You are on page 1of 32

Ore Geology Reviews 61 (2014) 1–32

Contents lists available at ScienceDirect

Ore Geology Reviews


journal homepage: www.elsevier.com/locate/oregeorev

Review

The chemistry of hydrothermal magnetite: A review


Patrick Nadoll a,⁎, Thomas Angerer b, Jeffrey L. Mauk c, David French d, John Walshe a
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
U.S. 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

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

Article history: Magnetite (Fe3O4) is a well-recognized petrogenetic indicator and is a common accessory mineral in many ore
Received 1 May 2013 deposits and their host rocks. Recent years have seen an increased interest in the use of hydrothermal magnetite
Received in revised form 20 December 2013 for provenance studies and as a pathfinder for mineral exploration. A number of studies have investigated how
Accepted 20 December 2013
specific formation conditions are reflected in the composition of the respective magnetite. Two fundamental
Available online 14 January 2014
questions underlie these efforts — (i) How can the composition of igneous and, more importantly, hydrothermal
Keywords:
magnetite be used to discriminate mineralized areas from barren host rocks, and (ii) how can this assist explo-
Magnetite ration geologists to target ore deposits at greater and greater distances from the main mineralization? Similar to
Hydrothermal igneous magnetite, the most important factors that govern compositional variations in hydrothermal magnetite
Mineral deposits are (A) temperature, (B) fluid composition — element availability, (C) oxygen and sulfur fugacity, (D) silicate and
Provenance studies sulfide activity, (E) host rock buffering, (F) re-equilibration processes, and (G) intrinsic crystallographic controls
Indicator mineral such as ionic radius and charge balance. We discuss how specific formation conditions are reflected in the com-
Petrogenetic position 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 mag-
netite (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 d'Alene district, United States, (3) porphyry
Cu–(Au)–(Mo) deposits and associated (4) calcic and magnesian skarn deposits in the southwestern United
States 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.
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 N 1000 ppm) and display systematic variations. We propose a com-
bination 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 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 and Sn; (II) Ag–Pb–Zn veins (hydrothermal) — high
Mn and low Ga and Sn; (III) Mg-skarn (hydrothermal) — high Mg and Mn and low Al, Ti, Cr, Co, Ni and Ga; (IV)
skarn (hydrothermal) — high Mg, Al, Cr, Mn, Co, Ni and Zn and low Sn; (V) porphyry (hydrothermal) — high Ti
and V and low Sn; (VI) porphyry (igneous) — high Ti, V and Cr and low Mg; and (VII) Climax-Mo (igneous) —
high Al, Ga and Sn and low Mg and Cr.
© 2014 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1. Mineralogy and crystallography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2. Magnetite as a petrogenetic indicator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3. Magnetite in exploration and provenance studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4. Defining hydrothermal magnetite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2. Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1. Electron microprobe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

⁎ Corresponding author.

0169-1368/$ – see front matter © 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.oregeorev.2013.12.013
2 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

2.2. Laser ablation inductively coupled plasma mass spectrometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6


2.3. Data treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3. Geologic background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1. Banded iron formations and related iron ore deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1.1. Meso-, and Neoarchean Algoma-type BIF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1.2. Neoarchean–Paleoproterozoic Superior-type BIF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2. Ag–Pb–Zn veins (Lucky Friday, Coeur d'Alene district, Idaho) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.3. Porphyry Cu–(Au)–(Mo) and related skarn deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.3.1. American Southwest (Arizona, New Mexico) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.3.2. Ertsberg district (Papua, Indonesia) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4. Plutonic igneous rocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4.1. Climax-type Mo deposit (Henderson, Red Mountain), Colorado . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4.2. Inner Zone Batholith (Japan) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4. Magnetite chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.1. Deposit type comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.1.1. Banded iron formation and related iron ore deposits (158 analyses) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.1.2. Ag–Pb–Zn veins (30 analyses) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.1.3. Calcic and magnesian skarn (229 and 17 analyses) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.1.4. Porphyry Cu–(Au)–(Mo) deposits — hydrothermal and igneous (840 and 100 analyses) . . . . . . . . . . . . . . . . . . . . . 15
4.1.5. Plutonic igneous rocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.1.6. Magnetite from iron oxide copper–gold deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.2. Classification of orebody magnetite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.3. Controlling factors and processes for the composition of magnetite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3.1. Crystallographic factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3.2. Igneous processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3.3. Hydrothermal processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.4. Pseudomorphic magnetite alteration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4.1. Martitization and mushketovitization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5. Inclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
6. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

1. Introduction Fleet, 1981), where A represents a divalent cation such as Mg, Fe2 +,
Ni, Mn, Co, or Zn, and B represents a trivalent cation such as Al, Fe3+,
Magnetite (Fe3 O4 ) is an important petrogenetic indicator and Cr, V, Mn, or Ga (Lindsley, 1976a; Wechsler et al., 1984). Titanium,
pathfinder mineral with a wide array of applications including geophysical with a 4 + charge, can also occupy the B site when substitution is
studies, igneous petrology, provenance studies and mineral exploration coupled with a divalent cation (Wechsler et al., 1984). Octahedral
(e.g., Dupuis and Beaudoin, 2011; Ghiorso and Sack, 1991b; Grant, 1984/ sites in the magnetite structure are randomly occupied by subequal
85a; Lindsley, 1976a; McClenaghan, 2005; Razjigaeva and Naumova, numbers of ferric (Fe3+) and ferrous (Fe2+) iron atoms, whereas tetra-
1992). The compositional variability of magnetite in response to varying hedral sites are exclusively occupied by the smaller ferric iron atoms
formation conditions has been the focus of many studies over the last Fe3 +[Fe2 +Fe3 +]O4 (Lindsley, 1976a; Waychunas, 1991; Wechsler
few decades. Recent years have seen particular interest in hydrothermal et al., 1984) (Fig. 1A).
magnetite as a pathfinder mineral for exploration — facilitated by the Magnetite displays many phase transitions with other spinel group
development and improvement of analytical techniques such as laser minerals, including but not limited to, spinel (MgAl2O4), ulvöspinel
ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) (Fe2TiO4), chromite (FeCr2O4), galaxite (Mn,Mg)(Al,Fe)2O4, gahnite
and trace mode electron microprobe analysis (EMPA), which allow in- (ZnAl2O4), franklinite ((Zn, Fe,Mn)(Fe,Mn)2O4), and jacobsite (Mn,Fe,
situ measurements with increasingly lower detection limits (e.g., Mg)(Fe,Mn)2O4. Above 600 °C a continuous solid solution exists be-
Dupuis and Beaudoin, 2011; Longerich et al., 1996; Nadoll and Koenig, tween magnetite and ulvöspinel (titanomagnetitess — TixFe3 − xO4),
2011). We review compositional trends in magnetite from a variety of and their oxidation products (titanomaghemite), with coupled substi-
hydrothermal ore deposits and their host rocks and discuss how these tution of Ti4+ for Fe3+ in the octahedral sites and Fe2+ for Fe3+ in the
compositional variations can be used as a geochemical fingerprint. We tetrahedral sites (Buddington and Lindsley, 1964; Waychunas, 1991;
present previously published and unpublished minor and trace element White et al., 1994). Below 600 °C thermodynamic data are poorly
data for hydrothermal and igneous magnetite from (1) banded iron constrained and extensive miscibility gaps occur (Ghiorso and Sack,
formations (BIF) and related high-grade iron ore deposits in Western 1991b). A log fO2–T diagram with relevant buffers for the Fe–Si–O sys-
Australia, India, and Brazil, (2) Ag–Pb–Zn veins of the Coeur d'Alene tem and a schematic phase diagram for the system Fe–O–S in fO2–fS2
district, United States, (3) porphyry Cu–(Au)–(Mo) deposits and as- space are shown in Fig. 2A and B (redrawn after Frost, 1991a; Frost
sociated (4) calcic and magnesian skarn deposits in the southwest- et al., 1988). The minimum oxygen fugacity for any given temperature
ern United Sates (Chino, Cobre, Copper Flat, Safford, Morenci) and at which magnetite is stable is the iron-magnetite (IM) or magnetite–
Indonesia (Ertsberg district), and (5) plutonic igneous rocks from the wüstite (MW) buffer (Buddington and Lindsley, 1964; Frost, 1991a).
Henderson Climax-type Mo deposit, United States, and the un- The fayalite–magnetite–quartz (FMQ) buffer marks the limit above
mineralized Inner Zone Batholith granodiorite, Japan. which Fe is mostly incorporated into magnetite. Below the FMQ buffer,
iron is predominantly present in silicates. The quartz–iron–fayalite
1.1. Mineralogy and crystallography (QIF) buffer marks the limit below which Fe occurs in its native
state. The upper limit for magnetite stability is defined by the hematite–
Magnetite belongs to the space group Fd3m and has an inverse spi- magnetite (HM) buffer, beyond which hematite is the dominant oxide
nel structure with the general stoichiometry AB2O4 (Bragg, 1915; mineral (Buddington and Lindsley, 1964; Frost, 1991a; Grant, 1984/
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 3

Fig. 1. Cubic inverse spinel structure of magnetite and common cations that can substitute for Fe3+ in the tetrahedral sites and Fe2+/Fe3+ 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). The incorporation of foreign
cations is more likely when the substituting cation has a similar charge and ionic radius (±15–18% radius variations, after Goldschmidt's rule — Goldschmidt, 1954). Coupled substitution
is required for the incorporation of 4+ cations such as Ti4+ 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. High relative element abundance in the mineralizing fluid is a further controlling factor for
substitution of largely incompatible cations (e.g., Si, Al).
Data from Shannon (1976).

85a). Cation substitution is generally more likely to take place at lower used to help model the differentiation of magmatic intrusions (Toplis
oxygen fugacity (Buddington and Lindsley, 1964; Frost, 1991a; Lindsley, and Corgne, 2002). Similar applications are relevant for hydrothermal
1991). Magnesium, Mn, Zn, and Ni may substitute Fe2+, whereas Fe3+ conditions. For example, magnetite and coexisting hematite have been
can be replaced by Al, V and Cr (Fig. 1B, C) (e.g., Barnes and Roeder, used to define redox-potentials in hydrothermal fluids associated with
2001; Lindsley, 1991; Ramdohr, 1955; Righter et al., 2006a). Vanadium, ore deposits (Barnes, 1997; Otake et al., 2010).
which has a range of possible oxidation states depending on prevailing
oxygen fugacity conditions, is mainly present as V3 + in 1.2. Magnetite as a petrogenetic indicator
titanomagnetites but commonly shows variable minor amounts of
V4+ (Balan et al., 2006; Bordage et al., 2011; Toplis and Carroll, 1995). Magnetite is one of the most abundant oxide minerals in the conti-
Vanadium is incompatible at high oxygen fugacity levels due to its 5+ nental crust and has been recognized as an important indicator mineral
oxidation state. The strong oxygen fugacity-dependence of V can be for petrogenetic and geochemical studies since the early 20th century

Fig. 2. (A) Log fO2–T diagram showing relevant buffers for the Fe–Si–O system. HM: hematite–magnetite, FMQ: fayalite–magnetite–quartz, MW: magnetite–wüstite, IW: iron–wüstite, IM:
iron-magnetite, and QIF: quartz–iron–fayalite. (B) Schematic phase diagram for the system Fe–O–S in fO2–fS2 space. The geometry of the diagram remains essentially the same even
though fO2 and fS2 values vary significantly with temperature.
Panel A is redrawn after Frost (1991a) and Frost et al. (1988). Panel B is redrawn after Hall (1986) and Nadoll and Mauk (2011).
4 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

(Ramdohr, 1926). Indicator minerals such as magnetite are generally Oxide–silicate, oxide–oxide, and intra-oxide re-equilibration reactions
more resistant to weathering and transport than other coexisting min- are further important controls of the magnetite stability during cooling
eral phases (McClenaghan, 2005). Magnetite is a widespread and easily of plutonic and metamorphic rocks (e.g., Frost, 1991a, b, c; Frost and
identifiable accessory mineral in igneous, sedimentary, and metamor- Lindsley, 1991; Nadoll et al., 2012; Wones, 1989).
phic host rocks of a wide range of compositions (e.g., Annersten, Recent studies put further constraints on compositional trends of
1968; Buddington and Lindsley, 1964; Prins, 1972; Ramdohr, 1955; igneous magnetite and reinforced its petrogenetic significance. For
Rumble, 1976b; Shcheka et al., 1980; Vincent and Phillips, 1954) and example, the significance of magnetite from massive sulfide Ni–Cu–
can incorporate a large number of foreign cations (Bowles et al., 2011; PGE deposits as a sensitive gauge for fractionation processes in sulfide
Lindsley, 1976a; Wechsler et al., 1984) (Fig. 1). In combination with iso- melts has been investigated by Dare et al. (2012). They found that
topic studies, compositional patterns in magnetite can provide impor- minor and trace element concentrations in magnetite are governed by
tant constraints on petrogenetic factors such as temperature, pressure, the successive depletion of lithophile elements in the sulfide melt and
and oxygen/sulfur fugacity (Anderson et al., 2008; Chiba et al., 1989; the uptake of siderophile elements such as Co, Mo, Ni, Pb, Sn, and Zn
Frost, 1991b; Ghiorso and Sack, 1991b). Magnetite forms ideal pairs in co-crystallizing sulfides. Jenner et al. (2010) recently emphasized
with silicates, carbonates, and other oxides such as hematite or ilmenite the significant role that the onset of magnetite crystallization plays for
for geothermometry and geobarometry due to its distinct oxygen iso- the saturation of sulfides in a melt. Righter et al. (2006a) presented
tope fractionation factor (e.g., Buddington and Lindsley, 1964; new values for Ni, Co, and V partition coefficients between Cr-rich
Buddington et al., 1955; Chacko et al., 2001; Friedman and O'Neil, spinels and silicate melt. Igneous magnetite from carbonatites has
1977; Ghiorso and Sack, 1991a; Powell and Powell, 1977; Sauerzapf been investigated by a number of authors who found indicative minor
et al., 2008; Zheng and Simon, 1991). By virtue of its magnetic proper- and trace element signatures that can be attributed to controlling fac-
ties, magnetite has also been the focus of many geophysical studies tors such as oxygen fugacity and interdependencies among substituting
that investigated its use for geophysical mapping and exploration cations (Bagdasarov, 1989; Bailey and Kearns, 2002; Reguir et al., 2008).
(e.g., Akimoto, 1955; Behn et al., 2001; Clark, 2001; Grant, 1984/85b; Reguir et al. (2008) observed V and Mn concentrations to be directly
Malmqvist and Parasnis, 1972; McEnroe et al., 2001). controlled by oxygen fugacity. Furthermore, they present data that Zn
Minor and trace elements that have been reported in magnetite in igneous magnetite from carbonatites shows a covariation with Mn,
ranges from alkali and alkaline earth metals to transition metals, includ- suggesting a link between the two elements controlled by coupled
ing REEs, and metalloid and non-metals (e.g., Borisenko and Lapin, substitution.
1972; Dare et al., 2012; Dupuis and Beaudoin, 2011; Klemm et al., Ryabchikov and Kogarko (2006) constructed fO2–T diagrams for the
1985; Lindsley, 1991; McQueen and Cross, 1998; Nadoll et al., 2012; Russian Khibina magmatic system using Fe–Ti-oxide equilibria. Varying
Nielsen and Beard, 2000; Righter et al., 2006a; Singoyi et al., 2006). Mg, Al, Mn, Ti, and Cr concentrations in magnetite have been linked
However, a suite of elements, namely Mg, Al, Ti, V, Co, Ni, Zn, Cr, Mn, with the different intrusive stages at the Sokli complex in northeastern
Ga and Sn are commonly present in magnetite of all origins at concen- Finland by Lee et al. (2005). Laser ablation inductively coupled plasma
trations that can be detected by electron microprobe analysis (10 to mass spectrometry (LA-ICP-MS) trace element analyses of magnetite
N1000 ppm). We refer to these cations as magnetite elements. from Kiruna, Sweden, showed that V and Mn are the most abundant
Concentrations of foreign cations in magnetite are responsive to a trace elements in igneous magnetite, whereas Cu, Zn, and Pb are not
number of external factors that reflect formation conditions of the re- commonly incorporated (Müller et al., 2003).
spective host rock. Comprehensive overviews of physicochemical prop- The identification of characteristic minor and trace element signa-
erties of magnetite can be found in Bowles et al. (2011) and the Reviews tures in magnetite of predominantly hydrothermal origin commenced
in Mineralogy volumes on oxide minerals (Lindsley, 1991; Rumble, with the pioneering work of Borisenko and Lapin (1971, 1972) and
1976a). Many classic studies that focused on igneous magnetite show Shcheka et al. (1988), who were also among the first to apply multivar-
extensive compositional variability in response to seven main control- iate statistics to magnetite minor and trace element data in order to re-
ling factors: (1) source rock or magma composition, (2) temperature, veal underlying compositional trends that could be used to discriminate
(3) pressure, (4) cooling rate, (5) oxygen fugacity, fO2, (6) sulfur fugac- sample populations. Recent studies have presented data that suggest
ity, fS2, and (7) silica and sulfide activity (Buddington and Lindsley, that magnetite from magmatic–hydrothermal deposits displays sys-
1964; Frost and Lindsley, 1991; Ghiorso and Sack, 1991a; Haggerty, tematic variations in minor and trace element concentrations that can
1991a; Mollo et al., 2013; Whalen and Chappel, 1988). Additionally, be used to fingerprint ore deposits (Beaudoin and Dupuis, 2009;
crystallographic factors such as ionic radius and overall charge balance Dupuis and Beaudoin, 2011; Kamvong et al., 2007; Nadoll et al., 2012;
are important aspects that put decisive constraints on possible Rusk et al., 2009; Singoyi et al., 2006). Carew et al. (2006) and Rusk
substituting cations (Cornell and Schwertmann, 2003; Fleet, 1981; et al. (2010) presented evidence that Ti, V, and Mn concentrations in hy-
Goldschmidt, 1954; Wechsler et al., 1984) (Fig. 1). Magnetite is also drothermal magnetite from the Ernest Henry iron oxide copper gold
an important accessory mineral in metamorphic rocks. The composition (IOCG) deposit in Australia's Cloncurry region can be used to discrimi-
of metamorphic magnetite changes in response to mainly two factors: nate barren from mineralized host rocks. Singoyi et al. (2006) observed
temperature and oxygen fugacity (Frost, 1991c). The partitioning be- three distinct types of magnetite (referred to as types A, B, and C) from
havior and distribution of trace elements in magnetite depends on the selected volcanic-hosted massive sulfide (VMS), skarn, IOCG, and
metamorphic grade (Evans and Frost, 1975; Skublov and Drugova, Broken Hill-type clastic-dominated Pb–Zn deposits in Australia. Type
2003; Van Baalen, 1993). Low-grade metamorphic magnetite is compo- A commonly and consistently incorporates Mg, Al, Ti, V, Mn, Co, Ni,
sitionally homogeneous and has very low trace element concentrations Zn, Ga and Sn at levels above LA-ICP-MS detection limits, whereas
compared to most other types of hydrothermal magnetite (Frost, 1991c; type B displays lesser concentrations and more heterogeneous distribu-
Nadoll et al., 2012). This has been interpreted as a reflection of the low tion across different samples for Cr, As, Zr, Nb, Mo, REE, Ta, W and Pb.
formation temperatures (Nadoll et al., 2012). Although elements such Their type C magnetite hosts no detectable concentrations of Cu, Ag,
as Ti are essentially immobile under low temperatures, Van Baalen Se, Tl, Te, Bi and Au. Singoyi et al. (2006) further suggest that Sn/Ga
(1993) showed that Ti can be mobilized on a length scale of meters and Al/Co ratios can help to distinguish magnetite from VMS, skarn,
even in low-grade metamorphic rocks. Verlaguet et al. (2006) came to IOCG, and Broken Hill-type clastic-dominated Pb–Zn deposits; an ap-
similar conclusions for the comparably immobile element Al. Further- proach that was later adopted by Kamvong et al. (2007). McQueen
more, Al contents in spinel minerals from metamorphic rocks contain- and Cross (1998) demonstrated how trace element variations in hydro-
ing silicates are governed by pressure-, temperature-, and water thermal magnetite hosted in contact metasomatic skarn deposits can
fugacity-sensitive equilibria involving chlorite (Evans and Frost, 1975). provide characteristic signatures and consequently, help to target
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 5

magnetite-bearing ore deposits. Most recently, Duan et al. (2012) variations in Ti, Mn, Cr, V, Ni, Co, Zr, Sn, Zn, Pb, and Cu concentrations
documented systematic compositional variations among different in detrital magnetite to trace the source rocks of sediments, comparing
types of hydrothermal magnetite, tracing the evolution of the magmatic– them to a set of standard igneous magnetites first described by
hydrothermal mineralization in the Washan porphyry-style Fe deposit, Shcheka et al. (1980). Similarly, Grigsby (1990) used petrographic and
East China. Huang et al. (2013) described the compositions of hydro- chemical variations in detrital magnetite to discriminate among felsic, in-
thermal magnetite from the magmatic–hydrothermal Cihai Fe deposit termediate, and mafic volcanic and plutonic source rocks.
in China. They compared their results with the magnetite classification
scheme suggested by Dupuis and Beaudoin (2011) and showed that 1.4. Defining hydrothermal magnetite
magnetite from the Cihai deposit is depleted in V and Ti compared to
magnetite from Fe–Ti–V deposits and has Mg, Al, Ti, V, Cr, Co, Ni, Mn, The term hydrothermal magnetite can be rather ambiguous espe-
Zn, Ga, and Sn concentrations that exceed those commonly found in cially when considering the complex geological and mineralogical
magnetite from skarn deposits. relationships found in many hydrothermal ore deposits (Fig. 3). Mul-
tiple vein generations, large- to small-scale overprints of multiple al-
1.3. Magnetite in exploration and provenance studies teration stages, fluid/rock interaction, and secondary weathering
processes translate into a variety of different types of hydrothermal
Mining and exploration companies have recognized the great magnetite. These are commonly synonymously addressed in publi-
potential of employing the geochemistry of magnetite as an explora- cations. However, it is important to consider the different types of
tion tool for many years. There is a large body of unpublished work hydrothermal magnetite when attributing specific minor and trace
that investigated the use of magnetite as a pathfinder or indicator min- element concentrations to a specific deposit or type of mineraliza-
eral. One of the main challenges in mineral exploration is to detect geo- tion. One of the main challenges for researchers that investigate hy-
chemical signatures of ore deposits at greater and greater distances drothermal magnetite is to diligently record the large variety of
from the main mineralization. Magnetite, among other detrital heavy minor and trace element compositions in magnetite that reflect a di-
minerals such as ilmenite, rutile, garnet, and zircon has been a valuable verse formation and alteration history, and at the same time invoke
asset to mineral exploration and provenance studies (e.g., Dupuis and diagnostic trends that can be used to reliably discriminate different
Beaudoin, 2011; Nadoll, 2011; Nadoll et al., 2012; Razjigaeva and ore deposits, as well as barren zones from mineralized zones.
Naumova, 1992; Shcheka et al., 1982). Hydrothermal and igneous mag-
netite can provide vectors to mineralized areas where the primary min- 2. Methods
eralogical context is lost, deeply covered, or highly altered, similar to the
use of rutile as a resistate indicator mineral (Anand and Butt, 2010; 2.1. Electron microprobe
Borisenko and Lapin, 1972; Buddington and Lindsley, 1964; Hutton,
1950; McQueen and Cross, 1998; McQueen and Whitbread, 2002; Electron microprobe analysis (EMPA) were carried out at North
Rumble, 1976a and references therein, Shcheka et al., 1982; Triebold Ryde (Sydney) on a Cameca Camebax™ automated wave length disper-
et al., 2007; Zack et al., 2004). Razjigaeva and Naumova (1992) used sive spectroscopy scanning electron microprobe with an acceleration

Fig. 3. The complexity of hydrothermal ore deposits translates into a variety of different types of hydrothermal 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 magnetite in terms
of process-controlled compositional variability, it is not only important to be able to discriminate 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 hy-
drothermal deposit types.
6 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

voltage of 15 kV and a beam current of 20 nA. The analytical conditions Mg24, Al27, Si28, S34, Ca44, Ti47, V51, Cr52, Mn55, Fe57, Co59, Ni60,
for each element are given in Table 1. Line overlap corrections were per- Cu63, Zn66, Ga71, Ge72, As75, Se82, Sr88, Y89, Nb93, Mo98, Ag107,
formed for the overlap of Ti Kβ on VKα, V Kβ on CrKα, Cr Kβ on MnKα In115, Sn118, Sb121, Ba138, Ce140, W184, Au197, Tl205, Pb208 and
and Mn Kβ on FeKα. ZAF corrections were applied for all elements. Mi- Th232. This list includes a number of non-spinel elements, such as Si,
croprobe results were accepted only if the compound weight percent Ca, and the REEs, which can be helpful for data screening and the iden-
total was between 98.5 and 101.5%. The ferric iron content of each anal- tification of analyses that are affected by inclusions. Laser ablation data
ysis is calculated based on the assumption of stoichiometry and an ideal were processed by non-automatic spectra reduction to exclude spectra
AB2O4 formula. Tin and Ga were not measured. or parts of spectra that are affected by mineral inclusions.
Back-scattered electron (BSE) and multi-element imaging with a
field emission gun scanning electron microscope (FEG-SEM) was un-
2.3. Data treatment
dertaken at CSIRO Earth Sciences and Resource Engineering, Perth,
with an acceleration voltage of 20 keV and a beam current of 2.9 nA.
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
2.2. Laser ablation inductively coupled plasma mass spectrometry
conditions, translate into a wide range of detection limits. The lowest
comparable element concentrations are unavoidably defined by the
The growing number of studies that employ LA-ICP-MS as the
greatest detection limit for the respective element when comparing
preferred analytical technique to collect ppm to sub-ppm trace ele-
minor and trace element concentrations across different datasets.
ment concentrations in magnetite have produced a body of evidence
Several statistical methods can be used for data analysis; we show box
that a matrix matched external standard is not required and that for
and whisker plots that display lower quartile, median, mean, upper
most conditions and spot sizes reference materials such as GSE-1G,
quartile, and the 5th and 95th percentile values, as well as discriminate
GRM278, BCR2-G, NIST610, and NIST612 are well suited and produce
outliers. The upper threshold values, representing the upper 95th per-
reliable results (Dare et al., 2012; Nadoll and Koenig, 2011; Savard
centile of the data, are usually not affected by the varying detection
et al., 2010). Detection limits at ppm to ppb levels shed light on subtle
limits and therefore represent a robust approach to compare data
but characteristic compositional variations for elements of interest that
from different sources. To eliminate detection limit related minor and
have formerly been hampered by greater detection limits or limitations
trace element variations, element ratios have also been employed for
to the number of elements measured.
data comparison.
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 3. Geologic background
data. The ferric iron content of each analysis was calculated based on
the assumption of stoichiometry and an ideal AB2O4 formula. This section provides a geologic overview of the investigated areas
Table 2 shows the average laser ablation ICP-MS reporting limits and deposits. Locality maps are shown in Figs. 4 and 5. A complete list
(USGS) and detection limits (UQAC, CODES — Angerer et al., 2012, of samples including host rock and alteration types are shown in
2013). Thirty-three elements were included in the analyses: Na23, Table 3.

