

hydrothermal

monzodiorite

phyllic

GRS37-219

734566

9551406

651.15

previously unpublished

126506

18**

Porphyry

Ertsberg district

126516

19**

Porphyry

Ertsberg district

126406

16**

Porphyry

Ertsberg district

126412

13**

Porphyry

Ertsberg district

Dalam Andesite
Main
Grasberg
(Middle)
Main
Grasberg
(Early)

126418

7**

Porphyry

Ertsberg district

Dalam Diorite

126420

1**

Porphyry

Ertsberg district

126424

14**

Porphyry

Ertsberg district

126433

17**

Porphyry

Ertsberg district

Dalam Diorite
Main
Grasberg
(Middle)
Main
Grasberg
(Middle)

126534

7**

Porphyry

Ertsberg district

Dalam Andesite

126540

16**

Porphyry

Ertsberg district

126547

9**

Porphyry

Ertsberg district

Dalam Andesite
Main
Grasberg
(Middle)

126551

16**

Porphyry

Ertsberg district

Dalam Andesite

Intrusion

hydrothermal

andesite

phyllic

GRS37-260

734964

9551624

1.5

previously unpublished

hydrothermal

monzodiorite

potassic

GRS37-45

734426

9551264

100.4

previously unpublished

hydrothermal

monzodiorite

potassic

GRS37-45

734318

9551171

359.7

previously unpublished

hydrothermal

diorite

potassic

GRS37-45

734197

9551075

600.1

previously unpublished

hydrothermal

diorite

potassic

GRS37-45

734159

9551047

700.15

previously unpublished

hydrothermal

monzodiorite

phyllic

GRS37-59

734537

9551410

0.1

previously unpublished

hydrothermal

monzodiorite

potassic

GRS37-59

734633

9551537

159.1

previously unpublished

hydrothermal

andesite

phyllic

GRS93B

734220

9551567

0.1

previously unpublished

hydrothermal

andesite

phyllic

GRS93B

734220

9551567

251.5

previously unpublished

hydrothermal

monzodiorite

potassic

GRZ37-43

734408

9551236

83.6

previously unpublished

hydrothermal

andesite

potassic

GRZ37-43

734346

9551159

180

previously unpublished

Intrusion
Intrusion

Intrusion
Intrusion

Intrusion

65

ACCEPTED MANUSCRIPT
126554

20**

Porphyry

Ertsberg district

Dalam Andesite

hydrothermal

andesite

potassic

GRZ37-43

734296

9551097

262

126557

13**

Porphyry

Ertsberg district

Dalam Andesite

hydrothermal

andesite

potassic

GRZ37-43

734241

9551028

349.2

previously unpublished
previously unpublished

126560

23**

Porphyry

Ertsberg district

Dalam Andesite

hydrothermal

andesite

phyllic-potassic

GRZ37-43

734171

9550942

459.5

previously unpublished

126565

3**

Porphyry

Ertsberg district

Dalam Andesite

hydrothermal

andesite

phyllic

GRZ37-43

734074

9550822

620

previously unpublished

126344A

1**

Porphyry

Ertsberg district

Kay Intrusion

hydrothermal

n/a

phyllic

KI1-1

734337

9549191

571.15

previously unpublished

126344B

5**

Porphyry

Ertsberg district

Kay Intrusion

hydrothermal

n/a

propylitic

126566

22**

Porphyry

Ertsberg district

Dalam Andesite

hydrothermal

andesite

126568

24**

Porphyry

Ertsberg district

Dalam Andesite

hydrothermal

andesite

126574

24**

Porphyry

Ertsberg district

Dalam Andesite

hydrothermal

andesite

126345

1**

Porphyry

Ertsberg district

Dalam Volcaniclastics

hydrothermal

n/a

126350

18**

Porphyry

Ertsberg district

Ring Dyke

hydrothermal

n/a

126347

20**

Porphyry

Ertsberg district

Dalam Andesite

hydrothermal

andesite

126333

12**
**EMPA
data

Skarn

Ertsberg district

Ertsberg East Skarn

hydrothermal

skarn

*mine grid coordinates

previously unpublished

previously unpublished

phyllic

SGR-2

734030

9550703

50

previously unpublished

phyllic

SGR-2
surface
sample
surface
sample
surface
sample
surface
sample

734027

9550649

178

IP
T

571.15

9550724

CR

phyllic

potassic

US

propylitic

skarn

previously unpublished
previously unpublished
previously unpublished
previously unpublished
previously unpublished

