You are on page 1of 27

13.

20

Iron Oxide(CuAuREEPAgUCo) Systems

MD Barton, University of Arizona, Tucson, AZ, USA


2014 Elsevier Ltd. All rights reserved.

13.20.1
Introduction
13.20.1.1
Semantics and Postulated Origins
13.20.2
Geologic Context for IOCG Systems
13.20.2.1
Distribution in Space and Time
13.20.2.2
Geologic Settings
13.20.2.2.1
Association with igneous rocks (or lack thereof)
13.20.2.2.2
Framework lithologies and paleoclimate
13.20.3
Synopsis of Deposit Features
13.20.3.1
Deposit Types
13.20.3.1.1
Magnetite- and/or hematite-dominated deposits
13.20.3.1.2
Fe oxide-poor Cu(Au/Ag) deposits of proposed affinity to IOCG systems
13.20.3.1.3
Possible modern analogues
13.20.3.2
Grade, Size, and Form
13.20.3.3
Ore Mineralogy and Paragenesis
13.20.3.4
Minor Element Contents and Mineralogy: UTh, REE, CoNiV, ClFBr, B
13.20.4
Hydrothermal Alteration and System-scale Zoning
13.20.4.1
Types of Hydrothermal Alteration
13.20.4.1.1
Sodic to calcic alteration types
13.20.4.1.2
Carbonate-hosted alteration: Skarn and Fe oxide replacement
13.20.4.1.3
K-rich alteration: High-temperature and low-temperature types
13.20.4.1.4
Hydrolytic (acid) alteration
13.20.4.2
System- to Regional-Scale Spatial and Temporal Patterns
13.20.4.3
Extent of Metasomatism and Comparison with Porphyry/Alkaline Cu Systems
13.20.5
Petrologic and Geochemical Characteristics
13.20.5.1
Conditions of Formation
13.20.5.1.1
Depth
13.20.5.1.2
Temperature
13.20.5.1.3
Fluid inclusion compositions
13.20.5.1.4
Oxidation state, sulfidation state, and total sulfur
13.20.5.2
Tracer Studies and Sources of Components
13.20.5.2.1
Light stable isotopes: H, O, S, C, and B
13.20.5.2.2
Radiogenic isotopes: Sr, Nd, Pb, and Os
13.20.5.2.3
Halogens and noble gases: Ratios and isotopes
13.20.6
Summary of the IOCG Clan, Likely Origins, and the terrestrial Hydrothermal Environment
13.20.6.1
Discussion of Genesis
13.20.6.2
IOCG Systems and the Terrestrial Hydrothermal Environment
Acknowledgments
References

13.20.1

Introduction

The bulk composition (i.e., geochemistry) of the mineralization defines the iron oxide(CuAuREEPAgUCo) clan.
The numerous and geologically diverse deposits of this group
contain >10% of low-Ti Fe oxides in combination with geochemically elevated contents of Cu, Au, REE, P, U, Ag, and Co.
They tend to be structurally or stratigraphically controlled and
they are temporally and spatially associated with intense and
voluminous NaCaK metasomatism. When considered as a
group, they lack the well-defined tectonic and igneous correlations that are characteristic of well-understood magmatichydrothermal systems. These distinctive traits challenge us to

Treatise on Geochemistry 2nd Edition

515
516
517
517
518
518
519
519
519
520
522
522
522
523
524
525
525
525
526
526
527
527
528
529
529
529
530
530
531
531
532
533
534
534
535
536
536
536

understand what geologic processes and settings lead to this


geochemical specialization and, more generally, to see how
this family of deposits fits in the broad spectrum of hydrothermal systems.
Although ores of this kind have been known for centuries
and mined for iron, copper, and other commodities, it was not
until the 1980s that the fundamental similarities of these disparate deposits were pointed out (e.g., Hauck, 1990; Hitzman
et al., 1992; Meyer, 1988). This recognition followed a resurgence of interest in the copper-rich variants motivated by the
discovery in 1975 of the Olympic Dam deposit in South
Australia, which is currently the worlds largest U resource
and one of the largest Cu and Au deposits. Previously, and to