Table 1
Laser ablation ICP-MS operation conditions for the Université du Québec à Chicoutimi (UQAC), U.S. Geological Survey Denver (USGS), and CODES University of Tasmania.

LA-ICP-MS UQAC USGS CODES

Location Université Chicoutimi, US Geological Survey, Denver, USA ARC Centre of Excellence in Ore Deposits,
Québec, Canada University of Tasmania, Australia
Laser ablation system RESOlution M-50 (Excimer 193 nm) New Wave Research UP-193 FX New Wave Research UP-213 (Excimer 213 nm)
(Excimer 193 nm)
ICP-MS Agilent 7700× Quadrupole ICP-MS PerkinElmer DRC-e ICP Agilent HP4500 Quadrupole ICP-MS
Pulse frequency 5 Hz 4 Hz 5 Hz
Scan type Line Spot Spot
Stage speed 4 μm/s n/a n/a
Beam size 15–33 μm 15–100 μm 15–49 μm
Energy density 5 J/cm2 ~5 J/cm2 ~3.3 J/cm2
Analysis line raster ~240 μm long n/a n/a
Gas blank 30 s 45–60 s 30 s
Signal 60 s 120 s 60 s
Internal standard Fe (EMPA) Fe (EMPA) Fe (EMPA)
Reference material GSE-1G, BC-28 GSE-1G, NIST-610 STDGLb2, NIST-610

EMPA — CSIRO, North Ryde, Australia

Element (X-ray line) Standard Analyzing crystal

Mg (Kα) Spinel TAP


Al (Kα) Spinel TAP
Ti (Kα) Rutile PET
V (Kα) V metal LiF
Cr (Kα) Cr metal LiF
Mn (Kα) Mn metal LiF
Fe (Kα) Hematite LiF
Co (Kα) Co metal LiF
Ni (Kα) Ni metal LiF
Zn (Kα) Zn metal LiF
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 7

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

Laboratory North Ryde USGSa UQACb CODESb

Spot size b2 μm 100 μm 50 μm 25 μm 15 μm 10 μm 33 μm 25 μm 47 μm 35 μm 22 μm 15 μm

Na 2.98 14 48 148 394


Mg 218 1.97 10 48 166 452 0.04 0.06 0.1 0.1 0.2 0.3
Al 494 12.1 60 219 561 1240 0.18 0.29 0.5 0.6 0.7 0.9
Si 933 2790 8550 25,400 47,800 2000 3590
S 247 215 234 342 873
Ca 4060 10,100 26,900 64,400 145,000 7.26 13.6 398 472 1210 927
Ti 330 9.13 15 75 160 330 0.04 0.08 0.2 0.2 0.4 0.5
V 406 1.51 14 35 70 170 0.03 0.05 0.2 0.2 0.4 0.5
Cr 269 7.56 36 120 920 2460 0.63 1.09 3.53 3.08 7.22 8.02
Mn 336 1.34 2 10 29 70 0.56 0.95 0.81 0.96 1.24 1.54
Co 350 0.38 2 8 20 50 0.01 0.01 0.07 0.08 0.09 0.2
Ni 300 1.48 12 47 133 330 0.03 0.05 0.41 0.5 1.07 0.4
Cu 0.95 3 11 30 70 0.02 0.03 1.17 1.09 10.63 3.69
Zn 563 3.37 14 60 190 370 0.1 0.17 1.31 2.51 10.31 8.15
Ga 0.43 1 6 15 32 0.04 0.07 0.12 0.19 0.07 0.10
Ge b1 4 15 56 130
As 1.97 10 34 106 320 0.09 0.16 1.50 4.88 11.97 15.97
Se 21.2 108 224 1720 505
Sr 0.13 1 2 9 19 0.01 0.10 0.04 0.04 0.12 0.11
Y b1 1 2 7 17 0.07 0.06 0.01 0.02 0.09 0.07
Nb 0.08 0.5 2 7 15 n/a n/a 0.02 0.05 0.1 0.2
Mo 0.24 2 8 22 57 0.01 0.02 0.2 0.22 0.49 0.65
Ag 0.41 2 8 23 50
In 0.09 0.4 1 4 11
Sn 1.21 2 7 23 51 0.03 0.05
Sb 0.28 2 6 19 46
Ba 0.07 1 3 8 18 0.44 0.84 0.09 0.21 0.54 0.74
Ce 0.05 0.5 2 6 19 0.06 0.11 0.01 0.17 0.06 0.06
Au 0.29 1 5 107 63
Tl 0.1 1 5 56 53
W 0.17 0.3 0.03 0.08 0.21 0.2
Pb 0.51 2 5 19 35 0.01 0.01 0.02 0.04 0.06 0.19
Th 0.08 1 3 9 22 0.06 0.1 0.02 0.06 0.04 0.06
a
10× standard deviation.
b
3× standard deviation.

3.1. Banded iron formations and related iron ore deposits 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-
Banded iron formations (BIFs) are the most important sedimen- ore) has several forms: (1) laminated bands made of amalgamated,
tary host rock for iron ore. BIFs have been deposited predominantly anhedral to subhedral grains (Fig. 6A), (2) disseminated grains with-
during Eoarchean to Paleoproterozoic times, and are preserved in in, but obscuring BIF laminae (Fig. 6B), and (3) layer-discordant
two main tectonostratigraphic settings — volcanic rock dominated veins, shear zones and cleavage filling. Magnetite grain sizes com-
granite–greenstone terranes (Algoma-type BIF) or continental shelf monly increase with metamorphic grade without modifying textural
environments (Superior-type BIF). Despite their difference in strati- equilibrium relationships with neighboring quartz (Klein, 2005). The
graphic successions, both Algoma- and Superior-type BIF are mineral- presented data set includes 158 LA-ICP-MS analyses of metamorphic
ogically similar, consisting of iron oxide, siderite to ankerite–dolomite, and hydrothermally altered magnetite from various major iron ore
or iron silicate laminae interlayered with chert/quartz laminae districts.
(Fig. 6A) (Klein, 2005; Trendall and Blockley, 1970). Both Algoma-
and Superior-type BIF locally host high-grade iron ore bodies of signifi-
cant economic size, in which iron underwent an enrichment process by 3.1.1. Meso-, and Neoarchean Algoma-type BIF
post-depositional alteration processes. The processes and geological The data set includes 57 LA-ICP-MS analyses representing magnetite
conditions for enrichment of iron in BIF can vary from district to district, from Mesoarchean Algoma-type BIF from the Yilgarn craton, Western
and several genetic models have been proposed (Angerer and Australia: Koolyanobbing and Marda greenstone belts (Angerer et al.,
Hagemann, 2010; Clout, 2003; Evans et al., 2013; Gruner, 1937; 2012, 2013), and Weld Range (previously unpublished data, kindly pro-
Lobato et al., 2008; Morris, 1985; Ramanaidou, 2009; Rosière et al., vided by Paul Duuring, The University of Western Australia) (Fig. 4A).
2008; Taylor et al., 2001). Unaltered and unweathered BIF typically con- The sample set consists of metamorphic and hydrothermally altered
tains diagenetic or metamorphic magnetite and hematite, and a domi- magnetite. Peak burial metamorphic temperatures are recorded to
nance of one over the other reflects the metamorphic grade and range between 200 and 350 °C (Angerer et al., 2012, 2013; Duuring
oxygen fugacity in the rock (Klein, 2005). Magnetite is only a minor and Hagemann, 2013). Temperatures of the various hydrothermal over-
constituent in the most economic zones of many BIF-hosted iron ore de- prints are not well constraint, but isotopic equilibriums calculations
posits, which is due to pervasive oxidation and replacement by other (Angerer, unpublished data) suggest generally lower temperatures
minerals. However, magnetite occurs in deeper sections of ore deposits, than peak temperature (200 to 300 °C). Twenty-one analyses are from
below the weathering front, where it commonly forms medium- to the Mesoarchean Iron Ore Group of Jharkhand-Orissa region, India
high-grade “proto-ore”, and in these zones complex mineral associa- (Bhattacharya et al., 2007) (Fig. 4C). Peak metamorphic overprint is
tions, including iron-rich carbonates and/or silicates are commonly ob- reported as lower greenschist facies, i.e., ~ 270–350 °C (Bhattacharya
served (Angerer et al., 2012; Duuring and Hagemann, 2012; Figueiredo et al., 2007).
8 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

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., 2012, 2013), 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 Jharkhand–Orissa region, India (Bhattacharya et al., 2007).

3.1.2. Neoarchean–Paleoproterozoic Superior-type BIF with peak metamorphic conditions in the area (Doughty and
Neoarchean to Paleoproterozoic Superior-type BIF is represented Chamberlain, 1996; Nadoll et al., 2012; Sack and Lichtner, 2009). Sider-
by 39 magnetite analyses from the Marra Mamba and Mt Sylvia ite and calcite are the predominant gangue minerals in mineralized
Formations in the Hamersley Province, Western Australia (Angerer veins. Hematite and magnetite are common accessory minerals in
et al., unpublished data), and three unpublished analyses from the the host rocks, but they can also occur as rare accessory minerals in
Caué Formation, Quadrilátero Ferrífero (QF), Brazil (kindly provided hydrothermal veins (Fryklund, 1964; Nadoll et al., 2012). Our database
by Ana-Sophie Hensler, The University of Western Australia). Peak includes 30 previously published LA-ICP-MS analyses of magnetite
burial metamorphic temperatures range between 200 and 300 °C in (Nadoll et al., 2012), from the Gold Hunter calcite and siderite veins
the Hamersley Province (Klein, 2005), and 250 to 350 °C in the QF (Fig. 6C) of the Lucky Friday mine.
(Rosière and Rios, 2004). Hydrothermally altered magnetite from vari-
ous alteration zones in the Tom Price high-grade iron ore deposit is rep- 3.3. Porphyry Cu–(Au)–(Mo) and related skarn deposits
resented by 38 previously unpublished magnetite analyses (kindly
provided by Warren Thorne, The University of Western Australia) The geology and magmatic–hydrothermal processes involved in the
(Fig. 4B). The hydrothermal fluids producing magnetite in the Tom formation of porphyry and associated skarn deposits have been de-
Price deposit were basinal brines (107–142 °C, 25.5 wt.% equiv; NaCl– scribed by many authors (e.g., Audetat et al., 2008; Einaudi, 1981b;
CaCl2-rich) (Thorne et al., 2004). Seedorff et al., 2005; Sillitoe, 2010; Titley, 1981; Titley and Beane,
1981). Multiple stages of mineralization, alteration and in most cases
3.2. Ag–Pb–Zn veins (Lucky Friday, Coeur d'Alene district, Idaho) a diverse suite of plutonic, volcanic and (meta)-sedimentary host
rocks are common characteristics of porphyry and related skarn de-
The Lucky Friday mine is located in the Coeur d'Alene mining district posits (e.g., Beane and Titley, 1981; Einaudi, 1981b; Halsall, 1983;
of northern Idaho, which is one of the major Ag–Pb–Zn producers in the Lowell and Guilbert, 1970; Seedorff et al., 2005; Sillitoe, 2010; Titley,
world. The Ag–Pb–Zn replacement veins formed at 250–350 °C and 1981). Hydrothermal veins form at temperatures between 300 and
1–3 kbar in clastic burial metamorphic rocks of the Mesoproterozoic 800 °C (e.g., Beane and Titley, 1981; Einaudi, 1981a; Muntean and
Belt Supergroup (Beaudoin and Sangster, 1992; Landis and Hofstra, Einaudi, 2001; Rusk et al., 2008; Seedorff et al., 2005; Sillitoe, 2000;
1991; Leach et al., 1988). Vein formation temperatures correspond Sillitoe, 2010; Ulrich et al., 2001). Skarn deposits can form where
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 9

Fig. 5. Locations and schematic geology for (A) porphyry Cu (Mo) and associated skarn deposits in the southwestern United States (Garrity and Soller, 2009; Nadoll 2011), (B) the
Henderson Climax-type Mo deposit in the central United States (Garrity and Soller, 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.

porphyry systems intrude carbonate rocks (e.g., Beane and Titley, 1981; 2005). To assure that hydrothermal and igneous magnetite are properly
Einaudi, 1981b; Einaudi et al., 1981; Meinert et al., 2005). Magnetite of identified in the complex porphyry and skarn environment, a combina-
igneous and hydrothermal origin is a ubiquitous accessory mineral in tion of field observations and detailed microscopy as well as careful
porphyry Cu–(Au)–(Mo) and associated skarn deposits. Magnetite mineral separation has been undertaken.
commonly forms prior to and during the early potassic alteration
(Arancibia and Clark, 1996), but can also be found in later stage 3.3.1. American Southwest (Arizona, New Mexico)
magnetite-rich alteration as disseminated accessory minerals or in The Laramide Belt in the southwestern United States hosts porphyry
veins of varying assemblages (e.g., Beane and Titley, 1981; Gustafson and skarn deposits of different ages (Jurassic, Cretaceous, and Tertiary),
and Hunt, 1975; Sillitoe, 2000, 2010; Titley and Beane, 1981). In skarn dimensions, and metal content (Cu, Mo, Ag, Au) (e.g., Ahmad and Rose,
deposits, hydrothermal magnetite can be an abundant mineral phase, 1980; Enders, 2000; Halsall, 1983; McLemore et al., 1999; Robinson and
occurring as massive replacement mineralization as a result of Fe-rich Cook, 1966; Rose and Baltosser, 1966; Seedorff et al., 2005; Sillitoe,
fluids (e.g., Einaudi, 1981b; McQueen and Cross, 1998; Meinert et al., 2010; Titley, 1981; Titley and Beane, 1981). Formation temperatures
10
Table 3
List of samples including information about domain, magnetite type, host rock, alteration, location, and type and number (#) of analysis (LA-ICP-MS or EMPA).

Sample # Setting Location Domain Type Host rock Alteration DH — Mine Easting Northing Depth/ Data source
Level elevation

6499220 6 BIF-Iron ore Tom Price deposit Hamersley Province Metamorphic– Superior type BIF, Carbonate–talc– n/a n/a n/a n/a Previously unpublished
hydrothermal Brockman Fma chlorite magnetite (W. Thorne, UWA)
6499240 7 BIF-Iron ore Tom Price deposit Hamersley Province Metamorphic– Superior type BIF, Carbonate–talc– n/a n/a n/a n/a Previously unpublished
hydrothermal Brockman Fma chlorite magnetite (W. Thorne, UWA)
NTD9113 9 BIF-Iron ore Tom Price deposit Hamersley Province Metamorphic– Superior type BIF, Siderite–silicate n/a n/a n/a n/a Previously unpublished
hydrothermal Brockman Fma magnetite (W. Thorne, UWA)
6499320 5 BIF-Iron ore Tom Price deposit Hamersley Province Metamorphic– Superior type BIF, Siderite–silicate n/a n/a n/a n/a Previously unpublished
hydrothermal Brockman Fma magnetite (W. Thorne, UWA)
NTD1173 11 BIF-Iron ore Tom Price deposit Hamersley Province Metamorphic– Superior type BIF, Siderite–silicate n/a n/a n/a n/a Previously unpublished
hydrothermal Brockman Fma magnetite (W. Thorne, UWA)
HH188-208.8 9 BIF-Iron ore Hashimoto Hamersley Province Metamorphic Superior type BIF, Least-altered n/a n/a n/a n/a Previously unpublished
deposit Mt Sylvia Fma (T. Angerer, UWA)
WHB1113-232.7 4 BIF-Iron ore Mt Whaleback Hamersley Province Metamorphic Superior type BIF, Magnetite– n/a n/a n/a n/a Previously unpublished
deposit Mt Sylvia Fma martite (T. Angerer, UWA)
WHB1113-307.2 6 BIF-Iron ore Mt Whaleback Hamersley Province Metamorphic Superior type BIF, Magnetite– n/a n/a n/a n/a Previously unpublished
deposit Mt Sylvia Fma martite (T. Angerer, UWA)
MG368-183.6 5 BIF-Iron ore Mesa Gap deposit Hamersley Province Metamorphic Superior type BIF, Least-altered n/a n/a n/a 183.6 Previously unpublished

P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32


Marra Mamba Fmb (T. Angerer, UWA)
MG368-181 8 BIF-Iron ore Mesa Gap deposit Hamersley Province Metamorphic Superior type BIF, Least-altered n/a n/a n/a 181 Previously unpublished
Marra Mamba Fmb (T. Angerer, UWA)
77D153a 8 BIF-Iron ore B deposit Hamersley Province Metamorphic Superior type BIF, Least-altered n/a n/a n/a n/a Previously unpublished
Marra Mamba Fmb (Curtin University)
FSD 122-137.85 3 BIF-Iron ore Cerro do Curral Iron Quadrangle Metamorphic Superior type BIF, Hydrothermal- n/a n/a n/a n/a Previously unpublished
(supergene- Caue Fmb supergene (A.-S. Hensler UWA)
altered)
bn-w10-13_ 1 BIF-Iron ore Weld Range, Yilgarn Craton Metamorphic Algoma type BIFc Dolomite-altered WRRC0253d 582,437 7,027,013 111 Previously unpublished
K1-7b Beebyn deposit magnetite (P. Duuring, UWA)
bn-w9-101_k1 5 BIF-Iron ore Weld Range, Yilgarn Craton Metamorphic Algoma type BIFc Least-altered WRRC1076d 580,196 7,026,154 152.7 Previously unpublished
Beebyn deposit (P. Duuring, UWA)
A-1 9 BIF-Iron ore Koolyanobbing Yilgarn Craton Metamorphic– Algoma type BIFc Siderite-altered Surface sample n/a n/a 70 Angerer et al. (2012)
A deposit hydrothermal
c
KK17-31 4 BIF-Iron ore Koolyanobbing Yilgarn Craton Metamorphic– Algoma type BIF Talc-altered KPDDH017 n/a n/a 182.75 Angerer et al. (2012)
K deposit hydrothermal magnetite
K901 5 BIF-Iron ore Koolyanobbing Yilgarn Craton Metamorphic– Algoma type BIFc Talc-altered Surface sample n/a n/a 270 Angerer et al. (2012)
K deposit hydrothermal magnetite
K910 3 BIF-Iron ore Koolyanobbing Yilgarn Craton Metamorphic– Algoma type BIFc Talc-altered Surface sample n/a n/a 270 Angerer et al. (2012)
K deposit hydrothermal magnetite
c
KP-50 6 BIF-Iron ore Koolyanobbing Yilgarn Craton Metamorphic– Algoma type BIF Hydrothermal- Surface sample n/a n/a 100 Angerer et al. (2012)
K deposit hydrothermal- supergene
supergene
KK17-31 2 BIF-Iron ore Koolyanobbing Yilgarn Craton Metamorphic Algoma type BIFc Least-altered KPDDH017 n/a n/a 182.75 Angerer et al. (2012)
K deposit
KK17-17/18 3 BIF-Iron ore Koolyanobbing Yilgarn Craton Metamorphic Algoma type BIFc Least-altered KPDDH017 n/a n/a 165.3 Angerer et al. (2012)
K deposit
c
K110 10 BIF-Iron ore Koolyanobbing Yilgarn Craton Metamorphic- Algoma type BIF Magnetite- Surface sample n/a n/a 270 Angerer et al. (2012)
K deposit contactmetamorphic hematite ore
c
W2-331.7 6 BIF-Iron ore Windarling W2 Yilgarn Craton Metamorphic– Algoma type BIF Dolomite-altered W2DDH007 n/a n/a 331.7 Angerer et al. (2013)
deposit hydrothermal magnetite
W2-343.4 2 BIF-Iron ore Windarling W2 Yilgarn Craton Metamorphic– Algoma type BIFc Dolomite-altered W2DDH008 n/a n/a 343.4 Angerer et al. (2013)
deposit hydrothermal magnetite
W2-377.1 2 BIF-Iron ore Windarling W2 Yilgarn Craton Metamorphic– Algoma type BIFc Hematite– W2DDH009 n/a n/a 377.1 Angerer et al. (2013)
deposit hydrothermal magnetite ore
c
W2-380.8 1 BIF-Iron ore Windarling W2 Yilgarn Craton Metamorphic– Algoma type BIF Magnetite– W2DDH010 n/a n/a 380.8 Angerer et al. (2013)
deposit hydrothermal hematite ore
c
W2-345.5 10 BIF-Iron ore Windarling W2 Yilgarn Craton Metamorphic Algoma type BIF Least-altered W2DDH011 n/a n/a 345.5 Angerer et al. (2013)
deposit
H/1 5 BIF-Iron ore Gandhamardan Metamorphic Algoma type BIFb Least-altered n/a n/a n/a n/a Bhattacharya et al. (2007)
H/2 6 BIF-Iron ore Deo Nala Metamorphic– Algoma type BIFb Altered by granitic n/a n/a n/a n/a Bhattacharya et al. (2007)
hydrothermal fluids?
H/4 5 BIF-Iron ore Garumahishani Metamorphic Algoma type BIFb Least-altered n/a n/a n/a n/a Bhattacharya et al. (2007)
H/5 5 BIF-Iron ore Garumahishani Metamorphic Algoma type BIF Least-altered n/a n/a n/a n/a Bhattacharya et al. (2007)
(silicate facies)b
563 GM6#1G3 12 IZB Japan Magnetite Igneous Granodiorite Least-altered n/a n/a n/a n/a Spong (1998), Nadoll
granodiorite series granitoid et al. (2012)
AU58431 8 Ag-Pb-Zn Lucky Friday Gold Hunter Hydrothermal Calcite vein Siderite 4900 Level 591,793 5,258,197 −1493 Nadoll et al. (2012)
AU58451 9 Ag-Pb-Zn Lucky Friday Gold Hunter Hydrothermal Siderite vein Siderite 5900 Level 18,060.79 24,482.08 −2464.87 Nadoll et al. (2012)
AU58484 4 Ag-Pb-Zn Lucky Friday Gold Hunter Hydrothermal Siderite vein Siderite 61-05 29 Nadoll et al. (2012)
AU58493 9 Ag-Pb-Zn Lucky Friday Gold Hunter Hydrothermal Siderite/siltite Siderite 61-05 421.5 Nadoll et al. (2012)
AU60865 8 Mg-Skarn Chino (Santa Rita) Hydrothermal Magnesian skarn Skarn 1991 1701d 606d 1330 Nadoll (2011)
AU60869 9 Mg-Skarn Chino (Santa Rita) Hydrothermal Magnesian skarn Skarn 1991 1701d 606d 1454 Nadoll (2011)
AU60908 7 Porphyry Chino (Santa Rita) Hydrothermal Granodiorite Potassic–sericitic D2245 30 Nadoll (2011)
AU60916 6 Porphyry Chino (Santa Rita) Hydrothermal Quartzite Calcic sodic 2305 2478.1d 867.4d 2148 Nadoll (2011)
AU60895 8 Porphyry Cobre Hydrothermal Hornfels Propylitic 772,378 3,637,481 Nadoll (2011)
AU61017 6 Porphyry Morenci Hydrothermal Diabase Argillic 3898 −8000d 9800d 1044 Nadoll (2011)
AU61018 2 Porphyry Morenci Hydrothermal Diabase Argillic 3898 −8000d 9800d Nadoll (2011)
AU61022 7 Porphyry Morenci-Producer Hydrothermal Diorite Sericitic 3925 −17170d 11750d 240 Nadoll (2011)
AU61011 8 Porphyry Morenci-Sun Ridge Hydrothermal Granite Sericitic 3944 −16200d 21600d 1502 Nadoll (2011)
AU61015 7 Porphyry Morenci-Sun Ridge Hydrothermal Granite Sericitic 3944 −16200d 21600d 1555 Nadoll (2011)