PT

ED

MA
N

Mesoproterozoic

9549191

734032

CE

Neoarchean

734337

SGR-2

AC

Paleoproterozoic

KI1-1

phyllic

66

ACCEPTED MANUSCRIPT
TABLE 4

0.95
10000
3210
11100
219
318
32300
19.4
150
214
8.74
17.5

Mg-Skarn - hydrothermal
#
Min
Max
Med
17
5140
26700
21900
9
1100
2240
1810
4
15.2
86.1
19.2
8
15
281
20.1
0
<dl
<dl
<dl
17
972
1610
1390
4
2.64
4.09
3.44
0
<dl
<dl
<dl
9
175
311
220
10
25.9
43.4
28.4
10
67.6
99.1
82.5

0.95
26700
2240
86.1
281
<dl
1610
4.09
<dl
311
43.4
99.1

Skarn - hydrothermal
#
Min
Max
203
49.7
59400
214
580
31700
183
134
7180
51
38.2
2510
56
270
4720
229
350
50400
59
31.3
990
41
61.3
920
194
79.4
15100
36
6.68
134
6
8.85
53.1

Med
30700
9500
3020
518
458
17500
440
490
7640
27.7
23.4

0.95
54000
16700
6410
2310
3800
46100
950
800
13300
129
53.1

0.95
24700
7810
19600
3800
696
8890
646
724
3090
115
70.1

Porphyry - igneous
#
Min
Max
65
103
3940
75
459
9720
94
191
97800
96
367
6660
14
194
3290
99
115
10500
62
4.95
131
28
21.7
448
54
48
2430
72
15.1
108
10
2.3
67.6

Med
478
2260
1650
1970
544
464
45.6
77.8
456
59.4
7.78

0.95
1940
7770
76300
3960
3290
7080
83.1
434
2000
98.3
67.6

Climax-Mo - igneous
#
Min
Max
2
227
508
41
2000
19100
41
553
11200
9
74.1
1160
0
<dl
<dl
41
737
12500
13
10.6
78.2
1
403
403
34
253
3300
41
94.6
280
35
42.5
210

0.95
508
17900
10700
1160
<dl
11300
78.2
403
2920
259
175

IZB - igneous
#
Min
9
302
12
1260
12
50
12
2210
5
32.6
12
441
12
31.9
12
33
12
50
7
15.3
0
<dl

Med
543
1820
168
3600
68.2
1540
76.1
73.6
249
69.9
<dl

0.95
844
3330
419
4140
198
3200
502
192
2080
150
<dl

CR

US

MA
N

ED

IP
T

Med
129
362
154
136
69.6
80.2
9.3
85.2
58.4
3.83
2.36

PT

Med
840
2300
1930
1470
370
1000
380
400
530
47.7
7.78

Ag-Pb-Zn - hydrothermal
#
Min
Max
29
4.86
11500
30
145
4490
30
9.25
21900
26
1.09
232
20
5.59
320
30
21.4
45000
18
1.06
19.4
20
1.11
152
30
17.6
216
27
0.05
9.4
23
0.17
19.6

CE

Mg, ppm
Al, ppm
Ti, ppm
V, ppm
Cr, ppm
Mn, ppm
Co, ppm
Ni, ppm
Zn, ppm
Ga, ppm
Sn, ppm