http://dx.doi.org/10.1016/B978-0-08-095975-7.01123-2

515

516

Iron Oxide(CuAuREEPAgUCo) Systems

some extent since then, a variety of names and contrasting


classifications have been applied to these deposits (Cox and
Singer, 2007). Beginning about 2000 (e.g., Porter, 2000), this
group has been termed iron oxide(coppergold) (IOCG) after
the several possible principal commodities. Divisions within
the IOCG clan have been proposed by a number of authors
based on economic metals, deposit mineralogy, and possible
origins (e.g., Williams, 2010a; Williams et al., 2005).
This term is useful because it unifies a widespread and
formerly ill-defined group of deposits that share distinctive
chemical and geologic characteristics (Hitzman et al., 1992).
These include a link to moderate- to high-temperature brines
but which lack globally coherent tectonic or magmatic correlations (Barton and Johnson, 1996). The definition given
above emphasizes a characteristic set of element enrichments
(not ore grades) coupled with voluminous hypogene iron
oxides. This approach is more inclusive than other definitions
that have specified temporal, tectonic, economic, or magmatic
attributes (e.g., Groves et al., 2010; Hitzman et al., 1992;
Meyer, 1988; Williams et al., 2005). Nevertheless, it still serves
to distinguish this group from sedimentary iron deposits,
magmatic irontitanium oxide deposits, and other iron
oxide-bearing hydrothermal deposits, such as many members
of the porphyry copper clan (Chapter 13.14). In particular,

Magma-derived

this usage does not restrict IOCG to those deposits containing


economic concentrations of copper and(or) gold (cf. Williams
et al., 2005), economics being a factor that reflects a societal
rather than geologic distinction even within a single hydrothermal system.

13.20.1.1

Semantics and Postulated Origins

Debate arises both from semantics (definitions and what to


include) and from contrasting interpretations of origin.
Various workers have included deposits lacking voluminous
iron oxides, notably magnetite-bearing porphyry Cu(Au)
deposits, Cu(Ag) deposits hosted in volcanic terranes with
accessory hematite, and some pyrrhotite  pyrite Cu(Au)
deposits that occur in the same areas as the magnetite/hematite-dominated deposits (see Haynes, 2000; Williams, 2010a).
Connections between these other FeCuAu-bearing ore types
and the IOCG clan as defined here remain the topic of ongoing
discussion and reflect uncertainties in both geology and genesis.
The geologic diversity (in contrast to the geochemical similarity) of IOCG systems has contributed to multiple genetic
hypotheses (Figure 1): (1) magmatic-hydrothermal wherein
the key aqueous fluids are of magmatic origin, (2) terrestrial
hydrothermal wherein key fluids are basinal or surficial

Surface- or basin-derived

Surface or
basin-derived
waters

Metamorphic-derived

Surface-derived
brines
Cu(Au)
Fe-oxide

Surface-derived
waters
Fe-Oxide
Cu(Au)
(depends
on trap)

Cu(Au)
Fe-oxide

Early, barren
Fe oxide

Regional
Na(Ca)

Deep basinal
fluids

Basinal
fluids

Local magmatic
K-silicate alteration
Magmatic
fluids

Older,
Cl-rich
rocks

Metamorphic fluids

Magmatic
fluids

(generated from or salinity added by


passing through Cl-rich rocks)

Specialized magmas
(alkaline or otherwise)

Cu(Au) mineralization
Cpy/Bn Hm/Mt Py

Fe-oxide mineralization
Mt(Ap)/Hm Py Cpy

Igneous heat cation source


(other heat sources possible)
H+ alteration
Ser(Mu)/Chl + Qz + Hm

Na(Ca) alteration
Na plag/Scap + Cpx/Act/Chl

Tectonic/metamorphic drive
for fluid production and flow

K alteration (type I)
Biot/Kfsp Mt/Hm + Act/Cpx

K alteration (type II)


Kfsp + Hm (Biot + Mt)