P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32


AU60967 8 Porphyry Safford-Dos Hydrothermal Andesite Propylitic RL-1 20437d 4050d 1534 Nadoll (2011)
Pobres
AU60962 9 Porphyry Safford-Dos Hydrothermal Andesite Propylitic RL-1 20437d 4050d 1464 Nadoll (2011)
Pobres
d d
AU60976 7 Porphyry Safford-Dos Hydrothermal Granodiorite Propylitic RL-1 20437 4050 2027 Nadoll (2011)
Pobres
AU60940 2 Porphyry Safford-Lone Star Hydrothermal Andesite Potassic–argillic LS07-0153 1489 Nadoll (2011)
AU60942 8 Porphyry Safford-Lone Star Hydrothermal Andesite Potassic–argillic LS06-0083 41,201.81d 8799.35d 2173 Nadoll (2011)
AU60952 7 Porphyry Safford-Lone Star Hydrothermal Andesite Potassic–sericitic LS06-0083 41,201.81d 8799.35d 1411.5 Nadoll (2011)
oxidized
AU60947 8 Porphyry Safford-Lone Star Hydrothermal Andesite Propylitic LS07-0154 1932 Nadoll (2011)
AU60990 8 Porphyry Safford-San Juan Hydrothermal Andesite Calcic sodic SJP-64 27600d 4109d 1085 Nadoll (2011)
AU60991 10 Porphyry Safford-San Juan Hydrothermal Andesite Calcic sodic SJP-64 27600d 4109d 1141 Nadoll (2011)
AU60859 9 Skarn Chino (Santa Rita) Hydrothermal Skarn Skarn 2209 5810d 1310d 1430 Nadoll (2011)
AU60875 8 Skarn Chino (Santa Rita) Hydrothermal Skarn Skarn 2166 5400d 1000d 2152 Nadoll (2011)
AU60910 9 Skarn Chino (Santa Rita) Hydrothermal Skarn Skarn D2245 76.7 Nadoll (2011)
AU60930 8 Skarn Chino (Santa Rita) Hydrothermal Skarn Skarn 2330 −3832.9d 673.5d 1004 Nadoll (2011)
AU60898 5 Skarn Cobre Hydrothermal Skarn Skarn 772,341 3,637,425 Nadoll (2011)
AU60903 8 Skarn Cobre Hydrothermal Skarn Skarn 772,232 3,637,506 2003 Nadoll (2011)
AU60884 5 Skarn Copper Flat Hydrothermal Skarn Skarn 769,403 3,633,542 Nadoll (2011)
AU61024 7 Skarn Morenci-Shannon Hydrothermal Skarn Skarn −9959d 22266d 5300 Nadoll (2011)
onramp
AU60857 8 Porphyry Chino (Santa Rita) Igneous Granodiorite Sericitic 2209 5810d 1310d 1398 Nadoll (2011)
AU60914 8 Porphyry Chino (Santa Rita) Igneous Granodiorite Potassic 774,830 3,632,319 Nadoll (2011)
AU60919 8 Porphyry Chino (Santa Rita) Igneous Granodiorite Potassic 1689-x −115.07d 108.76d 783.5 Nadoll (2011)
breccia
d d
AU60921 8 Porphyry Chino (Santa Rita) Igneous Monzonite Potassic 1689-x −115.07 108.76 773 Nadoll (2011)
AU60893 8 Porphyry Cobre Igneous Granite Potassic 772,352 3,637,691 Nadoll (2011)
AU61033 8 Climax-Mo Henderson Igneous Leuco-rhyolite/ Hypogene NEHW 5180d 7405d 7150 Previously unpublished
granite level
AU61031 5 Climax-Mo Henderson igneous leuco-rhyolite/ Hypogene 662 957.4 Previously unpublished
granite
d d
AU61032 9 Climax-Mo Henderson Igneous Leuco-rhyolite/ Hypogene P87 5188 7442 7210 Previously unpublished
granite level
AU61030 8 Climax-Mo Henderson Igneous Leuco-rhyolite/ Hypogene U87 5030d 7164d 7270 Previously unpublished
granite level
AU61028 8 Climax-Mo Henderson Igneous Leuco-rhyolite/ Hypogene U93 4834d 7216d 7270 Previously unpublished
granite level

11
(continued on next page)
12
Table 3 (continued)
Sample # Setting Location Domain Type Host rock Alteration DH — Mine Easting Northing Depth/ Data source
Level elevation

AU61029 3 Climax-Mo Henderson Igneous Leuco-rhyolite/ Hypogene U93-1 4790d 7100d 7270 Previously unpublished
granite level
d d
AU61017 2 Porphyry Morenci Igneous Diabase Argillic 3898 −8000 9800 1044 Nadoll (2011)
AU61018 3 Porphyry Morenci Igneous Diabase Argillic 3898 −16300d 21400d Nadoll (2011)
AU61027 7 Porphyry Morenci-Garfield Igneous Diabase n/a −10365d 27015d 5150 Nadoll (2011)
onramp
AU61022 1 Porphyry Morenci-Producer Igneous Diorite Sericitic 3925 −17170d 11750d 240 Nadoll (2011)
AU61009 4 Porphyry Morenci-Sun Igneous Granite Sericitic– 3944 −16200d 21600d 1462 Nadoll (2011)
Ridge propylitic
AU60999 7 Porphyry Safford-Dos Igneous Granodiorite Potassic RL-26 20813d 4049d 1406 Nadoll (2011)
Pobres
d d
AU60957 6 Porphyry Safford-Lone Igneous Andesite Potassic- LS06-0083 41,201.81 8799.35 1434 Nadoll (2011)
Star oxidized
AU61007 4 Porphyry Safford-Lone Igneous Andesite Potassic- LS07-0140 1506 Nadoll (2011)
Star oxidized
AU60934 4 Porphyry Safford-Lone Igneous Andesite Propylitic LS07-0154 1919 Nadoll (2011)
Star
AU60936 6 Porphyry Safford-Lone Igneous Andesite Propylitic LS07-0151 1757.5 Nadoll (2011)

P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32


Star
AU60980 8 Porphyry Safford-San Juan Igneous Granodiorite Potassic– SJP-70 26742d 4066d 520 Nadoll (2011)
sericitic
AU60982 8 Porphyry Safford-San Juan Igneous Granodiorite Potassic SJP-70 26742d 4066d 720.7 Nadoll (2011)
126335 11e Porphyry Ertsberg district Karume Intrusion Hydrothermal Monzodiorite Propylitic CTS-1 735,386 9,550,242 162.5 Previously unpublished
126336 18e Porphyry Ertsberg district Karume Intrusion Hydrothermal Monzodiorite Phyllic CTS-1 735,386 9,550,242 177.2 Previously unpublished
126337 16e Porphyry Ertsberg district Karume Intrusion Hydrothermal Monzodiorite Propylitic CTS-1 735,386 9,550,242 376.9 Previously unpublished
126340 5e Porphyry Ertsberg district Karume Intrusion Hydrothermal Monzodiorite Potassic CTS-1 735,386 9,550,242 698.5 Previously unpublished
126312 7e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 736,966 9,548,885 17.05 Previously unpublished
126313 14e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 736,972 9,548,891 26.46 Previously unpublished
126314 5e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 736,976 9,548,896 30.65 Previously unpublished
126317 15e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 736,984 9,548,904 44.38 Previously unpublished
126318 17e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 737,002 9,548,923 70 Previously unpublished
126320 10e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 737,037 9,548,959 121.74 Previously unpublished
126323 19e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 737,070 9,548,993 168.46 Previously unpublished
126324 20e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 737,076 9,549,000 179.2 Previously unpublished
126325 12e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 737,080 9,549,004 183 Previously unpublished
126326 12e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 737,088 9,549,012 197.35 Previously unpublished
126327 11e Porphyry Ertsberg district Ertsberg Diorite Hydrothermal Diorite Phyllic FAS3-1 737,094 9,549,019 204.38 Previously unpublished
126328 17e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 737,098 9,549,023 212.8 Previously unpublished
126329 16e Porphyry Ertsberg district Ertsberg Diorite Hydrothermal Diorite Phyllic FAS3-1 737,101 9,549,026 215.1 Previously unpublished
126330 17e Porphyry Ertsberg district Ertsberg Diorite Hydrothermal Diorite Phyllic FAS3-1 737,104 9,549,029 221 Previously unpublished
126331 10e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn FAS3-1 737,106 9,549,032 222.3 Previously unpublished
126341 11e Porphyry Ertsberg district North Grasberg Hydrothermal Diorite Propylitic FNG-1 733,862 9,553,477 126.35 Previously unpublished
Intrusion
126342 12e Porphyry Ertsberg district North Grasberg Hydrothermal Diorite Phyllic FNG-1 733,857 9,553,435 265.4 Previously unpublished
Intrusion
e
126343 12 Porphyry Ertsberg district North Grasberg Hydrothermal Diorite Potassic FNG-1 733,852 9,553,399 379.3 Previously unpublished
Intrusion
126369 18e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite n/a GRS24 734,427 9,551,267 0.3 Previously unpublished
Intrusion (Early)
126375 16e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite n/a GRS24 734,427 9,551,267 159.9 Previously unpublished
Intrusion (Early)
e
126380 16 Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite n/a GRS32 734,513 9,551,321 2.7 Previously unpublished
Intrusion (Middle)
e
126385 15 Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite n/a GRS32 734,513 9,551,321 149.75 Previously unpublished
Intrusion (Middle)
126393 10e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite n/a GRS32 734,513 9,551,321 400 Previously unpublished
Intrusion (Middle)
126400 21e Porphyry Ertsberg district South Kali Hydrothermal Monzodiorite Potassic GRS32 734,513 9,551,321 610.05 Previously unpublished
126403 4e Porphyry Ertsberg district South Kali Hydrothermal Monzodiorite Potassic GRS32 734,513 9,551,321 690 Previously unpublished
126452 1e Porphyry Ertsberg district Dalam Diorite Hydrothermal Diorite Potassic GRS37-120 734,289 9,551,082 416.9 Previously unpublished
126456A 1e Porphyry Ertsberg district Dalam Diorite Hydrothermal Diorite Potassic GRS37-120 734,223 9,550,997 582.45 Previously unpublished
126456B 17e Porphyry Ertsberg district Dalam Diorite Hydrothermal Diorite Potassic GRS37-120 734,223 9,550,997 582.45 Previously unpublished
126459 16e Porphyry Ertsberg district Dalam Diorite Hydrothermal Diorite Phyllic– GRS37-120 734,152 9,550,908 763.15 Previously unpublished
potassic
126472 22e Porphyry Ertsberg district South Kali Hydrothermal Monzodiorite Phyllic GRS37-217 734,536 9,551,394 230.5 Previously unpublished
126476 18e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite n/a GRS37-217 734,590 9,551,461 400.5 Previously unpublished
Intrusion (Early)
126479 15e Porphyry Ertsberg district Dalam Diorite Hydrothermal Diorite Potassic GRS37-217 734,621 9,551,499 500.35 Previously unpublished
126482 20e Porphyry Ertsberg district Dalam Diorite Hydrothermal Diorite Potassic GRS37-217 734,662 9,551,549 630.25 Previously unpublished
126486 16e Porphyry Ertsberg district Border Phase Hydrothermal Diorite Massive sulfide GRS37-217 734,692 9,551,587 730.4 Previously unpublished
126489 1e Porphyry Ertsberg district Border Phase Hydrothermal Limestone Massive sulfide GRS37-217 734,726 9,551,629 841.3 Previously unpublished
126495 19e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite Phyllic GRS37-219 734,489 9,551,334 198.7 Previously unpublished
Intrusion (Late)
126503A 9e Porphyry Ertsberg district South Kali Hydrothermal Monzodiorite Phyllic GRS37-219 734,536 9,551,378 500.1 Previously unpublished
126503Bi 5e Porphyry Ertsberg district South Kali Hydrothermal Monzodiorite Phyllic GRS37-219 734,536 9,551,378 500.1 Previously unpublished
126503Bii 6e Porphyry Ertsberg district South Kali Hydrothermal Monzodiorite Phyllic GRS37-219 734,536 9,551,378 500.1 Previously unpublished
126506 18e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite Phyllic GRS37-219 734,566 9,551,406 651.15 Previously unpublished
Intrusion (Middle)
126516 19e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Phyllic GRS37-260 734,964 9,551,624 1.5 Previously unpublished

P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32


126406 16e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite Potassic GRS37-45 734,426 9,551,264 100.4 Previously unpublished
Intrusion (Middle)
126412 13e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite Potassic GRS37-45 734,318 9,551,171 359.7 Previously unpublished
Intrusion (Early)
e
126418 7 Porphyry Ertsberg district Dalam Diorite Hydrothermal Diorite Potassic GRS37-45 734,197 9,551,075 600.1 Previously unpublished
126420 1e Porphyry Ertsberg district Dalam Diorite Hydrothermal Diorite Potassic GRS37-45 734,159 9,551,047 700.15 Previously unpublished
126424 14e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite Phyllic GRS37-59 734,537 9,551,410 0.1 Previously unpublished
Intrusion (Middle)
126433 17e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite Potassic GRS37-59 734,633 9,551,537 159.1 Previously unpublished
Intrusion (Middle)
e
126534 7 Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Phyllic GRS93B 734,220 9,551,567 0.1 Previously unpublished
126540 16e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Phyllic GRS93B 734,220 9,551,567 251.5 Previously unpublished
126547 9e Porphyry Ertsberg district Main Grasberg Hydrothermal Monzodiorite Potassic GRZ37-43 734,408 9,551,236 83.6 Previously unpublished
Intrusion (Middle)
126551 16e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Potassic GRZ37-43 734,346 9,551,159 180 Previously unpublished
126554 20e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Potassic GRZ37-43 734,296 9,551,097 262 Previously unpublished
126557 13e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Potassic GRZ37-43 734,241 9,551,028 349.2 Previously unpublished
126560 23e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Phyllic– GRZ37-43 734,171 9,550,942 459.5 Previously unpublished
potassic
126565 3e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Phyllic GRZ37-43 734,074 9,550,822 620 Previously unpublished
126344A 1e Porphyry Ertsberg district Kay Intrusion Hydrothermal n/a Phyllic KI1-1 734,337 9,549,191 571.15 Previously unpublished
126344B 5e Porphyry Ertsberg district Kay Intrusion Hydrothermal n/a Propylitic KI1-1 734,337 9,549,191 571.15 Previously unpublished
126566 22e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Phyllic SGR-2 734,032 9,550,724 1 Previously unpublished
126568 24e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Phyllic SGR-2 734,030 9,550,703 50 Previously unpublished
126574 24e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Phyllic SGR-2 734,027 9,550,649 178 Previously unpublished
126345 1e Porphyry Ertsberg district Dalam Volcaniclastics Hydrothermal n/a Phyllic Surface sample Previously unpublished
126350 18e Porphyry Ertsberg district Ring Dyke Hydrothermal n/a Potassic Surface sample Previously unpublished
126347 20e Porphyry Ertsberg district Dalam Andesite Hydrothermal Andesite Propylitic Surface sample Previously unpublished
126333 12e Skarn Ertsberg district Ertsberg East Skarn Hydrothermal Skarn Skarn Surface sample Previously unpublished
a
Paleoproterozoic.
b
Neoarchean.
c
Mesoproterozoic.
d
Mine grid coordinates.
e
EMPA data.

13
14 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

range from 300 to 800 °C in porphyry deposits and approximately from Batholith (Spong, 1998) have previously been used for comparison of
300 to 500 °C in the associated skarn deposits (e.g., Ahmad and Rose, magnetite from porphyry and associated skarn deposits, and can be
1980; Audetat and Pettke, 2006; Beane and Titley, 1981; Manske and employed as a baseline that highlights relative enrichment or depletion
Paul, 2002; Sillitoe, 2010). The data presented here are from five classic of foreign cations in magnetite (Nadoll, 2011; Nadoll et al., 2012).
porphyry and related skarn deposits (Fig. 5A): (1) Morenci, Arizona,
which is North America's largest producer of copper and one of the larg- 4. Magnetite chemistry
est open pit mines in the world; (2) the Safford district, Arizona, with its
four porphyry Cu deposits; (3) Chino (Santa Rita), New Mexico, which A total of 1415 EMPA and LA-ICP-MS analyses, of which 1103
was one of the world's first low-grade open pit copper mines, (4) the have not been previously published, represent the database for our
Cobre porphyry Cu and skarn deposit, New Mexico, and (5) the Copper geochemical evaluation. Summary statistics are shown in Table 4
Flat porphyry and Zn-skarn deposit, New Mexico. Our database includes and the corresponding box and whisker plots are shown in Fig. 7.
previously published data for 100 analyses of igneous magnetite and Probability plots (Fig. 8) and an upper threshold value radar plots
118 analyses of hydrothermal magnetite from these porphyry de- (Fig. 9) are essentially derivates of the box and whisker plots, but
posits. Furthermore, 59 analyses of hydrothermal magnetite from help to identify compositional trends by displaying different aspects of
calcic skarn, and 17 analyses of hydrothermal magnetite from mag- the data. Probability plots help to distinguish sub-populations within
nesian skarn are presented here. All of these data have been previ- groups and radar plots concisely illustrate relative variations for the
ously presented in a PhD thesis (Nadoll, 2011). 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
3.3.2. Ertsberg district (Papua, Indonesia) for magnetite based on their common incorporation into the magnetite
The Ertsberg district is among the world's largest Cu–Au resources structure at detectable limits and their data availability across EMPA
districts and contains a number of porphyry and skarn Cu–Au–(Mo) and LA-ICP-MS datasets. Further important elements, such as Si, Pb,
deposits. The Grasberg Igneous Complex is located within the Ertsberg W, and Nb, that can, in some cases, also be useful discriminators, are
district, in the central highlands of New Guinea in the province of discussed where applicable.
Papua (formerly Irian Jaya), Indonesia (Fig. 5D). Pliocene intrusions of
dioritic to monzonitic compositions were emplaced in deeply folded
siliciclastic and carbonate host rocks of Tertiary age to produce two 4.1. Deposit type comparison
porphyry orebodies (ca. 3 Ma; Mathur et al., 2000) with disseminated
and stockwork-type mineralization and sulfide-rich skarn at the The descriptive illustration of compositional variations in magnetite
margins of the igneous complex (Fig. 5E) (MacDonald and Arnold, from various geological settings, including hydrothermal ore deposits,
1994; Pollard and Taylor, 2002; Pollard et al., 2005). Grasberg is one of lays the foundation for a more abstract classification scheme that is
the biggest Cu and Au mines in the world (MacDonald and Arnold, used to discriminate different types of mineral deposits and, hopefully
1994; Meinert et al., 1997; Pollard et al., 2005). The main Grasberg will help exploration geologists to target mineralized areas.
intrusion (ca. 3 Ma) is characterized by successive stages of intense
K-feldspar, biotite, and magnetite alteration (Pollard and Taylor, 4.1.1. Banded iron formation and related iron ore deposits (158 analyses)
2002). The nearby Ertsberg East Skarn (ca. 2.6 Ma; Pollard et al., Magnetite in BIF and iron ore deposits commonly have only trace
2005) is a calcium–magnesium silicate skarn hosted by carbonates amounts of Fe-substituting cations. Seventy to 90% of the analyses
of the Tertiary Waripi and Faumai Formations (Mertig et al., 1994). show concentrations of less than 10 ppm for Ti, V, Cr, Co, Ni, and Zn
Magnetite of igneous and hydrothermal origin is a widespread accesso- and merely 40% of the data exceed 1 ppm for Ga. Only Mg, Al, Si (not
ry to major mineral phase in the intrusive rocks and hydrothermal shown), and Mn consistently have concentrations above 10 ppm in
veins, as well as in the associated skarn. A total of 892 previously unpub- magnetite from BIF and iron-ore deposits. In particular, Mg and Si
lished EMPA analyses of hydrothermal magnetite from the Ertsberg show trends towards greater concentrations, with median values of
district are included in our database. Of these, 722 represent hydrother- 556 ppm and 8000 ppm, respectively (Figs. 7, 8, Table 4). Higher values
mal magnetite from the Grasberg porphyry and associated intrusive with positive correlation among Si, Mg, Al, Ca, and/or Mn are common
rocks and 59 are from hydrothermal skarn magnetite. An overview of but may reflect contamination from ubiquitous micro to nanometer
the geology and sample locations is shown in Fig. 5 and Table 3. 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
3.4. Plutonic igneous rocks S. The color-coded field emission gun SEM image in Fig. 10 illustrates
that mineral inclusions can be an abundant feature in magnetite grains
3.4.1. Climax-type Mo deposit (Henderson, Red Mountain), Colorado and have the potential to contribute and skew chemical analyses.
A set of igneous magnetites (41 analyses) is from the Henderson Microprobe data from several BIFs from North America (Dupuis and
Climax-type Mo deposit (Fig. 5B), which is the world's largest operating Beaudoin, 2011) show median concentrations of Si 1400, Zn 600, Al 350,
molybdenum mine (Carten et al., 1993), located on the east side of the Cu 340, Ni 310, Cr 250, Mn 180, V 160, Ti 150, Mg 130, and Ca 90 ppm,
Continental Divide 50 miles west of Denver, Colorado. The stock, and thus yield enriched concentrations compared to our LA-ICP-MS data
represented by a leucorhyolite/leucogranite intrusive complex with (Fig. 12). Control analyses to compare EPMA (Dupuis and Beaudoin,
high-temperature quartz–molybdenum veins and fractures, is potassically 2011) and ICPMS (Chicoutimi) data revealed that, at ICPMS values
altered and enveloped by distal quartz–sericite–pyrite (QSP) alteration b50 ppm, EPMA data of Zn, Cu, Ni, V, and partly Cr are at least one
(Seedorff and Einaudi, 2004; Shannon et al., 2004). Magnetite is a common order of magnitude higher than ICPMS data. Thus, data for Zn, Cu, Ni, Cr,
disseminated accessory mineral in the igneous host rocks, and, although no and V in BIF (Dupuis and Beaudoin, 2011) may be overestimated as a re-
data for hydrothermal magnetite from this deposit are presented here, it sult of insufficiently low microprobe detection limits.
also occurs in hydrothermal veins. Metamorphic magnetite in unaltered BIF rarely contains more than
10 ppm Co and 50 ppm Ni and has Co/Ni ratios ranging from 0.1 to 1
3.4.2. Inner Zone Batholith (Japan) (this study, Bajwah et al., 1987). No significant differences between
The un-mineralized granodiorite of the Inner Zone Batholith in Algoma type and Superior type BIF have been observed. In comparison
southwest Japan (Fig. 5C) is one of the classic magnetite-series granitic to metamorphic magnetite, hydrothermally altered magnetite in low
rocks first described by Ishihara (1977). Median minor and trace ele- grade metamorphic hematite–chert BIF from the Carajas Serra Norte de-
ment concentrations for igneous magnetite from the Inner Zone posits, Brazil, is characterized by average Co and Ni concentrations that
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 15