Porphyry - hydrothermal
#
Min
Max
583
110
32200
750
384
15900
733
176
37900
715
57.8
18200
78
270
2240
748
77.2
19300
143
8.47
760
183
15.9
1090
563
20.1
6210
86
12.9
151
14
2.53
70.1

0.95
2160
841
105
15.7
48.9
251
5.75
31.2
30.7
2.4
1.14

AC

Mg, ppm
Al, ppm
Ti, ppm
V, ppm
Cr, ppm
Mn, ppm
Co, ppm
Ni, ppm
Zn, ppm
Ga, ppm
Sn, ppm

BIF - Iron ore, hydrothermal


#
Min
Max
Med
98
14.8
2880
556
137
0.27
1510
86.5
134
0.8
117
17.8
135
0.26
40.5
4.83
73
1.14
356
5.17
137
3.66
507
41.8
124
0.03
21.9
0.8
120
0.27
36.4
2.36
104
1.15
38.9
4.87
107
0.01
5.2
0.93
37
0.17
1.48
0.3

Med
367
5220
4040
272
<dl
5910
31.5
403
651
181
88.2

Max
844
3330
419
4140
198
3200
502
192
2080
150
<dl

67

ACCEPTED MANUSCRIPT
TABLE 5

Basalt-Andesite-Dacite
Basalt-Hawaiite
Basalt-Latite
Basalt-Trachyte-Hawaite
Dacite
Granodiorite
High Silica Rhyolite

Low Silica Rhyolite


Mugearite
n/a
Pyroclastics
Quartz Latite
Rhyolite

Tholeiite
Trachyandesite
Trachyte

Fe

Ga
24.1

Mg
4.1

3.2
2

0.2
0.12

153

2.1

10.4

24.7
2.16

166

17
4.3

28.6

62.7
222

8.1
2.6

1.4
97
96
82.1
96
29

12
19.4

6.11

6.87
33.9
6.87
26

3.1

87

2.8

8.4

15.5
18.4

0.23

16.5
25

4.32

Zn

0.7
0.86

10.3

150.4

Ti
2.9

32
3.5
5.2

1.5

202

Pb

5.33

172
275
4.2

149

Ni

1.81

19.2

850

210
62.9

Nb

MA
N

19.4
7.4

Mo

US

23

CR

433
37
13.3
3.4

Mn
5.72

IP
T

Cr

ED

Basalt-Anakaramite
Basalt-Andesite

Co

PT

Basalt

Al

CE

Andesite-Dacite

Reference
Ewart & Griffin 1994
Luhr & Carmichael 1980
Ewart et al. 1973
Nielsen & Beard 2000
Horn et al. 1994
Lemarchand et al. 1987
Lindstrom 1976
Nielsen 1992
Righter et al. 2006
Nielsen et al. 1994
Esperana et al. 1997
Haskin et al. 1966
Klemme et al. 2006
Reid 1983
Okamoto 1979
Lemarchand et al. 1987
Leeman et al. 1978
D'Orazio et al. 1998
Villemant et al. 1981
Ewart & Griffin 1994
Latourrette et al. 1991
Ewart & Griffin 1994
Mahood & Hildreth 1983
Streck & Grunder 1997
Ewart & Griffin 1994
Lemarchand et al. 1987
Kretz et al. 1999
Toplis & Corgne 2002
Dudas et al. 1973
Ewart & Griffin 1994
Bacon & Druitt 1988
Nash & Crecraft 1985
Tacker & Candela 1987
Leeman 1974
Luhr et al. 1984
Lemarchand et al. 1987
Mahood & Stimac 1990

AC

Rock/Melt Type
Andesite

14.1
15
49.7
49.5
44.4
30

27
29

12
0.71

26.6
130

0.8

58.9
63
61.8
50.1

29
1.62

8.52

11
1.5

0.6

1.6
3.2

30.9

33.3

9.3
16.2

60
115.2
80

30
218

12

65

22
0.73

1.58

13
34.1
39.4
75.2
15
215

31.6
72
41.7
21

154
8

17
11

68

You might also like