Figure 1 Alternative hydrothermal origins and architectures for IOCG systems illustrating possible fluid sources, paths, and distribution of alteration
and ores (modified from Barton MD and Johnson DA (2004) Footprints of Fe-oxide (CuAu) systems: SEG 2004 Predictive Mineral Discovery
Under Cover. Centre for Global Metallogeny, University of Western Australia Special Publication 33: 112116). Deposits show far more varied
geometries that are illustrated here. Multiple fluid sources are possible in all cases. (a) Magmatic source implies distinctive composition and proximity to
source; regional Na(Ca) is coincidental. (b) Evaporitic source implies coeval or older brine source, necessary but indifferent to type of heat source, and
available upper crustal plumbing. (c) Metamorphic source implies metaevaporitic (or conceivably mantle) Cl source with regional plumbing. A fourth
hypothesis involving immiscible Fe oxideP-rich liquids is not illustrated (see text) but has some parallels with (a).

Iron Oxide(CuAuREEPAgUCo) Systems

nonmagmatic brines circulated by igneous or crustal heat, (3)


metamorphic-hydrothermal wherein fluids are derived from
distinctive crustal sources by metamorphic devolatilization
and waterrock interaction at depth, and (4) fundamentally
magmatic wherein the key ore-forming fluid is an immiscible
volatile-bearing iron oxide melt. A further possibility, not
developed here, is that some occurrences represent the serendipitous superposition of CuAu mineralization on older,
genetically unrelated ironstones (e.g., Davidson and Large,
1994; Mark et al., 2006a; Skirrow and Walshe, 2002). In all
hydrothermal models, involvement of a second, near-surface
fluid is likely and may either mix with a deeper-sourced fluid or
be a later overprint; in many cases, this second fluid may be
necessary to supply sulfur for ore formation and might also
introduce some metals (Figure 1; e.g., Gow et al., 1994; Haynes
et al., 1995; Williams et al., 2010).
Characteristics and interpretations of IOCG systems have
been presented in many papers beginning with the early synthesis papers (Barton and Johnson, 1996; Einaudi and Oreskes,
1990; Hauck, 1990; Hitzman et al., 1992), continuing through
the comprehensive volumes edited by Porter (2000, 2002,
2010b), various other compilations (e.g., Corriveau and
Mumin, 2010; Skirrow and Davidson, 2007; Yang et al.,
2011), and the global review by Williams et al. (2005). These
publications are but a handful of over 300 papers and theses
with an IOCG focus that have been published from 2000 to
2011. Most earlier reviews have either taken a systematic review
of deposit types or focused on geochemical evidence and
processes.
This chapter reviews the nature of the IOCG clan based on
the extensive geologic and geochemical studies on these systems

517

published since the late 1980s. It briefly compares IOCG systems and related deposits to other hydrothermal ore-forming
systems and considers alternative origins in light of available
evidence. It concludes that IOCG systems comprise a diverse
family in which many but not all members form as shallow
terrestrial, brine-dominated hydrothermal systems and that
IOCG systems have compelling analogies in other terrestrial
hydrothermal ore-forming environments. (Hydrothermal fluids
are derived from magmatic, metamorphic, marine, and a compositionally diverse group of near-surface terrestrial (including
basinal) sources. Dominance by the last group defines terrestrial
hydrothermal (i.e., T > 200  C) systems as used here.)

13.20.2

Geologic Context for IOCG Systems

The variability in IOCG systems merits an overview of their


temporal, spatial, and geologic settings before moving to
deposit features and system-scale patterns.

13.20.2.1

Distribution in Space and Time

IOCG provinces span the globe and range in age from Archean
to Cenozoic and include several modern analogues (Figure 2).
Except for Antarctica, districts on all continents have produced
Cu and(or) Fe with or without other commodities. Nearly all
regions have both Cu-poor and Cu-rich deposits, although the
proportions vary greatly. Economically, the current most important regions, in order of age, are the central Andes (Cu Fe;
Mesozoic; Sillitoe, 2003), the lower Yangtze (Fe  Cu; Mesozoic;
Pan and Dong, 1999; Yang et al., 2011); the southern Urals

Figure 2 Global distribution of IOCG provinces showing ages, representative abundances of sodic(calcic), potassic alteration, and economic copper
(gold). See Williams et al. (2005) for references and deposits.