exceed 500 and 350 ppm, respectively, and Co/Ni ratios greater than 1 Distinguishing compositional features that help discriminate between
(Figueiredo e Silva et al., 2009). magnetite of igneous and hydrothermal origin are described in the fol-
lowing section. In comparison to magnetite from other settings, igneous
4.1.2. Ag–Pb–Zn veins (30 analyses) porphyry magnetite has the highest median V and Cr concentrations
Median minor and trace element concentrations in hydrothermal and the highest upper threshold value for Ti (7.63 wt.%, Fig. 7). Hydro-
magnetite from the Lucky Friday Ag–Pb–Zn veins are comparatively thermal porphyry magnetite has characteristically high Co and Ni concen-
low, with the lowest median concentration for Mg observed in the trations that are comparable only to those found in hydrothermal skarn
data (Fig. 7, Table 4). Median concentrations for Al, Cr, Mn, Co, Zn, and magnetite. Both igneous and hydrothermal porphyry magnetite have me-
Ga are also characteristically low compared to most other magnetite dian V concentrations that exceed those found in other types of magnetite
occurrences; only magnetite from BIF and iron ore deposits has more by an order of magnitude. Furthermore, characteristically high Ga
depleted Al, Zn and Ga concentrations. contents are a feature for both types, and are only exceeded by igneous
The minor and trace element concentrations in hydrothermal mag- magnetite from Henderson Climax-Mo deposit.
netite from the Lucky Friday Ag–Pb–Zn veins range widely. In particular,
Mg, Mn, and Ni have minimum and maximum concentrations that 4.1.4.1. Hydrothermal vs. igneous porphyry magnetite. When comparing
cover four orders of magnitude. This compositional variation can partly hydrothermal and igneous magnetite it is important to acknowledge
be attributed to the different vein types that occur in the Coeur d'Alene that many important spinel elements such as Ti, V and Cr can be
mining district. Hydrothermal magnetite from siderite veins displays highly variable among different types of igneous host rocks and re-
higher Mg and Mn concentrations compared to magnetite in calcite flect a wide range of whole rock compositions and cooling histories.
veins as described in detail by Nadoll et al. (2012). This distinguishing Magnetite from mafic and ultramafic igneous rocks often has higher
feature is illustrated in the probability plots in Fig. 8, which, in addition concentrations of these elements than magnetite crystallizing from
to Mn (split at 0.5), Mg (split at 0.4), and Ni (split at 0.38), shows two felsic to intermediate magmas (e.g., Dare et al., 2012; Grigsby, 1990;
distinct sub-groups for V (split at 0.5). Lindsley, 1991). Magnesium, Ti, V, Cr, and Co are useful discriminator el-
ements that can differentiate porphyry magnetite of igneous and hy-
4.1.3. Calcic and magnesian skarn (229 and 17 analyses) drothermal origin. Titanium is one of the best discriminators: only
Hydrothermal magnetite from calcic and magnesian skarn has the 10% of hydrothermal porphyry magnetites contain more than 1 wt.%
highest recorded median Mg concentrations. Gallium concentrations Ti, whereas 30% of igneous magnetite from porphyry deposits contain
are comparable in both types. Calcic and magnesian skarn magnetite more than 1 wt.% Ti. Similarly, lesser V concentrations are more indica-
display distinct variations in their Mn, Ti, V, Co, and Zn concentrations. tive of hydrothermal porphyry magnetite than for igneous porphyry
These elements have concentrations that are one to two orders of magnetite. Approximately 50% of hydrothermal porphyry magnetites
magnitude greater in calcic skarn magnetite than in magnesian skarn have V concentrations below 1000 ppm, whereas only 20% of igneous
magnetite. Furthermore, median Mn and Zn concentration in magnetite porphyry magnetites have V concentrations below that level (Fig. 8).
from calcic skarn are an order of magnitude greater than in any other Chromium concentrations are greater in igneous porphyry magnetite
type of magnetite, with 60% exceeding 10,000 ppm for Mn and 80% of than in hydrothermal occurrences. Hydrothermal porphyry magnetite
the measured Zn concentrations exceeding 1000 ppm. This trend is has median Cr concentrations of 370 ppm, whereas igneous porphyry
also reflected in the upper threshold values for calcic skarn magnetite, magnetite has a median Cr content of 544 ppm. The difference is
which significantly exceed the values for Mg, Mn and Zn measured in more evident in their upper threshold values: 3290 ppm for igneous
any other type of magnetite. Magnesian skarn magnetite has the lowest and 426 ppm for hydrothermal magnetite (Fig. 9, Table 4). Cobalt con-
overall upper threshold values for Ti and Co in the dataset (Table 4, centrations display the opposite trend, with greater concentrations in
Fig. 7) and Ni and Cr concentrations lie below the detection limit. hydrothermal porphyry magnetite. Approximately 60% of hydrother-
Upper threshold values for Mn are lower in magnesian skarn magnetite mal porphyry magnetites have Co concentrations greater than
than in most other magnetite types. Magnesian skarn magnetite incor- 100 ppm, whereas Co concentrations in excess of 100 ppm could
porates characteristically high Sn concentrations with a median of only be recorded in two analyses (Fig. 8). Magnesium concentrations
82.5 ppm and an upper threshold value of 99.1 ppm (Table 4, Fig. 7). can also help to distinguish hydrothermal from igneous porphyry
Igneous magnetite from the Henderson Climax-type Mo deposit is the magnetite. Hydrothermal occurrences display a much larger range
only other magnetite occurrence with comparable Sn concentrations. of Mg concentrations, with concentrations above 1000 ppm and
The probability plot in Fig. 8 shows that magnetite hosted in magnesian below 100 ppm being indicative for hydrothermal occurrences. Igne-
skarn has two distinct sub-populations with characteristically different ous porphyry magnetite shows less variability in Mg concentrations
Mg concentrations — illustrated by the split at 0.48 in the probability (100–1000 ppm) (Table 4, Fig. 7).
plot. Nadoll (2011) have linked the two populations to magnetite
from outside and inside the chalcocite enrichment zone of the Chino 4.1.5. Plutonic igneous rocks
(Santa Rita, USA) skarn. Similarly magnetite from calcic skarn can be di-
vided into sub-populations based on characteristic variations in Mg and 4.1.5.1. Climax-type Mo (Henderson, Red Mountain) (41 analyses). Igne-
Mn. Low-Mg (b1 wt.%) and low-Mn (90% of sub-population below ous magnetite from the Henderson Climax-type Mo deposit has one
2000 ppm) skarn magnetite also shows a tendency to incorporate less of the most distinct compositional signatures. Gallium concentrations
Al and Ti than the rest of the skarn magnetites. This trend could not are an order of magnitude higher than in any other type of magnetite
yet be linked to a specific locality, host rock, or alteration zone but it and Sn is only present at comparable concentrations in hydrothermal
has to be noted that most Ertsberg skarn magnetites fall outside the magnetite from magnesian skarn (Fig. 11). Median Sn concentrations
low-Mg–Mn population. reach 88.2 ppm and the upper threshold value is 175 ppm. Tin is an el-
ement that is not commonly present at concentrations exceeding trace
4.1.4. Porphyry Cu–(Au)–(Mo) deposits — hydrothermal and igneous levels (≤ 10 ppm) and therefore the high concentrations in igneous
(840 and 100 analyses) magnetite from Henderson are characteristic and can serve as a unique
Igneous and hydrothermal magnetite from the porphyry Cu– compositional fingerprint. Molybdenum could not be detected. Median
(Au)–(Mo) deposits display similar trends for several magnetite ele- Ti concentrations are the highest in the dataset, and Al and Mn concen-
ments. Median and upper threshold Al, Mn, Zn, Ga, and Sn concentrations trations are greater than in most other settings and are only exceeded
are similar, and, although those elements can help to classify local sample by hydrothermal magnetite from skarn deposits. Chromium could not
sub-populations, their use as overall discriminators seems to be limited. be detected in magnetite from the Henderson Climax-type Mo deposit.
16 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

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 mag-
netite 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 Friday mine, Coeur d'Alene 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). (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–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 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, Qz: quartz, Sd: siderite, Ser: sericite, Sp: sphalerite, Srp: serpentine.

Overall, the minor and trace element signature of igneous magnetite 4.1.5.2. Inner Zone Batholith (12 analyses). Minor and trace element
from Henderson is distinctive and can be easily discriminated from concentrations in igneous magnetite from the Inner Zone Batholith
other magnetite occurrences. However, due to a very limited number (IZB) show similar trends to other igneous magnetites (Figs. 7, 8,
of samples, it is an open question if this distinctive signature in igneous Table 4). Concentrations of elements such as Mg, Al, Co, Ni, Zn, and Ga
magnetite is also reflected in hydrothermal magnetite from this deposit. are comparable to those found in igneous porphyry magnetite. Igneous
If so, that would be a clear indicator for a strong host rock control. magnetite from the Inner Zone Batholith is characteristically high in V,
Table 4
Summary statistics for selected hydrothermal and igneous magnetite occurrences based on EMPA and LA-ICP-MS analyses. The corresponding graph is shown in Fig. 7.

BIF-Iron ore, hydrothermal Ag–Pb–Zn — hydrothermal Mg-Skarn — hydrothermal Skarn — hydrothermal

# Min Max Med 0.95 # Min Max Med 0.95 # Min Max Med 0.95 # Min Max Med 0.95

Mg, ppm 98 14.8 2880 556 2160 29 4.86 11,500 129 10,000 17 5140 26,700 21,900 26,700 203 49.7 59,400 30,700 54,000
Al, ppm 137 0.27 1510 86.5 841 30 145 4490 362 3210 9 1100 2240 1810 2240 214 580 31,700 9500 16,700

P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32


Ti, ppm 134 0.8 117 17.8 105 30 9.25 21,900 154 11,100 4 15.2 86.1 19.2 86.1 183 134 7180 3020 6410
V, ppm 135 0.26 40.5 4.83 15.7 26 1.09 232 136 219 8 15 281 20.1 281 51 38.2 2510 518 2310
Cr, ppm 73 1.14 356 5.17 48.9 20 5.59 320 69.6 318 0 bdl bdl bdl bdl 56 270 4720 458 3800
Mn, ppm 137 3.66 507 41.8 251 30 21.4 45,000 80.2 32,300 17 972 1610 1390 1610 229 350 50,400 17,500 46,100
Co, ppm 124 0.03 21.9 0.8 5.75 18 1.06 19.4 9.3 19.4 4 2.64 4.09 3.44 4.09 59 31.3 990 440 950
Ni, ppm 120 0.27 36.4 2.36 31.2 20 1.11 152 85.2 150 0 bdl bdl bdl bdl 41 61.3 920 490 800
Zn, ppm 104 1.15 38.9 4.87 30.7 30 17.6 216 58.4 214 9 175 311 220 311 194 79.4 15,100 7640 13,300
Ga, ppm 107 0.01 5.2 0.93 2.4 27 0.05 9.4 3.83 8.74 10 25.9 43.4 28.4 43.4 36 6.68 134 27.7 129
Sn, ppm 37 0.17 1.48 0.3 1.14 23 0.17 19.6 2.36 17.5 10 67.6 99.1 82.5 99.1 6 8.85 53.1 23.4 53.1

Porphyry — hydrothermal Porphyry — igneous Climax-Mo — igneous IZB — igneous

# Min Max Med 0.95 # Min Max Med 0.95 # Min Max Med 0.95 # Min Max Med 0.95

Mg, ppm 583 110 32,200 840 24,700 65 103 3940 478 1940 2 227 508 367 508 9 302 844 543 844
Al, ppm 750 384 15,900 2300 7810 75 459 9720 2260 7770 41 2000 19,100 5220 17,900 12 1260 3330 1820 3330
Ti, ppm 733 176 37,900 1930 19,600 94 191 97,800 1650 76,300 41 553 11,200 4040 10,700 12 50 419 168 419
V, ppm 715 57.8 18,200 1470 3800 96 367 6660 1970 3960 9 74.1 1160 272 1160 12 2210 4140 3600 4140
Cr, ppm 78 270 2240 370 696 14 194 3290 544 3290 0 bdl bdl bdl bdl 5 32.6 198 68.2 198
Mn, ppm 748 77.2 19,300 1000 8890 99 115 10,500 464 7080 41 737 12,500 5910 11,300 12 441 3200 1540 3200
Co, ppm 143 8.47 760 380 646 62 4.95 131 45.6 83.1 13 10.6 78.2 31.5 78.2 12 31.9 502 76.1 502
Ni, ppm 183 15.9 1090 400 724 28 21.7 448 77.8 434 1 403 403 403 403 12 33 192 73.6 192
Zn, ppm 563 20.1 6210 530 3090 54 48 2430 456 2000 34 253 3300 651 2920 12 50 2080 249 2080
Ga, ppm 86 12.9 151 47.7 115 72 15.1 108 59.4 98.3 41 94.6 280 181 259 7 15.3 150 69.9 150
Sn, ppm 14 2.53 70.1 7.78 70.1 10 2.3 67.6 7.78 67.6 35 42.5 210 88.2 175 0 bdl bdl bdl bdl

17
18 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

with the highest reported median concentration of 3600 ppm. Titanium Rusk et al. (2009) and Rusk et al. (2010) reported a common incorpora-
and Cr concentrations on the other hand are significantly lower tion of Mg, Al, Ti, V, Cr, Mn, Co, Ni, Zn, and Ga at detectable concentra-
compared to porphyry magnetite. Median concentrations for these tions and confirmed the high Mn concentrations in hydrothermal
elements lie at 168 ppm and 68.2 ppm, respectively. magnetite from the Ernest Henry deposit. However, their results
showed that hydrothermal magnetite associated with mineralization is
4.1.6. Magnetite from iron oxide copper–gold deposits not relatively depleted in V compared to barren hydrothermal occur-
Iron oxide copper–gold (IOCG) deposits have recently undergone in- rences. Rusk et al. (2010) show that Ti/Mn ratios in hydrothermal mag-
tense study (e.g., Barton and Johnson, 2004; Sillitoe, 2003; Williams netite present the best discriminator between barren and mineralized
et al., 2005). Magnetite is an abundant mineral in many IOCG orebodies, systems, and between different deposits. However, Zhang et al. (2009)
and, although we present no data for magnetite from IOCG deposits, point out that no systematic compositional trends exist in magnetite
several studies have reported chemical trends in magnetite from these from IOCG deposits that directly link minor and trace element signatures
deposits. The IOCG field defined by Dupuis and Beaudoin (2011) in of magnetite with paragenesis, rock type, or mineralization.
their Ti + V vs. Al + Mn diagram plots at slightly lower Ti + V and
Al + Mn concentrations than those commonly found in hydrothermal 4.2. Classification of orebody magnetite
magnetite from porphyry deposits (Fig. 12). Carew (2004) investigated
magnetite from the Cloncurry region and the Ernest Henry IOCG deposit. A number of researchers have suggested diagrams to discriminate
He found characteristically high Sn and Mn concentrations and low V, Ti, magnetite from different types of mineral deposits (Beaudoin and
Mg, Si, Cr, and Zn contents compared to barren magnetite occurrences. Dupuis, 2009; Dupuis and Beaudoin, 2011; Kamvong et al., 2007;

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 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.
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 19

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.
20 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

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.

Nadoll, 2011; Nadoll et al., 2012; Singoyi and Zaw, 2001). Nadoll et al. factor score plots. They established five factors that can discriminate be-
(2012) present data for low temperature hydrothermal and burial tween igneous, hydrothermal, and burial metamorphic magnetite,
metamorphic magnetite from the Belt Supergroup. They show that var- namely (F1) Mg–Mn, (F2) Ga–Zn–Cr, (F3) Co–Ni–V, (F4) Al, and (F5)
iations in minor and trace elements between these two types of magne- Ti. Their data for hydrothermal magnetite from Ag–Pb–Zn veins,
tite are subtle but can be resolved using multi-element radar plots and which is included in this study, suggest that hydrothermal magnetite
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 21

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 hydrothermally altered host rock.

from mineralized areas is enriched in Mg and Mn in comparison to igneous magnetite from the barren Inner Zone Batholith. Igneous and
the host rock magnetite. Singoyi et al. (2006) and Kamvong et al. hydrothermal magnetites from porphyry deposits show a similar Ti/V
(2007) proposed that Sn/Ga can be used to discriminate different ratio with over 90% of the data plotting below 10, and median ratios
types of hydrothermal magnetite. Our data show a distinct grouping of 0.9 and 2.1 respectively. Hydrothermal magnetite from BIF (median
of the various hydrothermal magnetite populations (Fig. 11), however, 3.5) and Ag–Pb–Zn veins (median 5.9) also show similar Ti/V ratios.
Sn concentrations commonly lie below detectable limits or, when the Ti/V ratios for hydrothermal skarn magnetite are the third highest in
analytic resolution yields data, concentrations lie below 10 ppm. the dataset with a median value of 4.3. Apart from hydrothermal
Dupuis and Beaudoin (2011) recently proposed that Ni + Cr vs. porphyry magnetite which plots at similar Ti + V/Al + Mn ratios
Si + Mg, Al/(Zn + Ca) vs. Cu/(Si + Ca), Ni/(Cr + Mn) vs. Ti + V, to igneous magnetite from barren and mineralized host rocks (~ 1),
and Ca + Al + Mn vs. Ti + V diagrams can be used to discriminate other hydrothermal magnetite, from calcic and magnesian skarn,
among a wide variety of mineral deposits, including iron oxide–cop- BIF and Ag–Pb–Zn veins, has distinctly lower ratios — between
per–gold (IOCG), banded iron formation (BIF), porphyry Cu, Fe–Cu 0.01 and 0.5 (Fig. 13C–D).
skarn, Ni–Cu–PGE, Cu–Zn–Pb volcanogenic massive sulfide (VMS), Upper threshold values are a robust approach when comparing differ-
and Archean porphyry Cu–Au. To test these suggested deposit fields we ent types of hydrothermal and igneous magnetite analyzed by different
plotted our data in their Ni/(Cr + Mn) vs. Ti + V and (Ca) + Al + Mn labs and under different operating conditions. They are usually not affect-
vs. Ti + V diagrams (Fig. 12). We decided to omit Ca because it contrib- ed by varying detection or reporting limits across a combined dataset.
utes negligible amounts to the overall cation sum — typically one to two Upper threshold values for the presented data span two to four orders
orders of magnitude lesser concentrations than the Al + Mn cation of magnitude and display several compositional trends among the differ-
sum. Furthermore, LA-ICP-MS detection limits for Ca are often high and ent types of magnetite (Fig. 9). There is, for example, a distinctive differ-
due to the relatively large LA beam size Ca is commonly attributed to ence between igneous and hydrothermal magnetite in regard to their
inclusions where present at greater concentrations (Nadoll and Koenig, upper threshold concentrations for Mg. Igneous magnetite has values in
2011; Nadoll et al., 2012). However, other workers have used Ca concen- the lower 1000 ppm to upper 100 ppm region, whereas hydrothermal
trations in magnetite determined by trace mode EMPA to discriminate magnetite across all occurrences lies at or above 1 wt.%. The highest
skarn and porphyry deposits (e.g., Dupuis and Beaudoin, 2011; Ray and upper threshold values for Mg as well as for Mn, Co, Ni, and Zn can be
Webster, 2007). Median values for the individual types of hydrothermal found in hydrothermal magnetite from skarn. Upper threshold values
magnetite roughly confirm the porphyry and skarn fields laid out by for Ti, V, Cr, Co, Ni, and Ga are lowest in hydrothermal BIF-Iron ore mag-
Dupuis and Beaudoin (2011). However, in particular hydrothermal por- netite. Igneous magnetite from Henderson has the lowest upper thresh-
phyry magnetite extends into the neighboring IOCG, skarn, Kiruna, and old value for Mg and the highest for Ga among all occurrences. Igneous
Fe–Ti, V fields. Hydrothermal magnetite from BIF-Iron ore plots at signif- porphyry magnetite can be discriminated from the other types of magne-
icantly lesser Ti + V values and extends past their proposed BIF field on tite by its high upper threshold values for Ti and Cr. Vanadium, Mn, Ni, Zn,
the Al + Mn axis by several orders of magnitude. We suggest extending and Ga values in igneous porphyry magnetite are comparable to upper
the BIF field toward lower Ti + V and Al + Mn values (Fig. 12A). The threshold values observed in hydrothermal porphyry magnetite. The
graph in Fig. 12A also shows the skarn and porphyry fields suggested by overall trends in upper threshold values for the different settings can be
Nadoll (2011). Although hydrothermal magnetites from skarn are neatly summarized as follows: (I) BIF (hydrothermal) — low Al, Ti, V, Cr, Mn,
enclosed by their proposed field boundaries, the hydrothermal porphyry Co, Ni, Zn, Ga and Sn; (II) Ag–Pb–Zn veins (hydrothermal) — high Mn
magnetite extends into the skarn field. The corresponding data that these and low Ga and Sn; (III) Mg-skarn (hydrothermal) — high Mg and Mn
deposit-type fields are based on are also part of our database (Table 3). and low Al, Ti, Cr, Co, Ni and Ga; (IV) skarn (hydrothermal) — high Mg,
Although element ratio diagrams can help to discriminate magnetite Al, Cr, Mn, Co, Ni and Zn and low Sn (V) porphyry (hydrothermal) —
from different mineral deposits, their field boundaries still need refine- high Ti and V and low Sn; (VI) porphyry (igneous) — high Ti, V and Cr
ment and stringent verification. In Fig. 13A–D probability plots and and low Mg; (VII) Climax-Mo (igneous) — high Al, Ga and Sn, low Mg
boxplots show Ti/V and (Ti + V)/(Al + Mn) ratios for magnetite from and Cr. We propose a combination of the discussed discriminant plots
the variety of investigated mineral deposits and geological settings. and upper threshold concentrations to discriminate hydrothermal and ig-
The plots illustrate a distinct trend for two igneous magnetite occur- neous magnetite from different deposit types.
rences. The highest Ti/V ratio (N5) is present in igneous magnetite One of the primary objectives of studies investigating the chemistry
from Henderson, whereas the lowest Ti/V ratio (b 0.1) can be found in of magnetite from an exploration perspective is to reliably discriminate
22 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

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, 2011; Nadoll et al., 2012). However, further research is re-
quired to address the most fundamental question — How can the chem-
istry of magnetite help to clearly distinguish prospective areas from
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

In the following section, we review and discuss the various factors


and processes that can directly or indirectly affect the composition of ig-
neous and hydrothermal magnetite. It has not yet been experimentally
determined whether the same factors that control the composition of
high temperature igneous magnetite apply to cooler temperature hy-
drothermal magnetite. However, first-order principles suggest that ma-
trix (melt/fluid) composition, temperature, cooling rate, pressure,
oxygen fugacity, sulfur fugacity, and silica activity, which apply to igne-
ous conditions (600–1200 °C), also apply equally to cooler temperature
hydrothermal and metamorphic conditions (200–650 °C). Relative var-
iations 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 magnetite elements are likely to display comparable trends
under hydrothermal conditions.
Fig. 11. Gallium vs. Sn diagram for the different types of magnetite (values in parts per
4.3.1. Crystallographic factors million). Igneous magnetite from the Henderson Climax-type Mo deposit and hydrother-
Crystallographic factors such as ionic radius and the overall charge mal Mg-skarn magnetite from Santa Rita have the highest Sn concentrations. Igneous
magnetite from Henderson also has characteristically high Ga concentrations. Consistent
balance place further constraints on possible substituting cations
detectable Sn levels were also found in magnetite from Ag–Pb–Zn veins and BIF-Iron
(Cornell and Schwertmann, 2003; Fleet, 1981; Goldschmidt, 1954; ore deposits but do not exceed trace amounts (b10 ppm). Overall this plot suggests that
Wechsler et al., 1984) (Fig. 1). In general, siderophile and lithophile el- lower concentrations of Ga and Sn are linked with decreasing temperatures. Tin and
ements that have similar ionic radii to Fe2+ and Fe3+ are more readily particularly Ga concentrations are significant lower in cooler temperature magnetite —
incorporated into magnetite than chalcophile elements or cations of i.e., magnetite from low temperature BIF and iron ore deposits plots at lower Ga and Sn
concentrations than higher temperature porphyry magnetite (see text and Fig. 13).
mismatching ionic radii when other controlling factors such as temper-
ature, 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.
magnetite (Nielsen and Beard, 2000; Simon et al., 2008). Most
4.3.2. Igneous processes recently Dare et al. (2012) presented a comprehensive compilation
Petrogenetic implications of compositional variations in magne- of magnetite/melt partition coefficients for igneous conditions.
tite from igneous rocks of all compositions have been the center of Magnetite partition coefficients for divalent cations such as Mg, Mn,
many geochemical studies for more than 50 years (Anderson et al., Ni, and Co have been shown to be largely independent of oxygen fugacity
2008; Bagdasarov, 1989; Borisenko and Lapin, 1972; Buddington and (Toplis and Corgne, 2002). Instead their dependency is coupled with
Lindsley, 1964; Buddington et al., 1955; Cornell and Schwertmann, ionic radius, site occupancies, and composition of the coexisting melt.
2003; Frost and Lindsley, 1991; Ghiorso and Sack, 1991a; Grant, 1984/ Silver, Au and Pb have large ionic radii compared to Fe and require
85a; Klemm et al., 1985, 2006; Lee et al., 2005; Lindsley, 1976b; coupled substitution. The incorporation of these cations in magnetite in-
Nielsen and Beard, 2000; Nielsen et al., 1994; Powell and Powell, volves structural alterations of the cubic spinel cell and is therefore less
1977; Prins, 1972; Ramdohr, 1955; Sauerzapf et al., 2008; Shcheka likely (Fleet, 1981; Wechsler et al., 1984).
et al., 1980; Wones, 1989). Source rock or magma composition, temper- Concentrations of elements such as Cr, Ni and Ti in igneous mag-
ature, pressure, cooling rate, oxygen fugacity, sulfur fugacity, silica and netite can significantly exceed those found in hydrothermal magne-
sulfide activity govern the composition of different types of igneous tite. However, igneous magnetite can also have low to negligible
magnetite (Buddington and Lindsley, 1964; Dare et al., 2012; Frost concentrations of these elements for a number of reasons other
and Lindsley, 1991; Ghiorso and Sack, 1991a; Haggerty, 1991a; than the availability of the respective element in the melt. For example,
Whalen and Chappel, 1988). Many studies have investigated sub-solidus re-equilibration during cooling of plutonic rocks leads to
magnetite–mineral partition coefficients, and magnetite–melt distribu- low-Ti magnetite that is also depleted in other minor and trace ele-
tion coefficients are well established experimentally for igneous mag- ments (Frost and Lindsley, 1991; Wones, 1989). This is particularly ap-
netite for a range of conditions and melts/host rocks (EarthRef.org, parent in felsic plutonic rocks that are commonly depleted in minor and
2013 and references therein, Ewart and Griffin, 1994; LaTourrette trace elements (Czamanske et al., 1977; Grigsby, 1990). Furthermore,
et al., 1991; Okamoto, 1979; Toplis and Corgne, 2002). These coeffi- foreign cations will partition strongly into coexisting mineral phases,
cients can vary across three to five orders of magnitude for relevant such as sulfides and silicates, under specific T–fO2–fS2 conditions,
magnetite elements depending mostly on temperature, host rock/melt resulting in depleted minor and trace element concentrations in the
composition, and oxygen/sulfur fugacity (Fig. 15, Table 5). The compo- corresponding magnetite. Although Cr, Ni and Ti are good pointers for
sition of the respective magnetite can also exert a certain control on igneous magnetite it is important to consider the evolution of the host
mineral/melt partition coefficients. For example, Ti-rich magnetite has rock and its mineral assemblage. For example, ilmenite exsolutions in
been described as a more likely host for Cu, Nb and Ta than Ti-poor igneous magnetite will render the surrounding magnetite Ti-poor.
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 23

Fig. 12. (A) Ti + V vs. Al + Mn plot with deposit fields proposed by Dupuis and Beaudoin (2011) and Nadoll (2011) — values in weight percent. Whereas the data confirm the proposed
skarn field, the porphyry and BIF data 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 are 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 tem-
perature 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.

Dare et al. (2012) have shown that large laser ablation beam sizes that magnetite has the highest median concentrations for Mg, Al, Mn, Co,
incorporate the host magnetite as well as any occurring ilmenite exso- Ni, and Zn among all types of magnetite, which might reflect one or
lutions can provide an approximation of the original magmatic compo- all of the following: (1) elevated temperatures compared to other
sition of these magnetites. types of hydrothermal magnetite (Fig. 12B), (2) high concentrations
Measured partition coefficients for igneous Fe–(Ti)-oxide minerals of these elements in the original hydrothermal fluids, and/or (3) strong
indicate that elements such as REE, Y, Sr, U, Th, Mo, Sb, and W are highly fluid–host rock interaction. Experimental data that measure the relative
incompatible (Klemme et al., 2006). Indeed there is little to no evidence significance of these factors are not yet available for hydrothermal
in our data that these elements are incorporated into magnetite at levels magnetite. However, (1) formation temperature and (2) the composi-
exceeding common LA-ICP-MS detection limits. However, other authors tion of the original hydrothermal fluid are less likely to be the defining
have reported the incorporation of Mo and W into Fe-oxide phases, such physicochemical control for Fe-substitution in magnetite because they
as magnetite, at significant levels (Candela, 1997). Mahood and Hildreth match the temperatures and fluid compositions responsible for porphy-
(1983) published partition coefficients for Mo (D ~ 2 to D ~ 20) in igne- ry mineralization in the adjacent igneous rocks. Instead, (3) significant
ous magnetite that are comparable to those for other transition metals fluid–host rock interaction is a more likely explanation for the enrich-
like Co and Mn (Table 5). The interdependencies among elements can ment of a number of elements in skarn magnetite (Einaudi et al.,
have further affects on the incorporation of foreign cations in magnetite. 1981; Meinert, 1987; Nadoll, 2011). This case exemplifies the complex
For example, Simon et al. (2008) reported that the addition of Ti to end- controls on the geochemistry of magnetite which might be further com-
member magnetite can increase the incorporation of Au and Cu. plicated by multiple successive stages of alteration and alteration zona-
tion (Nadoll, 2011; Nadoll et al., 2010).
4.3.3. Hydrothermal processes In spite of these complexities, compositional trends can be put into a
It is often difficult to attribute specific element enrichments or de- process-oriented context and increasing amounts of experimental and
pletions to one singular physicochemical factor. For example, skarn fluid inclusion data can be consulted to support these observations.
24 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

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 con-
centration of elements in the respective types of magnetite and therefore represent a robust approach for comparing multiple data sets. Panels (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 (Figs. 7 and 8) the Ti/V ratio is relatively
constant. The (Ti + V)/(Al + Mn)-ratio plot confirms the decreasing 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.

For example, several authors have shown that even low Au concentra- are commonly enriched in K, Ca, Mn, Fe, Cl, Br, Cu, Pb, Zn, and As (e.g.,
tions (sub-ppm) in magnetite can be significant because of the relative- Meinert et al., 2005). Manganese and Zn are commonly incorporated
ly large grain size and abundance of Fe–Ti oxides in many hydrothermal into hydrothermal magnetite at levels above 100 ppm, whereas the
ore deposits (e.g., Larocque et al., 2002). Most recently Tauson et al. chalcophile elements Cu, As, Mo, Ag, Au, and Pb typically have concen-
(2012) showed experimental evidence that the partition coefficient trations that are less than their detection limits.
for Au between magnetite and hydrothermal solution is 1.0 ± 0.3 Tin is rarely present at concentrations above 100 ppm and, if detect-
[450 °C, 1 kbar (100 MPa)], indicating that Au is not an incompatible el- able, is incorporated at levels at or below 10 ppm in igneous or hydro-
ement in magnetite. In comparison, the distribution coefficient of Au in thermal magnetite of all origins. However, igneous magnetite from the
magnetite is much lower than in pyrite and arsenopyrite (Tauson et al., Henderson Climax-type Mo deposits and hydrothermal magnetite
2011). Tauson et al. (2011) also observed that Au is mostly enriched in from magnesian skarn display consistently higher Sn concentrations,
the outermost layer of magnetite crystals (~500 nm). exceeding 100 ppm (Table 4, Fig. 11), which is most likely controlled
by the composition of the respective fluid or fluid/host rock interaction.
4.3.3.1. Fluid composition — element availability. Element availability in Sodium, Si, S, Ca, Cu, Ge, As, Se, Sr, Y, Mo, Ag, In, Sb, Ba, Ce, Au, Tl, Pb, and
hydrothermal fluids provides a first order control for minor and trace Th are not commonly present (tens to hundreds of ppm) in magnetite
element concentrations in hydrothermal magnetite. Data for spinel ele- from porphyry and associated skarn deposits in spite of their availability
ment concentrations in hydrothermal fluids, especially fluids related to (Nadoll, 2011). Other authors have reported Cu, As, Mo, Ag, Au and Pb to
porphyry and skarn deposits, have become increasingly available. Rusk be incorporated in magnetite from hydrothermal ore deposits (Angerer
et al. (2009) and Baker et al. (2004) list concentrations of relevant ele- and Hagemann, 2010; Carew et al., 2006; Lee et al., 2005; Lindsley,
ments that range from b 100 to over 100 ppm for Cu and 100 to several 1991; Simon et al., 2008), but reported concentrations are typically at
100 ppm for Mn, Zn, Pb, and As. Ilton and Eugster (1989) suggested that or below the 100 ppm limit for Cu, As, Mo, and Pb and even lower for
even low concentrations of these elements in hydrothermal magnetite Au and Ag. Those concentrations still lie at or below the quantification
can be an indication for a strong enrichment of these elements in the limit for many analytical techniques. Elements such as Na, Sr, Y, Ba, Ce,
corresponding hydrothermal fluid. Element availability in hydrother- Tl, and Th are highly incompatible in magnetite (e.g., Nielsen et al.,
mal fluids is a first order control for minor and trace element concentra- 1994) and even though elements such as Na and Sr may have enriched
tions in hydrothermal magnetite. Elements such as Mg, Mn, Co, and Zn concentrations in hydrothermal fluids, they are not likely to be incorpo-
that are otherwise commonly incorporated into magnetite at significant rated into magnetite (Fig. 1).
levels (100s–1000s ppm) under hydrothermal conditions can be used Seawater is the mineralizing fluid for primary and metamorphic
as indicators for the composition of a hydrothermal fluid if not offset (pre-hydrothermal alteration) iron oxides in BIF. In terms of absolute
by secondary effects such as host rock buffering. If a specific hydrother- element concentrations there is no strictly coherent discrimination
mal magnetite occurrence is depleted in these elements, it is likely that possible between Algoma- and Superior-type BIF. The Algoma dataset
the fluid was equally depleted, if other conditions such as oxygen and (Koolyanobbing, Marda, and Weld Range greenstone belts) tends to
sulfur fugacity were favorable for the incorporation of these foreign have greater Ga and lesser Ti and Mn than the Superior dataset
cations into the magnetite structure. (Hamersley Province and QF), but there is considerable overlap. A con-
Hydrothermal fluids can contain spinel elements such as Mn, Fe, Cu, sistently lower Ti/V ratio in Algoma-type BIF magnetite (Figs. 13, 14A)
Zn, As, Mo, Ag, Au, and Pb (Kesler, 2005; Meinert et al., 2005; Roedder may reflect changing availability of weathered material, from V-
and Bodnar, 1997; Yardley, 2005). These elements are commonly highly enriched ultramafic to Ti-enriched mafic, or changed oxidation states
enriched in fluids compared to their concentration in igneous host rocks of seawater that led to variation of Ti and V solubility. Similarly, high
(Audetat et al., 2000, 2008; Hedenquist and Lowenstern, 1994; Rusk concentrations of Mn (3700 ppm), Ti (800 ppm), Ni (580 ppm), Cr
et al., 2004; Ulrich et al., 1999). Fluids associated with skarn deposits (39 ppm), and V (17 ppm) in magnetite from the Garumahishani
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 25

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. (1998), Dudas et al. (1973), Esperanca et al. (1997), Ewart and Griffin (1994), Ewart
et al. (1973), 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) and Villemant et al.
(1981).

Rock/melt type Reference Al Co Cr Fe Ga Mg Mn Mo Nb Ni Pb Ti V Zn

Andesite Ewart and Griffin (1994) 24.1 4.1 5.72 2.9 15.5
Luhr and Carmichael (1980) 433 28.6 62.7 18.4
Andesite–dacite Ewart et al. (1973) 37 222
Nielsen and Beard (2000) 0.23
Basalt Horn et al. (1994) 13.3 3.2 8.1
Lemarchand et al. (1987) 3.4 2 1.4 2.6
Lindstrom (1976) 23 97
Nielsen (1992) 1.81 96 12 6.87
Righter et al. (2006a,b) 19.4 82.1 33.9
Basalt–anakaramite Nielsen et al. (1994) 96 19.4 6.87
Basalt–andesite Esperanca et al. (1997) 7.4 153 29 26 3.1
Haskin et al. (1966) 0.7
Klemme et al. (2006) 2.1 10.4 0.86
Reid (1983) 87
Basalt–andesite–dacite Okamoto (1979) 24.7 166 19.2 5.33 16.5
Basalt–hawaiite Lemarchand et al. (1987) 2.16
Basalt–latite Leeman et al. (1978) 172 25 27
Basalt–trachyte–hawaïite D'Orazio et al. (1998) 17 275 32 29
Villemant et al. (1981) 0.2 4.3 4.2 1.5 3.5 12
Dacite Ewart and Griffin (1994) 149 2.8 14.1 5.2 0.71 26.6
Granodiorite Latourrette et al. (1991) 0.12 850 10.3 15 130
High silica rhyolite Ewart and Griffin (1994) 202 4.32 49.7 0.8 58.9
Mahood and Hildreth (1983) 210 49.5 29 63
Streck and Grunder (1997) 62.9 8.4 44.4 61.8
Low silica rhyolite Ewart and Griffin (1994) 150.4 6.11 30 1.62 50.1
Mugearite Lemarchand et al. (1987) 8.52
n/a Kretz et al. (1999) 11 1.6 13
Toplis and Corgne (2002) 1.5 0.6 3.2 9.3 34.1
Pyroclastics Dudas et al. (1973) 60 16.2 39.4
Quartz latite Ewart and Griffin (1994) 115.2 12 30.9 33.3 1.58 75.2
Rhyolite Bacon and Druitt (1988) 80 30 15
Nash and Crecraft (1985) 218 65 22 215
Tacker and Candela (1987) 0.73
Tholeiite Leeman (1974) 31.6
Trachyandesite Luhr et al. (1984) 72 154 17
Trachyte Lemarchand et al. (1987) 41.7
Mahood and Stimac (1990) 21 8 11

silicate facies iron-formation from the Archean Iron Ore Group of directly linked with temperature in igneous systems (Nielsen et al.,
Jharkhand–Orissa region, India, is attributed to the mixing of 1994; Toplis and Carroll, 1995), with greater concentrations indicat-
mafic–ultramafic debris derived from weathering of oceanic crust, ing hotter temperatures. Even though Ti, and to some degree Al,
and/or post-sedimentary influx of high-temperature metalliferous seem to follow this trend for hydrothermal magnetite from BIF-
brines produced from fluid–rock interaction in the ultramafic to related ore and Ag–Pb–Zn veins, concentrations of these elements
mafic oceanic crust (Bhattacharya et al., 2007). Pecoits et al. (2009) in hydrothermal porphyry and skarn magnetite are comparable to
analyzed magnetite in BIF from the Brockman Iron Formation by those found in igneous magnetite. This suggests that temperatures
LA-ICP-MS revealing that Ni/Cr ratios show a close relationship to in these settings were not significantly different or that the concentra-
underlying mafic igneous rocks, providing evidence of a mafic- tions observed are not solely controlled by temperature. On a deposit
dominated source. Similarly, an ultramafic metal source has been in- scale, Ti and V concentrations can be two to three orders of magnitude
ferred for the Paleoarchean Isua BIF that shows high Ni/Cr ratios. The lower in hydrothermal magnetite from calcic and magnesian skarn than
Ti/V ratio is partially disturbed in hydrothermally altered magnetite in most igneous magnetites (Nadoll, 2011; Ray and Webster, 2007). El-
from BIF-hosted iron ore deposits (Fig. 14B). In the Hamersley and ements such as Al and Ti are considered to be immobile at sub-
India dataset hydrothermally altered magnetite is comparatively low magmatic temperatures (Van Baalen, 1993; Verlaguet et al., 2006),
in Ti and high in V. and are typically not detected in hydrothermal fluids. Their incorpora-
tion in magnetite is largely temperature controlled (Nielsen et al.,
4.3.3.2. Temperature. Temperatures recorded for hydrothermal alter- 1994; Toplis and Carroll, 1995) and therefore more likely under high
ation and vein formation generally lie below magmatic tempera- temperature igneous conditions. Our data confirm this assumption
tures. They are approximately 100 to 300 °C for BIF-hosted high- and show that Al and Ti have on average considerably greater concen-
grade ore (Thorne et al., 2008), 250 to 350 °C in the Coeur d'Alene trations in igneous magnetite than for hydrothermal occurrences
district (e.g., Doughty and Chamberlain, 1996; Hayes, 1990; Landis (Fig. 12B). For example, Ti concentrations are two to three orders of
and Hofstra, 1991; Leach et al., 1988; Nadoll et al., 2012), 300 to magnitude lower in hydrothermal magnetite from calcic and magne-
800 °C in porphyry deposits, and 300 to 500 °C in the associated sian skarn than in most igneous magnetites. Similar trends have been
skarn deposits (e.g., Ahmad and Rose, 1980; Audetat and Pettke, observed in magnetite from the Heff Cu–Au skarn of British Columbia
2006; Beane and Titley, 1981; Manske and Paul, 2002; Sillitoe, (Ray and Webster, 2007). The affiliation of elements such as Ti with
2010) (Fig. 12B). Aluminum and Ti concentrations in magnetite are hotter-temperature formation conditions can have implications for
26 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

elements such as Au, Ag, and Cu that are more readily incorporated into state (Bordage et al., 2011; Toplis and Carroll, 1995; Toplis and
ulvöspinel than magnetite due to its greater Ti contents (Simon et al., Corgne, 2002). Mallmann and O'Neill (2009) demonstrate that the
2008). Magnetite has been described as an important host for Au in ig- partitioning behavior of homovalent cations such as Ga however, is
neous rocks and hydrothermal ore deposits (Gammons and Williams- not dependent on the prevailing oxygen fugacity conditions.
Jones, 1997; Larocque et al., 2002; Togashi and Terashima, 1997). The In sulfide-dominated environments sulfur fugacity can be the prima-
trend toward higher Ti concentrations with hotter formation tempera- ry control for the partitioning behavior of relevant elements (Dare et al.,
tures can have repercussions on the incorporation of elements that 2012), and it can affect Fe/Mg ratios in associated minerals (Heiligmann
are directly linked with Ti. et al., 2008). Coexisting sulfide minerals have considerably higher parti-
Partition coefficients are greatly temperature dependent (McIntire, tion coefficients than magnetite for chalcophile elements such as Cu, As,
1963). The data presented here show a trend that cooler temperature hy- Ag, and Pb (e.g., Cygan and Candela, 1995; Gaetani and Grove, 1997;
drothermal magnetite generally has lesser trace element concentrations. Lemarchand et al., 1987; Simon et al., 2008; Stimac and Hickmott,
High-temperature hydrothermal porphyry and skarn magnetite shows 1994). The data presented in the present study show no evidence that
relatively higher minor and trace element concentrations that for many Cu, As, Ag, and Pb are widely incorporated at detectable levels into mag-
elements are comparable with those found in igneous magnetite, where- netite under the conditions found in the examined geological settings.
as hydrothermal–metamorphic magnetite from BIF represents the low The chalcophile element Zn seems to be an exception, as it is commonly
end of the temperature spectrum, and has the lowest overall minor present at detectable levels in hydrothermal magnetite (several
and trace element concentrations among all investigated magnetites. 10–100 ppm), and in some geological settings, it can also form Zn spi-
Therefore, temperature must be considered a major governing factor nel, gahnite (ZnAl2O4). This trend seems to be in agreement with pub-
for the composition of hydrothermal magnetite. lished Zn partition coefficients for magnetite–matrix pairs (Fig. 15,
Magnetite can accommodate measurable trace element contents at Table 5) that are considerable higher than those for Cu or As (e.g.,
temperatures below 100 °C (Sidhu et al., 1978). Re-equilibration of hy- Ewart and Griffin, 1994). Calculations based on fluid inclusion data sup-
drothermal magnetite can also have significant repercussions on minor port the argument for low concentrations of chalcophile elements such
and trace element concentrations. Foreign cations in magnetite can be as Cu in hydrothermal magnetite (Ilton and Eugster, 1989; Sawkins and
continuously expelled by this process. This is analogous to sub-solidus Scherkenbach, 1981) — with a theoretical maximum of 10 ppm under
re-equilibration of plutonic magnetite during cooling (Frost, 1991c). the temperature and oxygen fugacity conditions commonly associated
Although only semi-quantitative, there is a visible temperature trend with porphyry Cu deposits (690 °C, oxygen fugacity: NNO buffer). This
in the data. Fig. 12B shows a Ti + V vs. Al + Mn plot where data points might be an explanation why Cu is usually not reported to have signifi-
have been attributed with approximate temperatures — from low- cant concentrations in magnetite.
temperature BIF and related iron ore (Thorne et al., 2008), medium tem- Siderophile elements such as Ni and in particular Mo are restricted
perature Ag–Pb–Zn veins (e.g., Doughty and Chamberlain, 1996; Hayes, to very low oxygen and sulfur fugacity conditions because they are oth-
1990; Leach et al., 1988), medium to high temperature hydrothermal erwise more likely to form pentlandite or molybdenite, respectively
skarn and porphyry magnetite, to high temperature igneous magnetite (Drabek, 1982). Similarly, platinum group elements (PGEs) are more
from porphyry and Climax-Mo deposits (e.g., Ahmad and Rose, 1980; likely to partition into sulfide minerals than into magnetite (e.g.,
Audetat and Pettke, 2006; Beane and Titley, 1981; Manske and Paul, Barnes et al., 2004).
2002; Sillitoe, 2010). Decreasing formation temperatures are reflected
in the chemistry of the corresponding magnetite with high temperature 4.3.3.4. Silica and sulfide activity. Interdependencies among coexisting
igneous magnetite plotting at high Ti + V and high Al + Mn and low mineral phases can be crucial for trace elements concentrations in mag-
temperature hydrothermal–metamorphic BIF magnetite plotting at low netite. Analogous to oxygen and sulfur fugacity, coexisting silicates and
Ti + V and Al + Mn values. Gallium and Sn concentrations seem to fol- sulfides exert a strong control on the composition of hydrothermal
low this trend with the highest concentrations in high temperature igne- magnetite (e.g., Cygan and Candela, 1995; Frost et al., 1988; Gaetani
ous magnetite from the Henderson Climax-type Mo deposits and the and Grove, 1997; Lemarchand et al., 1987; Schilling et al., 2011; Simon
lowest concentrations in BIF magnetite (Figs. 11, 12). The temperature et al., 2008; Stimac and Hickmott, 1994). For example, sulfide minerals
dependence of Ti, V, Al, Mn, and Ga can distinguish between low and will preferentially incorporate chalcophile elements (e.g., Cygan and
high T mineral systems. However, the other controlling factors discussed Candela, 1995; Simon et al., 2008) whereas lithophile elements such
here can also exert a significant control on cation substitution and in as Mg, Al and Ti have a strong preference to partition into silicates if con-
some cases mask other process-related compositional trends. For exam- ditions are favorable (Frost, 1991b; Simon et al., 2008; Toplis and
ple, seawater composition and/or seawater oxygen fugacity for magne- Corgne, 2002). Silicates also commonly have considerably higher parti-
tite formed in BIF, or re-equilibration processes that can expel cations tion coefficients for Pb than magnetite (Bea et al., 1994; Stimac and
over time. Hickmott, 1994). Measured partition coefficients for ilmenite and
ulvöspinel indicate that elements such as Mo are highly incompatible,
4.3.3.3. Oxygen and sulfur fugacity. Oxygen fugacity within a melt is the suggesting that their incorporation in magnetite is equally unlikely
major control on element partitioning between oxides, and between (Klemme et al., 2006). However, a number of studies have shown that
oxides and silicates at magmatic temperatures (e.g., Frost, 1991a). Ele- Mo can be a compatible element with partition coefficients as high as
ments, such as Cr and V, can occur in various valence states and there- 29 (Table 5).
fore their behavior is strongly linked to oxygen fugacity (Klemme
et al., 2006; Nielsen et al., 1994; Righter et al., 2006b; Ryabchikov and 4.3.3.5. Host rock buffering. Host rock buffering is another important con-
Kogarko, 2006) (Fig. 1). Minor and trace element compositions of oxides trol for compositional variations in hydrothermal magnetite, particular-
such as titanomagnetite are highly dependent on the prevailing oxygen ly where large scale metasomatic processes are involved. It has been
fugacity (Andersen and Lindsley, 1988; Buddington and Lindsley, 1964; shown that elements like Mg and Mn can successively be enriched in
Ghiorso and Sack, 1991b; Toplis and Carroll, 1995). However, the effect hydrothermal fluids by extensive fluid/rock interactions (Einaudi
at cooler temperature hydrothermal conditions is less well constrained. et al., 1981; Meinert, 1987). This is commonly reflected in high Mg
Several studies reported that the incorporation of foreign cations is and Mn concentrations in hydrothermal magnetite occurring in skarn.
more likely at lower oxygen fugacity (Buddington and Lindsley, 1964; In particular, magnetite from magnesian skarn shows this trend and
Frost, 1991a; Lindsley, 1991). Vanadium is usually preferentially incor- points towards host rock buffering as a major control for hydrothermal
porated into the magnetite structure at low oxygen fugacity and be- magnetite in metasomatic environments such as skarn deposits. Ilton
comes incompatible at high oxygen fugacity due to its 5 + oxidation and Eugster (1989) suggest that the formation of massive magnetite
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 27

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, color-coded by deposition age. (B) 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 unaltered magnetite.

in skarn can lead to significant enrichment of Mn relative to Fe in the re- 4.4.1. Martitization and mushketovitization
maining hydrothermal fluid. If hydrothermal alteration affecting magnetite is caused by an oxi-
Compared to igneous magnetite, skarn contains magnetite with dizing or acidic fluid, then martitization (hematite after magnetite)
higher Mg, Mn and Zn concentrations (McQueen and Cross, 1998; may occur (Ohmoto, 2003). The reversed alteration, magnetite after he-
Nadoll et al., submitted for publication, e.g., Rusk et al., 2009). This matite under reducing fluid–rock reaction, is called mushketovitization.
may reflect the primary fluid composition or host rock buffering via sig- Martitization and mushketovitization are commonly pseudomorphic
nificant fluid/host rock interaction (Baker et al., 2004, e.g., Einaudi et al., processes but can, depending on the governing oxygen fugacity and pH
1981; Meinert, 1987; Nadoll, 2011). conditions, represent considerable volumetric changes in the rock
In BIF-hosted high-grade iron ore deposits, magnetite is commonly (Mücke and Raphael Cabral, 2005). Two theoretically reversible chemical
chemically and texturally altered by hydrothermal fluid–rock interac- processes have been proposed for martitization and mushketovitization:
tion. Hence, magnetite can be used as an indicator mineral for hydro- one is a redox reaction (Eq. (1)), and the other is a non-redox transfor-
thermal alteration (Angerer et al., 2012). Hydrothermal alteration of mation (Eq. (2)) under acidic conditions (Ohmoto, 2003).
metamorphic magnetite has been discussed for the distinct magnetite
chemistry in the Neo Dala BIF in the Archean Iron Ore Group by 2Fe3 O4 þ 0:5O2aq ¼ 3Fe2 O3 ð1Þ
Bhattacharya et al. (2007). Here, elevated W, Pb, As, Mo, and Sn concen-
þ 2þ
trations were attributed to hydrothermal metasomatism related to the Fe3 O4 þ 2Haq ¼ Fe2 O3 þ Feaq þ H 2 O: ð2Þ
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 The lack of significant cation exchange during hydrothermal
concentrations in hydrothermally modified magnetite in carbonate- martitization supports the oxidation reaction, because there is little mo-
and silicate-altered BIF are significantly greater than those in magnetite bilization of Fe2+ and associated cations.
from associated least-altered BIF (up to 7 in altered versus b 1 in least- Magnetite is commonly subject to martitization in weathering zones
altered BIF) (Angerer et al., 2012; Angerer et al., 2013). Because the and supergene-modified ore deposits but also in oxidative hydrother-
Yilgarn craton deposits are hosted in granite–greenstone terranes, the mal zones. It is one of the major minerals in BIF and iron ore deposits.
magnetite chemistry through the evolution of the iron ore deposits is at- The understanding of cation exchange between magnetite and oxidiz-
tributed to changes from a more mafic–ultramafic (Ni) to dominantly ing fluids during martitization is important. It has been suggested that
granitic (Co, W) host rock chemistry (Angerer et al., 2012, 2013; most spinel elements, with the exception of Mg, are retained during
Figueiredo e Silva et al., 2009). martitization under moderate-temperature hydrothermal oxidation
(Angerer et al., 2012; Sidhu et al., 1981). Other workers suggest that
4.4. Pseudomorphic magnetite alteration the martitization process can expel divalent cations because of their in-
compatible valency and ionic radii (Cornell and Schwertmann, 2003).
Magnetite grains commonly show textures, crystal habits, and Martite in regolith environments contains lesser Fe and Ni and greater
oxidation–exsolution features, such as ilmenite lamellae and Si, Al, Ti, Cr than parent magnetite, but Zn, V and Mn remain relatively
martitization along spinel planes, that reflect physicochemical forma- unchanged. Additionally, Ti and Cr may substitute for Fe in the lattice,
tion conditions (e.g., Haggerty, 1991b; McQueen and Cross, 1998; but secondary phases such as kaolinite, gibbsite and halloysite are the
Mücke and Raphael Cabral, 2005). Ilmenite exsolutions (in igneous most likely hosts for Si and Al in supergene martite (Anand and
magnetite) and oxidation to hematite are common features in magne- Gilkes, 1984).
tite. The occurrence and abundance of these features must be assessed
by careful petrography prior to EPMA and LA-ICP-MS analyses. Grains 5. Inclusions
that show only slight to moderate martitization may be easily analyzed
with the small EPMA beam size (b2 μm). Laser ablation ICP-MS on the Mineral inclusions are common features in igneous and hydrother-
other hand, with considerably larger analytical spot sizes (commonly mal magnetite; they reflect magnetite formation conditions and give a
N10 μm), can be adversely affected by martite and ilmenite in magne- direct insight into the evolution of the corresponding host rock. Mineral
tite. However, magnetite with a low degree of martitization (b10% of inclusions are of particular interest when geological and mineralogical
the grain surface) will not negatively affect LA-ICP-MS analyses context have been removed or significantly altered, which is the case
(Nadoll and Koenig, 2011; Nadoll et al., 2012). for stream sediments or regolith cover. Some inclusions, such as Cu
28 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

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 depending on host rock or melt chemistry. A wide range of formation temperatures, pressures, oxygen and sulfur fugacity are inherently as-
sociated with varying host rocks or melts. A complete list of published partition coefficients and the corresponding references are shown in Table 5.

sulfide minerals, can be a direct pointer towards mineralized areas. overgrowth on Si-poor magnetite in Dales Gorge BIF, Hamersley Prov-
Wang et al. (2012) recently reported that magnetite with cassiterite in- ince, and other BIFs worldwide.
clusions is diagnostic of oxidized tin granites and hence magnetite from
stream sediments may serve as a strategic tracer for tin exploration in 6. Concluding remarks
those environments.
Despite their virtues for discriminating sample populations, mineral One of the main challenges for mineral exploration is the recognition
inclusions need to be thoroughly assessed because of their impact on of mineralized areas at increasingly greater distance from primary
EMPA and LA-ICP-MS analyses (Nadoll and Koenig, 2011). Critical as- mineralization. Minor and trace element data for magnetite from a vari-
sessment of whether elements are in genuine solid solution or are con- ety of mineral deposits are becoming more available, and help to further
tributed by micro- and nano-scale mineral inclusions is essential to elucidate trends that can be used as geochemical signatures for prove-
rigorously define the chemistry of magnetite. Magnetite commonly nance studies and mineral exploration. Published and previously unpub-
contains sulfide, phosphate, quartz, silicate, or carbonate inclusions in lished magnetite minor and trace element data presented here give
many geological settings including burial metamorphic, BIF, igneous, evidence for a number of systematic variations among hydrothermal
and hydrothermal environments (Figs. 6, 10). Laser ablation analyses magnetite from different deposit types and igneous magnetite from min-
of microdrilled (bulk) magnetite mesobands in the Neoarchean eralized and barren host rocks. The most important factors that govern
Temagami BIF, Canada, show large ranges of Fe (36 to 57 wt.%), Si (17 compositional variations in hydrothermal magnetite are (A) tempera-
to 34 wt.%), Al, Mg, and Ca (up to 1 wt.%), and P (660 to 2230 ppm), ture, (B) fluid composition — element availability, (C) oxygen and sulfur
and therefore clearly represent mineral mixtures (Bau and Alexander, fugacity, (D) silicate and sulfide activity, (E) host rock buffering, (F) re-
2009). Inclusions such as quartz, silicates (typically talc, stilpnomelane equilibration processes, and (G) intrinsic crystallographic controls such
or amphiboles), and carbonates (siderite, magnesite, and ankerite) as ionic radius and charge balance.
will contribute Mg, Al and Ca. Phosphates and sulfides contribute P, The main discriminator elements are Mg, Al, Ti, V, Cr, Mn, Co, Ni, Zn,
Ca, REE, S, and transition metals. In some cases, it is difficult to ascertain Ga, and Sn. Their common incorporation and characteristic variations
whether inclusions are responsible for elevated concentrations of non- can be used to fingerprint magnetite from different host rocks and de-
typical spinel elements. For example, magnetite from Henderson has posit types. The classification scheme proposed by Dupuis and
significant concentrations of the high field strength (HFS) element Nb, Beaudoin (2011) offers a good baseline for the discrimination of magne-
with median values of 19 ppm. It is debatable whether these Nb con- tite from various types of mineral deposits and ongoing research will
tents represent genuine solid solution, because Nb is not usually in- provide increasingly more data to refine their suggested deposit fields.
corporated in magnetite at high concentrations — up to 248 ppm in We propose a combination of the available discriminant plots such as
individual samples from Henderson. The Nb contents may be attrib- the Ti + V vs. Al + Mn plot and box and whisker plots for the suite of
uted to mineral inclusions of pyrochlore (approximate formula) (Na, magnetite elements can distinguish among magnetite from BIFs, Ag–
Ca)2Nb2O6(OH,F) or columbite [(Fe,Mn)(Nb,Ta)2O6]. Rutile, which Pb–Zn veins, porphyry and associated skarn deposits, Climax-Mo de-
contains Nb impurities (Nb2O5) and other HFS elements of up to several posits, and barren igneous host rocks. Low levels of Al, Ti, V, Cr, Co, Zn,
weight percent (Ghent, 1975; Waychunas, 1991), is also an abundant and Ga are particularly diagnostic for magnetite from BIFs, Mg-skarn,
and widespread mineral at Henderson (Seedorff and Einaudi, 2004), and Ag–Pb–Zn veins. Igneous and hydrothermal magnetite from por-
and is a potential candidate for mineral inclusions in magnetite. Howev- phyry deposits have remarkably similar compositional patterns but
er, the relatively low Ti concentrations in magnetite from Henderson can be discriminated based on Cr, Co, Ni, and Ga concentrations. Gal-
suggest that Sn and Nb concentrations are not attributed to rutile inclu- lium concentrations are consistently greater in igneous magnetite
sions, and careful petrography showed no rutile inclusions. Further- than in hydrothermal occurrences. Based on the established upper
more, Na and Ca concentrations could not be detected in samples that threshold concentrations for the suite of magnetite elements the differ-
contain high Nb concentrations. These considerations suggest that Sn ent deposit types display the following characteristics: (I) BIF (hydro-
and Nb are incorporated in the magnetite structure at Henderson, rath- thermal) — low Al, Ti, V, Cr, Mn, Co, Ni, Zn, Ga and Sn; (II) Ag–Pb–Zn
er than residing in mineral inclusions. Although Si is typically not signif- veins (hydrothermal) — high Mn and low Ga and Sn; (III) Mg-skarn
icantly incorporated into magnetite, Huberty et al. (2012) reported (hydrothermal) — high Mg and Mn, low Al, Ti, Cr, Co, Ni and Ga;
metamorphic silician magnetite (1 to 3 wt.% SiO2 in solid solution, (IV) skarn (hydrothermal) — high Mg, Al, Cr, Mn, Co, Ni and Zn and
determined by micro-XRD and transmission electron microscopy) low Sn; (V) porphyry (hydrothermal) — high Ti and V and low Sn;
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 29

(VI) porphyry (igneous) — high Ti, V and Cr and low Mg; and (VII) Bacon, C.R., Druitt, T.H., 1988. Compositional evolution of the zoned calcalkaline magma
chamber of Mount Mazama, Crater Lake, Oregon. Contrib. Mineral. Petrol. 98, 224–256.
Climax-Mo (igneous) — high Al, Ga and Sn and low Mg and Cr. Bagdasarov, Y.A., 1989. Composition of magnetites from carbonatite complexes and depth
In combination with petrographic studies, the chemistry of hy- facies of carbonatites. In: Bagdasarov, Y.A., Efimov, A.F. (Eds.), Forecasting and Evalu-
drothermal magnetite can assist exploration geologists to target pro- ating Carbonatites. IMGRE, Moscow, USSR, pp. 124–156.
Bailey, D.K., Kearns, S., 2002. High-Ti magnetite in some fine-grained carbonatites and the
spective areas in regolith- and glacigenic environments or stream magmatic implications. Mineral. Mag. 66, 379–384.
sediments distal to mineralization. The use of mineral chemistry to Bajwah, Z.U., Seccombe, P.K., Offler, R., 1987. Trace element distribution, Co:Ni ratios and
characterize host rocks and fingerprint hydrothermal ore deposits genesis of the Big Cadia iron–copper deposit, New South Wales, Australia. Mineral.
Deposita 22, 292–300.
strongly benefits from a sound petrographic description of magne- Baker, T., Van Achterberg, E., Ryan, C.G., Lang, J.R., 2004. Composition and evolution of ore
tite, especially in terms of paragenetic sequence and foreign mineral fluids in a magmatic–hydrothermal skarn deposit. Geology 32, 117–120.
inclusions characterization. Further refinement of magnetite classifi- Balan, E., De Villiers, J.P.R., Eeckhout, S.G., Glatzel, P., Toplis, M.J., Fritsch, E., et al., 2006. The
oxidation state of vanadium in titanomagnetite from layered basic intrusions. Am.
cation schemes and experimental data are needed to gain a better
Mineral. 91, 953–956.
understanding of how physicochemical factors, hydrothermal processes, Barnes, H.L., 1997. Geochemistry of Hydrothermal Ore DepositsThird ed. John Wiley &
and secondary enrichment processes are reflected in the composition of Sons Inc., New York.
the corresponding magnetite. Barnes, S.J., Roeder, P.L., 2001. The range of spinel composition in terrestrial mafic and ul-
tramafic rocks. J. Petrol. 42, 2279–2302.
Barnes, S.-J., Maier, W.D., Ashwal, L.D., 2004. Platinum-group element distribution in the
Acknowledgements Main Zone and Upper Zone of the Bushveld Complex, South Africa. Chem. Geol. 208,
293–317.
Barton, M., Johnson, D., 2004. Footprints of Fe-oxide(–Cu–Au) systems. SEG 2004: Predic-
We thank the iron ore research group at UWA (CET) — Paul Duuring, tive Mineral Discovery Under Cover. Centre for Global Metallogeny, Spec. Pub., 33.
Ana-Sophie Hensler, and Warren Thorne — for kindly providing data for The University of Western Australia 112–116.
Bau, M., Alexander, B.W., 2009. Distribution of high field strength elements (Y, Zr, REE, Hf,
this project, and Steffen Hagemann for helpful feedback. Thomas
Ta, Th, U) in adjacent magnetite and chert bands and in reference standards FeR-3
Angerer thanks the LabMaTer group at UQAC, Sarah-Jane Barnes, and FeR-4 from the Temagami iron-formation, Canada, and the redox level of the
Sarah Dare, Dany Savard, Sadia Mehdi, and the CODES laser ablation Neoarchean ocean. Precambrian Res. 174, 337–346.
team at UTAS, Sarah Gilbert and Leonid Danyushevsky, for granting Bea, F., Pereira, M.D., Stroh, A., 1994. Mineral/leucosome trace-element partitioning in a
peraluminous migmatite (a laser ablation-ICP-MS study). Chem. Geol. 117, 291–312.
access to their laser systems and helping with analyses and data reduc- Beane, R.E., Titley, S.R., 1981. Porphyry copper deposits, part II. Hydrothermal alteration
tion. Further acknowledgments go to industry partners, Cliffs Natural Re- and mineralization. Econ. Geol. 235–269 (75th Anniversary).
sources, Rio Tinto Iron Ore, and BHP Billiton Iron Ore for granting access Beaudoin, G., Dupuis, C., 2009. Iron-oxide trace element fingerprinting of mineral deposit
types. In: Corriveau, L., Mumin, H. (Eds.), Exploring for Iron Oxide Copper–Gold De-
to samples. We also thank Freeport McMoRan, and in particular Richard posits: Canada and Global Analogues, Short Course Volume. Geological Association
Leveille and Clyde Leys, for granting access to sites and providing drill of Canada Annual Meeting, pp. 107–121.
hole documentation for the Ertsberg district. The Ertsberg district data Beaudoin, G., Sangster, D.F., 1992. A descriptive model for silver–lead–zinc veins in clastic
metasedimentary terranes. Econ. Geol. 87, 1005–1021.
were collected as part of CSIRO/AMIRA project P408. We acknowledge Behn, G., Camus, F., Carrasco, P., Ware, H., 2001. Aeromagnetic signature of porphyry cop-
helpful comments and suggestions by Sarah Dare, Rosaline Figueiredo per systems in northern Chile and its geologic implications. Econ. Geol. 96, 239–248.
e Silva and Erin Marsh. For helping us with the visual content of this Bhattacharya, H.N., Chakrabortym, I., Ghosh, K.K., 2007. Geochemistry of some banded
iron-formations of the Archean Supracrustals, Jharkhand–Orissa region, India.
paper, we thank Travis Naughton. Any use of trade, product, or firm J. Earth Syst. Sci. 116, 245–259.
names is for descriptive purposes only and does not imply endorsement Bordage, A., Balan, E., Villiers, J.R., Cromarty, R., Juhin, A., Carvallo, C., et al., 2011. V oxida-
by the U.S. Government. tion state in Fe–Ti oxides by high-energy resolution fluorescence-detected X-ray ab-
sorption spectroscopy. Phys. Chem. Miner. 38, 449–458.
Borisenko, L.F., Lapin, A.V., 1971. O kontsentratsiyakh elementov-primesey v
References titanomagnetite i magnetite endogennykh mestorozhdeniy razlichnykh tipov
(Trace-element concentrations in titanomagnetite and magnetite from endogene de-
Ahmad, S.N., Rose, A.W., 1980. Fluid inclusions in porphyry and skarn ore at Santa Rita. posits of various types). Dokl. Akad. Nauk SSSR 196, 1441–1444.
New Mex. Econ. Geol. 75, 229–250. Borisenko, L.F., Lapin, A.V., 1972. Trace-element concentrations in titanomagnetite and
Akimoto, S., 1955. Magnetic properties of ferromagnetic minerals contained in igneous magnetite from magmatic, contact-metasomatic and hydrothermal deposits of differ-
rocks. Jpn. J. Geophys. 1, 1–31. ent types. Dokl. Earth Sci. Sect. 196, 217–220.
Anand, R.R., Butt, C.R.M., 2010. A guide for mineral exploration through the regolith in the Bowles, J.F.W., Howie, R.A., Vaughan, D.J., Zussman, J., 2011. Rock-forming Minerals —
Yilgarn Craton, Western Australia. Aust. J. Earth Sci. 57, 1015–1114. Non-silicates: Oxides, Hydroxides and Sulphides2nd ed. Geological Society, London.
Anand, R.R., Gilkes, R.J., 1984. Mineralogical and chemical properties of weathered mag- Bragg, W.H., 1915. The structure of the spinel group of crystals. Philos. Mag. 30, 305–315.
netite grains from lateritic saprolite. J. Soil Sci. 35, 559–567. Buddington, A.F., Lindsley, D.H., 1964. Iron–titanium oxide minerals and synthetic equiv-
Andersen, D.J., Lindsley, D.H., 1988. Internally consistent solution models for Fe–Mg–Mn– alents. J. Petrol. 5, 310–357.
Ti oxides; Fe–Ti oxides. Am. Mineral. 73, 714–726. Buddington, A.F., Fahey, J., Vlisdis, A., 1955. Thermometric and petrogenetic significance
Anderson, J.L., Barth, A.P., Wooden, J.L., Mazdab, F., 2008. Thermometers and of titaniferous magnetite. Am. J. Sci. 253, 497–532.
thermobarometers in granitic systems. Rev. Mineral. Geochem. 69, 121–142. Candela, P.A., 1997. A review of shallow, ore-related granites: textures, volatiles, and ore
Angerer, T., Hagemann, S.G., 2010. The BIF-hosted high-grade iron ore deposits in the Ar- metals. J. Petrol. 38, 1619–1633.
chean Koolyanobbing Greenstone Belt, Western Australia: structural control on Carew, M.J., 2004. Controls on Cu–Au Mineralisation and Fe-oxide metasomatism in the
synorogenic- and weathering-related magnetite-, hematite-, and goethite-rich iron Eastern Fold Belt, NW Queensland, Australia. [PhD] James Cook University,
ore. Econ. Geol. 105, 917–945. Townsville.
Angerer, T., Hagemann, S.G., Danyushevsky, L.V., 2012. Geochemical evolution of the Carew, M.J., Mark, G., Oliver, N.H.S., Pearson, N., 2006. Trace element geochemistry of
banded iron formation-hosted high-grade iron ore system in the Koolyanobbing magnetite and pyrite in Fe oxide (+/−Cu–Au) mineralised systems: insights into
greenstone belt. West. Aust. Econ. Geol. 107, 599–644. the geochemistry of ore-forming fluids. Geochim. Cosmochim. Acta 70, A83.
Angerer, T., Hagemann, S., Danyushevsky, L., 2013. High-grade iron ore at Windarling, Carten, R.B., White, W.H., Stein, H.J., 1993. High-grade granite-related molybdenum sys-
Yilgarn Craton: a product of syn-orogenic deformation, hypogene hydrothermal al- tems: classification and origin. Geol. Assoc. Can. Spec. Pap. 40, 521–554.
teration and supergene modification in an Archean BIF-basalt lithostratigraphy. Min- Chacko, T., Cole, D.R., Horita, J., 2001. Equilibrium oxygen, hydrogen and carbon isotope
eral. Deposita 48, 697–728. fractionation factors applicable to geologic systems. In: Valley, J.W., Cole, D.R.
Annersten, H., 1968. A mineral chemical study of a metamorphosed iron formation in (Eds.), Stable Isotope Geochemistry. Mineralogical Society of America, Washington,
northern Sweden. Lithos 1, 374–397. pp. 1–81.
Arancibia, O.N., Clark, A.H., 1996. Early magnetite–amphibole–plagioclase alteration- Chiba, H., Chacko, T., Clayton, R.N., Goldsmith, J.R., 1989. Oxygen isotope fractionations in-
mineralization in the Island Copper porphyry copper–gold–molybdenum deposit, volving diopside, forsterite, magnetite, and calcite: application to geothermometry.
British Columbia. Econ. Geol. 91, 402–438. Geochim. Cosmochim. Acta 53, 2985–2995.
Audetat, A., Pettke, T., 2006. Evolution of a porphyry-Cu mineralized magma system at Clark, D.A., 2001. A review of geological factors that control magnetic signatures of por-
Santa Rita, New Mexico (USA). J. Petrol. 47, 2021–2046. phyry and epithermal deposits; towards predictive magnetic exploration models.
Audetat, A., Guenther, D., Heinrich, C.A., 2000. Magmatic–hydrothermal evolution in a Vietnam 2001; IAGA-IASPEI First Joint Scientific Assembly; Abstracts.
fractionating granite: a microchemical study of the Sn–W–F-mineralized mole gran- Clout, J.M.F., 2003. Upgrading processes in BIF-derived iron ore deposits: implications for
ite (Australia). Geochim. Cosmochim. Acta 64, 3373–3393. ore genesis and downstream mineral processing. Trans. Inst. Min. Metall. Sect. B Appl.
Audetat, A., Pettke, T., Heinrich, C.A., Bodnar, R.J., 2008. Special paper: the composition of Earth Sci. 112, B89–B95.
magmatic–hydrothermal fluids in barren and mineralized intrusions. Econ. Geol. 103, Cornell, R.M., Schwertmann, U., 2003. The iron oxides: Structure, properties, reactions, oc-
877–908. currences, and uses, 2 ed. Weinheim, Wiley-VCH.
30 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

Cygan, G.L., Candela, P.A., 1995. Preliminary study of gold partitioning among pyrrhotite, Garrity, C.P., Soller, D.R., 2009. Database of the geologic map of North America; adapted
pyrite, magnetite, and chalcopyrite in gold-saturated chloride solutions at 600 to from the map by J.C. Reed, Jr. and others. U. S. Geol. Surv. Data Series 424.
7008 °C, 140 MPa, 1400 bars. In: Thompson, J.F.H. (Ed.), Magmas, Fluids, and Ore De- Ghent, E., 1975. Temperature, pressure, and mixed-volatile equilibria attending metamor-
posits, Short Course. Mineralogical Association of Canada, pp. 129–137. phism of staurolite–kyanite-bearing assemblages, Esplanade Range. B. C. Geol. Soc.
Czamanske, G.K., Wones, D.R., Eichelberger, J., 1977. Mineralogy and petrology of the in- Am. Bull. 86, 1654–1660.
trusive complex of the Pliny Range, New Hampshire. Am. J. Sci. 277, 1073–1123. Ghiorso, M.S., Sack, O., 1991a. Fe–Ti oxide geothermometry: thermodynamic formulation
Dare, S.A.S., Barnes, S.-J., Beaudoin, G., 2012. Variation in trace element content of magne- and the estimation of intensive variables in silicic magmas. Contrib. Mineral. Petrol.
tite crystallized from a fractionating sulfide liquid, Sudbury, Canada: implications for 108, 485–510.
provenance discrimination. Geochim. Cosmochim. Acta 88, 27–50. Ghiorso, M.S., Sack, O., 1991b. Thermochemistry of the oxide minerals. In: Lindsley, D.H.
D'Orazio, M., Armienti, P., Cerretini, S., 1998. Phenocryst/matrix trace-element partition (Ed.), Oxide Minerals: Petrologic and Magnetic Significance. Mineralogical Society
coefficients for hawaiite–trachyte lavas from the Ellittico volcanic sequence (Mt. of America, pp. 221–264.
Etna, Sicily, Italy). Mineral. Petrol. 64, 65–88. Goldschmidt, V.M., 1954. GeochemistryOxford University Press, Oxford.
Doughty, P.T., Chamberlain, K.R., 1996. Salmon River arch revisited: New evidence for Grant, F.S., 1984/85aa. Aeromagnetics, geology and ore environments, I. Magnetite in igne-
1370 Ma rifting near the end of deposition in the Middle Proterozoic belt basin. ous, sedimentary and metamorphic rocks: an overview. Geoexploration 23, 303–333.
Can. J. Earth Sci. 33, 1037–1052. Grant, F.S., 1984/85bb. Aeromagnetics, geology and ore environments, II. Magnetite and
Drabek, M., 1982. The system Fe–Mo–S–O and its geologic application. Econ. Geol. 77, ore environments. Geoexploration 23, 335–362.
1053–1056. Grigsby, J.D., 1990. Detrital magnetite as a provenance indicator. J. Sediment. Res. 60,
Duan, C., Li, Y.H., Yuan, S.D., Hu, M.Y., Zhao, L.H., Chen, X.D., et al., 2012. Geochemical char- 940–951.
acteristics of magnetite from Washan iron deposit in Ningwu ore district and its con- Gruner, J.W., 1937. Hydrothermal leaching of iron ores of the Lake Superior type — a mod-
straints on ore-forming. Acta Petrol. Sin. 28, 243–257. ified theory. Econ. Geol. 32, 121–130.
Dudas, M.J., Harward, M.E., Schmitt, R.A., 1973. Identification of dacitic tephra by activa- Gustafson, L.B., Hunt, J.P., 1975. The porphyry copper deposit at El Salvador. Chile Econ.
tion analysis of their primary mineral phenocrysts. Quat. Res. 3, 307–315. Geol. 70, 857–912.
Dupuis, C., Beaudoin, G., 2011. Discriminant diagrams for iron oxide trace element finger- Gutzmer, J., Beukes, N.J., de Kock, M.O., Netshiozwi, S.T., 2005. Origin of high-grade iron ores
printing of mineral deposit types. Mineral. Deposita 46, 319–335. at the Thabazimbi deposit, South Africa. In: AusIMM (Ed.), Proceedings Iron Ore
Duuring, P., Hagemann, S.G., 2012. Leaching of silica bands and concentration of magne- Conference 2005, 19–21 September 2005. Freemantle, Western Australia, pp. 57–65.
tite in Archean BIF by hypogene fluids: Beebyn Fe ore deposit, Yilgarn Craton, Haggerty, S.E., 1991a. Oxide mineralogy of the upper mantle. In: Lindsley Donald, H. (Ed.),
Western Australia. Mineral. Deposita 48, 341–370. Oxide Minerals: Petrologic and Magnetic Significance. Mineralogical Society of
Duuring, P., Hagemann, S., 2013. Leaching of silica bands and concentration of magnetite America, pp. 355–416.
in Archean BIF by hypogene fluids: Beebyn Fe ore deposit, Yilgarn Craton, Western Haggerty, S.E., 1991b. Oxide textures — a mini-atlas. In: Lindsley Donald, H. (Ed.), Oxide
Australia. Mineral. Deposita 48, 341–370. Minerals: Petrologic and Magnetic Significance. Rev. Mineral. Mineral. Soc. Am.
EarthRef.org, 2013. Geochemical Earth Reference Model (GERM) Partition Coefficient pp. 129–219.
(Kd) DatabaseEarthRef.org. Hall, A.J., 1986. Pyrite–pyrrhotine redox reactions in nature. Mineral. Mag. 50, 223–229.
Einaudi, M.T., 1981a. Description of skarns associated with porphyry copper plutons. In: Halsall, T.J., 1983. In: Titley, Spencer R. (Ed.), Advances in Geology of the Porphyry
Titley, S.R. (Ed.), Advances in Geology of the Porphyry Copper Deposits. The Univer- Copper Deposits: Southwestern North America. University of Arizona Press,
sity of Arizona Press, Tucson, pp. 139–183. Tucson, p. 560.
Einaudi, M.T., 1981b. General features and origin of skarns associated with porphyry cop- Haskin, L.A., Frey, F.A., Schmitt, R.A., Smith, R.H., 1966. Meteoritic, solar and terrestrial
per plutons. In: Titley, S.R. (Ed.), Advances in Geology of the Porphyry Copper De- rare-earth distributions. Phys. Chem. Earth 7, 167–321.
posits. The University of Arizona Press, Tucson, pp. 185–209. Hayes, T.S., 1990. A preliminary study of thermometry and metal sources of the Spar Lake
Einaudi, M.T., Meinert, L.D., Newberry, R.J., 1981. Skarn deposits. Economic Geology 75th stratabound copper–silver deposit, Belt Supergroup, Montana. U.S. Geological Survey
Anniversary Volume 317–391. Open-File Report 90-484 30.
Enders MS. The Evolution of supergene enrichment in the Morenci porphyry copper de- Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation of hydro-
posit, Greenlee County, Arizona. unpub. PhD dissert., University of Arizona. 2000: thermal ore deposits. Nature 370, 519–527.
252. Heiligmann, M., Williams-Jones, A.E., Clark, J.R., 2008. The role of sulfate–sulfide-oxide–
Esperanca, S., Carlson, R.W., Shirey, S.B., Smith, D., 1997. Dating crust–mantle separation: silicate equilibria in the metamorphism of hydrothermal alteration at the Hemlo
Re–Os isotopic study of mafic xenoliths from central Arizona. Geology 25, 651–654. Gold Deposit, Ontario. Econ. Geol. 103, 335–351.
Evans, B.W., Frost, B.R., 1975. Chrome–spinel in progressive metamorphism—a prelimi- Horn, I., Foley, S.F., Jackson, S.E., Jenner, G.A., 1994. Experimentally determined
nary analysis. Geochim. Cosmochim. Acta 39, 959–972. partitioning of high field strength- and selected transition elements between spinel
Evans, K.A., McCuaig, T.C., Leach, D., Angerer, T., Hagemann, S.G., 2013. Banded iron forma- and basaltic melt. Chem. Geol. 117, 193–218.
tion to iron ore: a record of the evolution of Earth environments? Geology 41, 99–102. Huang, X.-W., Zhou, M.-F., Qi, L., Gao, J.-F., Wang, Y.-W., 2013. Re–Os isotopic ages of py-
Ewart, A., Griffin, W.L., 1994. Application of proton-microprobe data to trace-element rite and chemical composition of magnetite from the Cihai magmatic–hydrothermal
partitioning in volcanic rocks. Chem. Geol. 117, 251–284. Fe deposit, NW China. Mineral. Deposita 1–22.
Ewart, A., Bryan, W.B., Gill, J.B., 1973. Mineralogy and geochemistry of the younger volca- Huberty, J.M., Konishi, H., Heck, P.R., Fournelle, J.H., Valley, J.W., Xu, H., 2012. Silician mag-
nic islands of Tonga, S.W. Pacific. J. Petrol. 14, 429–465. netite from the Dales Gorge Member of the Brockman Iron Formation, Hamersley
Figueiredo e Silva, R.C., Lobato, L.M., Rosiere, C.A., 2008. A hydrothermal origin for the Group, Western Australia. Am. Mineral. 97, 26–37.
jaspilite-hosted giant Sierra Norte deposits in the Cajajas Mineral Province, Para State, Hutton, C.O., 1950. Studies of heavy detrital minerals. Geol. Soc. Am. Bull. 61, 635–710.
Brazil. In: Hagemann, S., Rosière, C.A., Gutzmer, J., Beukes, N.J. (Eds.), Banded Iron Ilton, E.S., Eugster, H.P., 1989. Base metal exchange between magnetite and a chloride-
Formation-related High-grade Iron Ore: Society of Economic Geologists, pp. 255–290. rich hydrothermal fluid. Geochim. Cosmochim. Acta 53, 291–301.
Figueiredo e Silva, R.C., Lobato, L.M., Hagemann, S., Danyushevsky, L., 2009. Laser-ablation Ishihara, S., 1977. Magnetite-series and ilmenite-series granitic rocks. Min. Geol. 27,
ICP-MS analyses on oxides of hypogene iron ore from the giant Serra Norte jaspilite- 293–305.
hosted iron ore deposits, Carajás Mineral Province, Brazil. In: Williams, P.J. (Ed.), Pro- Jenner, F.E., O'Neill, H.S.C., Arculus, R.J., Mavrogenes, J.A., 2010. The magnetite crisis in the
ceedings of the 10th Biennial SGA Meeting of The Society for Geology Applied to evolution of arc-related magmas and the initial concentration of Au, Ag and Cu.
Mineral Deposits Townsville Australia 17th–20th August 2009. The Society for Geol- J. Petrol. 51, 2445–2464.
ogy Applied to Mineral Deposits, Townsville, pp. 570–572. Kamvong, T., Zaw, K., Siegele, R., 2007. PIXE/PIGE microanalysis of trace elements in hy-
Fleet, M., 1981. The structure of magnetite. Acta Crystallogr. B 37, 917–920. drothermal magnetite and exploration significance: a pilot study. 15th Australian
Friedman, I., O'Neil, J.R., 1977. Compilation of stable isotope fractionation factors of geo- Conference on Nuclear and Complementary Techniques of Analysis and 9th Vacuum
chemical interest. U. S. Geol. Surv. Prof. Pap. 12 (440-KK). Society of Australia Congress. University of Melbourne, Melbourne, Australia.
Frost, B.R., 1991a. Introduction to oxygen fugacity and its petrologic importance. In: Kesler, S.E., 2005. Ore-forming fluids. Elements 1, 13–18.
Lindsley Donald, H. (Ed.), Oxide Minerals: Petrologic and Magnetic Significance. Min- Klein, C., 2005. Some Precambrian banded iron-formations (BIFs) from around the world:
eralogical Society of America, pp. 1–9. their age, geologic setting, mineralogy, metamorphism, geochemistry, and origin.
Frost, B.R., 1991b. Magnetic petrology: factors that control the occurrence of magnetite in Am. Mineral. 90, 1473–1499.
crustal rocks. In: Lindsley, D.H. (Ed.), Oxide Minerals: Petrologic and Magnetic Signif- Klemm, D.D., Henckel, J., Dehm, R.M., Von Gruenewaldt, G., 1985. The geochemistry of
icance. Rev. Mineral. Mineral. Soc. Am. , pp. 489–509. titanomagnetite in magnetite layers and their host rocks of the eastern Bushveld
Frost, B.R., 1991c. Stability of oxide minerals in metamorphic rocks. In: Lindsley, D.H. Complex. Econ. Geol. 80, 1075–1088.
(Ed.), Oxide Minerals: Petrologic and Magnetic Significance. Reviews in Minerology, Klemme, S., Günther, D., Hametner, K., Prowatke, S., Zack, T., 2006. The partitioning of
Mineralogical Society of America, pp. 469–487. trace elements between ilmenite, ulvospinel, armalcolite and silicate melts with im-
Frost, B.R., Lindsley, D.H., 1991. Occurrence of iron–titanium oxides in igneous rocks. In: plications for the early differentiation of the moon. Chem. Geol. 234, 251–263.
Lindsley, D.H. (Ed.), Petrologic and Magnetic Significance, pp. 489–509. Kretz, R., Campbell, J.L., Hoffman, E.L., Hartree, R., Teesdale, W.J., 1999. Approaches to
Frost, B.R., Lindsley Donald, H., Andersen, D.J., 1988. Fe–Ti oxide–silicate equilibria: as- equilibrium in the distribution of trace elements among the principal minerals in a
semblages with fayalitic olivine. Am. Mineral. 73, 727–740. high-grade metamorphic terrane. J. Metamorph. Geol. 17, 41–59.
Fryklund, V.C., 1964. Ore deposits of the Coeur d'Alene district, Shoshone County, Idaho. Landis, G.P., Hofstra, A.H., 1991. Fluid inclusion gas chemistry as a potential minerals ex-
U. S. Geol. Surv. Prof. Pap. 445, 103. ploration tool: case studies from Creede, CO, Jerritt Canyon, NV, Coeur d'Alene dis-
Gaetani, G.A., Grove, T.L., 1997. Partitioning of moderately siderophile elements among trict, ID and MT, southern Alaska mesothermal veins, and mid-continent MVT's.
olivine, silicate melt, and sulfide melt: constraints on core formation in the Earth J. Geochem. Explor. 42, 25–59.
and Mars. Geochim. Cosmochim. Acta 61, 1,829–1,846. Larocque, A.C.L., Stimac, J.A., McMahon, G., Jackman, J.A., Chartrand, V.P., Hickmott, D., et
Gammons, C.H., Williams-Jones, A.E., 1997. Chemical mobility of gold in the porphyry– al., 2002. Ion-microprobe analysis of Fe-Ti oxides: optimization for the determination
epithermal environment. Econ. Geol. 92, 45–59. of invisible gold. Econ. Geol. 97, 159–164.
P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32 31

LaTourrette, T.Z., Burnett, D.S., Bacon, C.R., 1991. Uranium and minor-element partitioning Mukhopadhyay, J., Gutzmer, J., Beukes, N.J., Hayashi, K.I., 2008. Stratabound magnetite de-
in Fe–Ti oxides and zircon from partially melted granodiorite, Crater Lake, Oregon. posits from the eastern outcrop belt of Archaean Iron Ore Group, Singhbhum craton,
Geochim. Cosmochim. Acta 55, 457–469. India. Appl. Earth Sci. Trans. Inst. Min. Metall. B 117, 175–186.
Leach, D.L., Landis, G.P., Hofstra, A.H., 1988. Metamorphic origin of the Coeur d'Alene base- Müller, B., Axelsson, M.D., Ohlander, B., 2003. Trace elements in magnetite from Kiruna,
and precious-metal veins in the Belt Basin, Idaho and Montana. Geology 16, 122–125. northern Sweden, as determined by LA-ICP-MS. GFF 125, 1–5.
Lee, M.J., Lee, J.I., Moutte, J., 2005. Compositional variation of Fe–Ti oxides from the Sokli Muntean, J.L., Einaudi, M.T., 2001. Porphyry–epithermal transition: Maricunga Belt,
complex, north-eastern Finland. Geosci. J. 9, 1–13. Northern Chile. Econ. Geol. 96, 743–772.
Leeman, W.P., 1974. Experimental determination of partitioning of divalent cations be- Nadoll, P., Koenig, A.E., 2011. LA-ICP-MS of magnetite: methods and reference materials.
tween olivine and basaltic liquid. University of Oregon. J. Anal. At. Spectrom. 26, 1872–1877.
Leeman, W.P., Ma, M.S., Murali, A.V., Schmitt, R.A., 1978. Empirical estimation of magne- Nadoll, P., Mauk, J.L., 2011. Wüstite in a hydrothermal silver–lead–zinc vein, Lucky Friday
tite/liquid distribution coefficients for some transition elements. Contrib. Mineral. mine, Coeur d'Alene mining district. U.S.A. Am. Mineral. 96, 261–267.
Petrol. 65, 269–272. Nadoll, P., Mauk, J.L., Leveille, R., Koenig, A.E., 2010. Geochemistry of magnetite from Cu–
Lemarchand, F., Villemant, B., Calas, G., 1987. Trace element distribution coefficients in Mo porphyry + skarn, and Climax Mo deposits in the western United States. In: Pál-
alkaline series. Geochim. Cosmochim. Acta 51, 1071–1081. Molnár, E. (Ed.), IMA 2010. Department of Mineralogy, Geochemistry and Petrology,
Lindsley, D.H., 1976a. The crystal chemistry and structure of oxide minerals as exempli- University of Szeged, Budapest, p. 507.
fied by the Fe–Ti oxides. In: Rumble III, D. (Ed.), Oxide Minerals. Rev. Mineral. Miner- Nadoll, P., Mauk, J.L., Hayes, T.S., Koenig, A.E., Box, S.E., 2012. Geochemistry of magnetite
al. Soc. Am. , pp. L1–L60. from hydrothermal ore deposits and host rocks of the Mesoproterozoic Belt Super-
Lindsley, D.H., 1976b. Experimental studies of oxide minerals. In: Rumble III, D. (Ed.), group, United States. Econ. Geol. 107, 1275–1292.
Oxide Minerals. Rev. Mineral. Mineral. Soc. Am. , pp. L61–L88. Nadoll, P., 2011. Geochemistry of magnetite from hydrothermal ore deposits and host
Lindsley, D.H., 1991. Oxide minerals: petrologic and magnetic significance. Rev. Mineral. rocks - Case studies from the Proterozoic Belt Supergroup, Cu-Mo-porphyry + skarn
Mineral. Soc. Am. 25, 509. and Climax-Mo deposits in the western United States [PhD]: University of Auckland.
Lindstrom, D.J., 1976. Experimental Study of the Partitioning of the Transition Metals Nash, W.P., Crecraft, H.R., 1985. Partition coefficients for trace elements in silicic magmas.
Between Clinopyroxene and Coexisting Silicate LiquidsUniversity of Oregon. Geochim. Cosmochim. Acta 49, 2309–2322.
Lobato, L.M., Figueiredo e Silva, R.C., Hagemann, S.G., Thorne, W.S., Zuchetti, M., 2008. Hy- Nielsen, R.L., 1992. BIGD.FOR: a FORTRAN program to calculate trace-element partition
pogene alteration associated with high-grade banded iron formation-related iron coefficients for natural mafic and intermediate composition magmas. Comput.
ore. In: Hagemann, S.G., Rosière, C.A., Gutzmer, J., Beukes, N.J. (Eds.), Banded Iron Geosci. 18, 773–788.
Formation-related High-grade Iron Ore. Reviews in Economic Geology, pp. 107–128. Nielsen, R.L., Beard, J.S., 2000. Magnetite–melt HFSE partitioning. Chem. Geol. 164, 21–34.
Longerich, H.P., Jackson, S.E., Günther, D., 1996. Laser ablation-inductively coupled Nielsen, R.L., Forsythe, L.M., Gallahan, W.E., Fisk, M.R., 1994. Major- and trace-element
plasma-mass spectrometric transient signal data acquisition and analyte concentra- magnetite–melt equilibria. Chem. Geol. 117, 167–191.
tion calculation. J. Anal. At. Spectrom. 11, 899–904. Ohmoto, H., 2003. Nonredox transformations of magnetite–hematite in hydrothermal
Lowell, J.D., Guilbert, J.M., 1970. Lateral and vertical alteration-mineralization zoning in systems. Econ. Geol. 98, 157–161.
porphyry ore deposits. Econ. Geol. 65, 373–408. Okamoto, K., 1979. Geochemical study on magmatic differentiation of Asama Volcano,
Luhr, J.F., Carmichael, I.S.E., 1980. The Colima Volcanic complex, Mexico. Contrib. Mineral. central Japan. J. Geol. Soc. Jpn. 85, 525–535.
Petrol. 71, 343–372. Otake, T., Wesolowski, D.J., Anovitz, L.M., Allard, L.F., Ohmoto, H., 2010. Mechanisms of iron
Luhr, J.F., Carmichael, I.S.E., Varekamp, J.C., 1984. The 1982 eruptions of El Chichon volca- oxide transformations in hydrothermal systems. Geochim. Cosmochim. Acta 74,
no, Chiapas, Mexico: mineralogy and petrology of the anhydrite-bearing pumices. 6141–6156.
J. Volcanol. Geotherm. Res. 23, 69–108. Pecoits, E., Gingras, M.K., Barley, M.E., Kappler, A., Posth, N.R., Konhauser, K.O., 2009. Petrog-
MacDonald, G.D., Arnold, L.C., 1994. Geological and geochemical zoning of the Grasberg raphy and geochemistry of the Dales Gorge banded iron formation: paragenetic se-
Igneous Complex, Irian Jaya, Indonesia. J. Geochem. Explor. 50, 143–178. quence, source and implications for palaeo-ocean chemistry. Precambrian Res. 172,
Mahood, G., Hildreth, W., 1983. Large partition coefficients for trace elements in high- 163–187.
silica rhyolites. Geochim. Cosmochim. Acta 47, 11–30. Pollard, P., Taylor, R., 2002. Paragenesis of the Grasberg Cu–Au deposit, Irian Jaya,
Mahood, G.A., Stimac, J.A., 1990. Trace-element partitioning in pantellerites and trachytes. Indonesia: results from logging Section 13. Mineral. Deposita 37, 117–136.
Geochim. Cosmochim. Acta 54, 2257–2276. Pollard, P.J., Taylor, R.G., Peters, L., 2005. Ages of intrusion, alteration, and mineralization
Mallmann, G., O'Neill, H.S.C., 2009. The crystal/melt partitioning of V during mantle melt- at the Grasberg Cu–Au deposit, Papua, Indonesia. Econ. Geol. 100, 1005–1020.
ing as a function of oxygen fugacity compared with some other elements (Al, P, Ca, Powell, R., Powell, M., 1977. Geothermometry and oxygen barometry using coexisting
Sc, Ti, Cr, Fe, Ga, Y, Zr and Nb). J. Petrol. 50, 1765–1794. iron–titanium oxides: a reappraisal. Mineral. Mag. 41, 257–263.
Malmqvist, D., Parasnis, D.S., 1972. Aitik: geophysical documentation of a third- Prins, P., 1972. Composition of magnetite from carbonatites. Lithos 5, 227–240.
generation copper deposit in North Sweden. Geoexploration 10, 149–160. Ramanaidou, E.R., 2009. Genesis of lateritic iron ore from banded iron-formation in the
Manske, S.L., Paul, A.H., 2002. Geology of a major new porphyry copper center in the Su- Capanema mine (Minas Gerais, Brazil). Aust. J. Earth Sci. 56, 605–620.
perior (Pioneer) district, Arizona. Econ. Geol. 97, 197–220. Ramdohr, P., 1926. Beobachtungen an Magnetit, Ilmenit, Eisenglanz, und Uberlegungen
Mathur, R., Ruiz, J., Titley, S., Gibbins, S., Margotomo, W., 2000. Different crustal sources iiber das System FeO–Fe2O3–TiO2. Neues Jahrb. Mineral. 54A, 320–379.
for Au-rich and Au-poor ores of the Grasberg Cu–Au porphyry deposit. Earth Planet. Ramdohr, P., 1955. Die Erzminerale und ihre VerwachsungenAkademie-Verlag, Berlin.
Sci. Lett. 183, 7–14. Ray, G.E., Webster, I.C.L., 2007. Geology and chemistry of the low Ti magnetite-bearing
McClenaghan, M.B., 2005. Indicator mineral methods in mineral exploration. Geochem. Heff Cu–Au skarn and its associated plutonic rocks, Heffley Lake, South-Central Brit-
Explor. Environ. Anal. 5, 233–245. ish Columbia. Explor. Min. Geol. 16, 159–186.
McEnroe, S.A., Robinson, P., Panish, P.T., 2001. Aeromagnetic anomalies, magnetic petrol- Razjigaeva, N.G., Naumova, V.V., 1992. Trace element composition of detrital magnetite
ogy, and rock magnetism of hemo-ilmenite- and magnetite-rich cumulate rocks from from coastal sediments of northwestern Japan Sea for provenance study.
the Sokndal Region, South Rogaland, Norway. Am. Mineral. 86, 1447–1468. J. Sediment. Petrol. 62, 802–809.
McIntire, W.L., 1963. Trace element partition coefficients—a review of theory and applica- Reguir, E.P., Chakhmouradian, A.R., Halden, N.M., Yang, P., 2008. Early magmatic and
tions to geology. Geochim. Cosmochim. Acta 27, 1209–1264. reaction-induced trends in magnetite from the carbonatites of Kerimasi, Tanzania.
McLemore, V.T., Munroe, E.A., Heizler, M.T., McKee, C., 1999. Geochemistry of the Copper Can. Mineral. 46, 879–900.
Flat porphyry and associated deposits in the Hillsboro mining district, Sierra County, Reid, F., 1983. Origin of the rhyolitic rocks of the Taupo Volcanic Zone, New Zealand.
New Mexico, USA. J. Geochem. Explor. 67, 167–189. J. Volcanol. Geotherm. Res. 15, 315–338.
McQueen, K.G., Cross, A.J., 1998. Magnetite as a geochemical sampling medium: applica- Righter, K., Leeman, W.P., Hervig, R.L., 2006a. Partitioning of Ni, Co and V between spinel-
tion to skarn deposits. In: Eggleton, R.A. (Ed.), The State of the Regolith. Geological So- structured oxides and silicate melts: importance of spinel composition. Chem. Geol.
ciety of Australia, Brisbane, pp. 194–199. 227, 1–25.
McQueen, K.G., Whitbread, M.A., 2002. Mineralogical exploration: using element–host Righter, K., Sutton, S.R., Newville, M., Le, L., Schwandt, C.S., Uchida, H., et al., 2006b. An ex-
mineral associations in the search for ore. In: Roach, I.C. (Ed.), Regolith and Land- perimental study of the oxidation state of vanadium in spinel and basaltic melt with
scapes in Eastern Australia. CRC LEME, pp. 96–99. implications for the origin of planetary basalt. Am. Mineral. 91, 1643–1656.
Meinert, L.D., 1987. Skarn zonation and fluid evolution in the Groundhog Mine, Central Robinson, R.F., Cook, A., 1966. The Safford copper deposit, Lone Star mining district, Gra-
mining district, New Mexico. Econ. Geol. 82, 523–545. ham County, Arizona. In: Titley, S.R., Hicks, C.L. (Eds.), Geology of the Porphyry Cop-
Meinert, L.D., Hefton, K.K., Mayes, D., Tasiran, I., 1997. Geology, zonation, and fluid evolu- per Deposits, Southwestern North America. University of Arizona Press, Tucson,
tion of the Big Gossan Cu–Au skarn deposit, Ertsberg District, Irian Jaya. Econ. Geol. pp. 251–266.
92, 509–534. Roedder, E., Bodnar, R., 1997. Fluid inclusion studies of hydrothermal ore deposits. In:
Meinert, L.D., Dipple, G.M., Nicolescu, S., 2005. World skarn deposits. Economic Geology Barnes, H. (Ed.), Geochemistry of Hydrothermal Ore Deposits. John Wiley, New
100th Anniversary Volume. 299–336. York, pp. 657–698.
Mertig, H.J., Rubin, J.N., Kyle, J.R., 1994. Skarn Cu–Au orebodies of the Gunung Bijih Rose, A.W., Baltosser, W.W., 1966. The porphyry copper deposit at Santa Rita, New Mexico.
(Ertsberg) district, Irian Jaya, Indonesia. J. Geochem. Explor. 50, 179–202. In: Titley, S.R., Hicks, C.L. (Eds.), Geology of the Porphyry Copper Deposits — South-
Mollo, S., Putirka, K., Iezzi, G., Scarlato, P., 2013. The control of cooling rate on western North America. The University of Arizona Press, Tucson, pp. 205–220.
titanomagnetite composition: implications for a geospeedometry model applicable Rosière, C.A., Rios, F.J., 2004. The origin of hematite in high-grade iron ores based on infra-
to alkaline rocks from Mt. Etna volcano. Contrib. Mineral. Petrol. 165, 457–475. red microscopy and fluid inclusion studies: the example of the Conceição mine,
Morris, R.C., 1985. Genesis of iron ore in banded iron-formation by supergene and Quadrilátero Ferrífero, Brazil. Econ. Geol. 99, 611–624.
supergene-metamorphic processes — a conceptual model. In: Wolf, K.H. (Ed.), Hand- Rosière, C.A., Spier, C.A., Rios, F.J., Suckau, V.E., 2008. The itabirites of the Quadrilátero
book of Strata-bound and Stratiform Ore Deposits. Elsevier, Amsterdam, pp. 73–235. Ferrífero and related high-grade iron ore deposits: an overview. In: Hagemann, S.G.,
Mücke, A., Raphael Cabral, A., 2005. Redox and nonredox reactions of magnetite and he- Rosière, C.A., Gutzmer, J., Beukes, N.J. (Eds.), Banded Iron Formation-related High-
matite in rocks. Chem. Erde Geochem. 65, 271–278. grade Iron Ore, pp. 223–254.
32 P. Nadoll et al. / Ore Geology Reviews 61 (2014) 1–32

Rumble, D., 1976a. Oxide minerals. Rev. Mineral.: Mineral. Soc. Am., Short Course Notes 3, Tauson, V.L., Babkin, D.N., Pastushkova, T.M., Krasnoshchekova, T.S., Lustenberg, E.E.,
706. Belozerova, O.Y., 2011. Dualistic distribution coefficients of elements in the system min-
Rumble, D., 1976b. Oxide minerals in metamorphic rocks. In: Rumble III, D. (Ed.), Oxide eral–hydrothermal solution. I. Gold accumulation in pyrite. Geochem. Int. 49, 568–577.
Minerals. Rev. Mineral. Mineral. Soc. Am. , pp. R1–R20. Tauson, V.L., Babkin, D.N., Pastushkova, T.M., Akimov, V.V., Krasnoshchekova, T.S., Lipko,
Rusk, B.G., Reed, M.H., Dilles, J.H., Klemm, L.M., Heinrich, C.A., 2004. Compositions of mag- S.V., et al., 2012. Dualistic distribution coefficients of elements in the system
matic hydrothermal fluids determined by LA-ICP-MS of fluid inclusions from the por- mineral–hydrothermal solution. II. Gold in magnetite. Geochem. Int. 50, 227–245.
phyry copper–molybdenum deposit at Butte, MT. Chem. Geol. 210, 173–199. Taylor, D., Dalstra, H.J., Harding, A.E., Broadbent, G.C., Barley, M.E., 2001. Genesis of high-
Rusk, B.G., Reed, M.H., Dilles, J.H., 2008. Fluid inclusion evidence for magmatic–hydrother- grade hematite orebodies of the Hamersley province, Western Australia. Econ. Geol.
mal fluid evolution in the porphyry copper–molybdenum deposit at Butte, Montana. Bull. Soc. Econ. Geol. 96, 837–873.
Econ. Geol. 103, 307–334. Thorne, W., Hagemann, S., Barley, M., 2004. Petrographic and geochemical evidence for
Rusk, B.G., Oliver, N., Brown, A., Lilly, R., Jungmann, D., 2009. Barren magnetite breccias in hydrothermal evolution of the North Deposit, Mt Tom Price, Western Australia. Min-
the Cloncurry region, Australia; comparisons to IOCG deposits. In: Williams, P.J. (Ed.), eral. Deposita 39, 766–783.
Proceedings of the 10th Biennial SGA Meeting of the Society for Geology Applied to Thorne, W.S., Hagemann, S.G., Webb, A., Clout, J., 2008. Banded iron formation-related
Mineral Deposits. The Society for Geology Applied to Mineral Deposits, Townsville iron ore deposits of the Hamersley Province, Western Australia. In: Hagemann, S.G.,
Australia, pp. 656–658. Rosière, C.A., Gutzmer, J., Beukes, N.J. (Eds.), Banded Iron Formation-related High-
Rusk, B., Oliver, N., Cleverley, J., Blenkinsop, T., Zhang, D., Williams, P., et al., 2010. Physical grade Iron Ore, pp. 197–222.
and chemical characteristics of the Ernest Henry iron oxide copper gold deposit, Titley, S.R., 1981. Advances in geology of the porphyry copper deposits of southwestern
Australia; implications for IOGC genesis. In: Porter, T.M. (Ed.), Hydrothermal Iron North AmericaThe University of Arizona Press, Tucson.
Oxide Copper–Gold and Related Deposits: A Global Perspective. PGC Publishing, Ad- Titley, S.R., Beane, R.E., 1981. Porphyry copper deposits, part I. Geologic settings, petrolo-
elaide, pp. 201–218. gy, and tectogenesis. Econ. Geol. 214–235 (75th Anniversary).
Ryabchikov, I.D., Kogarko, L.N., 2006. Magnetite compositions and oxygen fugacities of the Togashi, S., Terashima, S., 1997. The behavior of gold in unaltered island arc tholeiitic
Khibina magmatic system. Lithos 91, 35–45. rocks from Izu–Oshima, Fuji, and Osoremaya Volcanic Areas, Japan. Geochim.
Sack, R.O., Lichtner, P.C., 2009. Constraining compositions of hydrothermal fluids in Cosmochim. Acta 61, 543–554.
equilibrium with polymetallic ore-forming sulfide assemblages. Econ. Geol. 104, Toplis, M.J., Carroll, M.R., 1995. An experimental study of the influence of oxygen fugacity
1249–1264. on Fe–Ti oxide stability, phase relations, and mineral–melt equilibria in ferro-basaltic
Sauerzapf, U., Lattard, D., Burchard, M., Engelmann, R., 2008. The titanomagnetite–ilmenite systems. J. Petrol. 36, 1137–1170.
equilibrium: new experimental data and thermo-oxybarometric application to the Toplis, M., Corgne, A., 2002. An experimental study of element partitioning between mag-
crystallization of basic to intermediate rocks. J. Petrol. 49, 1161–1185. netite, clinopyroxene and iron-bearing silicate liquids with particular emphasis on
Savard, D., Barnes, S.-J., Sunder Raju, P.V., 2010. Accurate LA-ICP-MS calibration for magnetite vanadium. Contrib. Mineral. Petrol. 144, 22–37.
analysis using multiple reference materials. Goldschmidt Conference Abstracts A914. Trendall, A.F., Blockley, J.G., 1970. The iron formations of the Precambrian Hamersley
Sawkins, F.J., Scherkenbach, D.A., 1981. High copper content of fluid inclusions in quartz Group, Western Australia; with special reference to the crocidolite. Geol. Surv.
from northern Sonora: implications for ore genesis theory. Geology 9, 37–40. West. Aust. Bull. 119 (365 pp.).
Schilling, J., Frost, B.R., Marks, M.A.W., Wenzel, T., Markl, G., 2011. Fe–Ti oxide–silicate Triebold, S., von Eynatten, H., Luvizotto, G.L., Zack, T., 2007. Deducing source rock litholo-
(QUIlF-type) equilibria in feldspathoid-bearing systems. Am. Mineral. 96, 100–110. gy from detrital rutile geochemistry: an example from the Erzgebirge, Germany.
Seedorff, E., Einaudi, M.T., 2004. Henderson porphyry molybdenum system, Colorado: I. Chem. Geol. 244, 421–436.
Sequence and abundance of hydrothermal mineral assemblages, flow paths of evolv- Ulrich, T., Gunther, D., Heinrich, C.A., 1999. Gold concentrations of magmatic brines and
ing fluids, and evolutionary style. Econ. Geol. 99, 3–37. the metal budget of porphyry copper deposits. Nature 399, 676.
Seedorff, E., Dilles, J.H., Proffett Jr., J.M., Einaudi, M.T., Zurcher, L., Stavast, W.J.A., et al., Ulrich, T., Gunther, D., Heinrich, C.A., 2001. The evolution of a porphyry Cu–Au deposit,
2005. Porphyry deposits: characteristics and origin of hypogene features. Econ. based on LA-ICP-MS analysis of fluid inclusions: Bajo de la Alumbrera, Argentina.
Geol. 251–298 (100th Anniversary Volume). Econ. Geol. 96, 1743–1774.
Shannon, R., 1976. Revised effective ionic radii and systematic studies of interatomic dis- Van Baalen, M.R., 1993. Titanium mobility in metamorphic systems: a review. Chem. Geol.
tances in halides and chalcogenides. Acta Crystallogr. A 32, 751–767. 110, 233–249.
Shannon, J.R., Nelson, E.P., Golden Jr., R.J., 2004. Surface and underground geology of the Verlaguet, A., Brunet, F., Goffé, B., Murphy, W.M., 2006. Experimental study and modeling
world-class Henderson molybdenum porphyry mine, Colorado. Geol. Soc. Am. Field of fluid reaction paths in the quartz–kyanite ± muscovite–water system at 0.7 GPa
Guid. 5. in the 350–550 °C range: implications for Al selective transfer during metamorphism.
Shcheka, S.A., Platkov, A.V., Vrzhosek, A.A., Levashov, G.B., Oktyabrsky, R.A., 1980. The Geochim. Cosmochim. Acta 70, 1772–1788.
trace element paragenesis of magnetite. Nauka 147. Villemant, B., Jaffrezic, H., Joron, J.-L., Treuil, M., 1981. Distribution coefficients of major
Shcheka, S.A., Zabelin, V.V., Chubarov, V.M., 1982. Magnetites and ferric hydroxides in and trace elements; fractional crystallization in the alkali basalt series of Chaîne des
sediments of the Japan and Philippine seas and their genetic information. Mar. Puys (Massif Central, France). Geochim. Cosmochim. Acta 45, 1997–2016.
Geol. 45, M23–M29. Vincent, E.A., Phillips, R., 1954. Iron–titanium oxide minerals in layered gabbros of the
Shcheka, S.A., Naumova, V.V., Wrzosek, A.A., 1988. Trace-element paragenesis in hydro- Skaergaard intrusion, East Greenland: part I. Chemistry and ore-microscopy.
thermal magnetite as indicators of the origin and ore potential of mineralization. Geochim. Cosmochim. Acta 6, 1–4 (IN1–IN2, 5–26).
Dokl. Akad. Nauk SSSR 294, 958–962. Wang, R., Yu, A.P., Chen, J., Xie, L., Lu, J.-J., Zhu, J.-C., 2012. Cassiterite exsolution with il-
Sidhu, P.S., Gilkes, R.J., Posner, A.M., 1978. The synthesis and some properties of Co, Ni, Zn, menite lamellae in magnetite from the Huashan metaluminous tin granite in south-
Cu, Mn and Cd substituted magnetites. J. Inorg. Nucl. Chem. 40, 429–435. ern China. Mineral. Petrol. 105, 71–84.
Sidhu, P.S., Gilkes, R.J., Posner, A.M., 1981. Oxidation and ejection of nickel and zinc from Waychunas, G.A., 1991. Crystal chemistry of oxides and oxyhydroxides. In: Lindsley, D.H.
natural and synthetic magnetites. Soil Sci. Soc. Am. J. 45, 641–644. (Ed.), Oxide Minerals: Petrologic and Magnetic Significance. Mineralogical Society of
Sillitoe, R.H., 2000. Gold-rich porphyry deposits: descriptive and genetic models and their America, pp. 11–61.
role in exploration and discovery. Rev. Econ. Geol. 13, 315–345. Wechsler, B.A., Lindsley, D.H., Prewitt, C.T., 1984. Crystal structure and cation distribution
Sillitoe, R.H., 2003. Iron oxide–copper–gold deposits: an Andean view. Mineral. Deposita in titanomagnetites (Fe3 − xTixO4). Am. Mineral. 69, 754–770.
38, 787–812. Whalen, J.B., Chappel, B.W., 1988. Opaque mineralogy and mafic mineral chemistry
Sillitoe, R.H., 2010. Porphyry copper systems. Econ. Geol. 105, 3–41. of I- and S-type granites of the Lachlan fold belt, southeast Australia. Am. Mineral.
Simon, A.C., Candela, P.A., Piccoli, P.M., Mengason, M., Englander, L., 2008. The effect of 73, 281–296.
crystal–melt partitioning on the budgets of Cu, Au, and Ag. Am. Mineral. 93, 1437–1448. White, A.F., Peterson, M.L., Hochella, M.F., 1994. Electrochemistry and dissolution kinetics
Singoyi, B., Zaw, K., 2001. A petrological and fluid inclusion study of magnetite–scheelite of magnetite and ilmenite. Geochim. Cosmochim. Acta 58, 1859–1875.
skarn mineralization at Kara, Northwestern Tasmania: implications for ore genesis. Whitney, D.L., Evans, B.W., 2010. Abbreviations for names of rock-forming minerals. Am.
Chem. Geol. 173, 239–253. Mineral. 95, 185–187.
Singoyi, B., Danyushevsky, L., Davidson, G.J., Large, R., Zaw, K., 2006. Determination of trace Williams, P.J., Barton, M.D., Johnson, D.A., Fontboté, L., 2005. Iron oxide copper–gold de-
elements in magnetites from hydrothermal deposits using the LA-ICP-MS technique. posits: geology, space–time distribution, and possible modes of origin. Econ. Geol.
Abstracts of Oral and Poster Presentations from the SEG 2006 ConferenceSociety of Bull. Soc. Econ. Geol. 100th Anniversary, 371–405.
Economic Geologists, Keystone, USA 367–368. Wones, D.R., 1989. Significance of the assemblage titanite + magnetite + quartz in gra-
Skublov, S., Drugova, G., 2003. Patterns of trace-element distribution in calcic amphiboles nitic rocks. Am. Mineral. 74, 744–749.
as afunction of metamorphic grade. Can. Mineral. 41, 383–392. Yardley, B.W.D., 2005. Metal concentrations in crustal fluids and their relationship to ore
Spong, P.L., 1998. Geochemistry of magnetite from convergent-margin plutonic rocks of formation, 100th anniversary special paper. Econ. Geol. 100, 613–632.
AustraliaUniversity of Auckland, Japan and New Zealand [MSc]. Zack, T., von Eynatten, H., Kronz, A., 2004. Rutile geochemistry and its potential use in
Stimac, J., Hickmott, D., 1994. Trace-element partition-coefficients for ilmenite, ortho- quantitative provenance studies. Sediment. Geol. 171, 37–58.
pyroxene and pyrrhotite in rhyolite determined by micro-pixel analysis. Chem. Zhang, D., Rusk, B., Oliver, N., 2009. Trace elements in sulfides and magnetite from the
Geol. 117, 313–330. Ernest Henry Iron Oxide–Copper–Gold deposit, Australia. Geological Society of
Streck, M.J., Grunder, A.L., 1997. Compositional gradients and gaps in high-silica rhyolites America Annual Meeting, Portland, Oregon, 18–21 October 2009. , 41, p. 85.
of the Rattlesnake Tuff, Oregon. J. Petrol. 38, 133–163. Zheng, Y.-F., Simon, K., 1991. Oxygen isotope fractionation in hematite and magnetite; a
Tacker, R.C., Candela, P.A., 1987. Partitioning of molybdenum between magnetite and theoretical calculation and application to geothermometry of metamorphic iron-
melt; a preliminary experimental study of partitioning of ore metals between silicic formations. Eur. J. Mineral. 3, 877–886.
magmas and crystalline phases. Econ. Geol. 82, 1827–1838.

You might also like