518

Iron Oxide(CuAuREEPAgUCo) Systems

(Fe; Paleozoic; Hawkins et al., 2010), South Australia


(Cu U Au; Mesoproterozoic; Hayward and Skirrow, 2010;
Skirrow and Davidson, 2007), north-central Australia
(Cu Au; Paleo to Mesoproterozoic; Williams, 1998; Williams
and Pollard, 2001), N Baltic (Fe  Cu; Paleoproterozoic;
Billstrom et al., 2010; Frietsch et al., 1997), and Carajas
(Cu Au; Late Archean; Xavier et al., 2010).
There is an irregular distribution in time: large Cu-rich
deposits occur widely in only a few regions (Carajas, South
Australia, central Andes) consistent with their overall scarcity.
However, the evidence (summarized in Barton and Johnson,
1996; Williams et al., 2005) does not support a fundamental
secular variation, such as a Proterozoic (or Precambrian) age as
originally proposed by Meyer (1988) and advocated by some
later workers (e.g., Groves et al., 2010). Phanerozoic examples
are numerous and exhibit the same general characteristics as
Precambrian systems. Hitzman (2000) and others have
suggested that Cu-poor, magnetite-rich deposits tend to be
older than Cu-rich deposits within individual regions, yet a
number of hydrothermal systems contain both Cu-rich and
Cu-poor, and there are ample counterexamples to a simple
pattern reflecting the protracted evolution in major regions
(e.g., Cloncurry, Mark et al., 2006a,b; Oliver et al., 2008; western
United States, Barton et al., 2011a). Thus, although distinctions
between provinces exist, the differences reflect many factors
including nature of the host rocks (e.g., more abundant carbonates in the Phanerozoic), nature of magmatism, levels of exposure, and later events (e.g., superimposed metamorphism). The
possible roles of these factors are noted later.
IOCG occurrences are far more numerous than the handful
of large systems that have drawn most economic and research
attention. Many regions have hundreds of distinct occurrences
compared to the few major deposits in the same areas. IOCG
systems represent a family of common processes that have
recurred over much of geologic time.

13.20.2.2

Geologic Settings

Regional settings are geologically active yet tectonically diverse;


and, with few exceptions, they have broadly coeval but compositionally varied magmatism. The host terrains are moderately to
highly oxidized with magnetite- to hematite-stable rocks
(Haynes, 2000), and most present evidence for coeval or older
evaporitic materials that is, surface/sedimentary brines and/or
(meta)evaporites (Barton and Johnson, 1996, 2000). Many continental tectonic settings can host economically significant Fe
oxide  Cu(Au) mineralization; in contrast, marine arcs and
ridges lack evidence for IOCG-type mineralization. Extensional
settings in arcs and rifts are most common (Hitzman et al., 1992),
although compressional settings appear to be widespread particularly for metamorphic-affiliated systems.
During the Phanerozoic, IOCG systems formed in compressional (Andes) and extensional arcs (Andes, southwestern
North America, Urals), back arcs (western United States,
Yangtze), intraplate (Siberia), and rift settings (Mesozoic eastern North America, northeast Africa, Salton Sea). Precambrian
districts appear to be similarly diverse. Nonetheless, the interpretation of settings remains unsettled in many regions even
for the same province (e.g., South Australia; cf. Groves et al.,
2010; Hayward and Skirrow, 2010). Regional metamorphic

(in part collisional?) settings are common in the later stages


of some Precambrian terrains (e.g., N Baltic, Rajasthan,
Cloncurry, Lufilian arc; see Figure 2) and possibly in the Mesozoic (Lund et al., 2011). Globally, this variability is amplified
by virtue of the fact that many (but not all) of these regions have
protracted histories of tectonism, and IOCG systems formed
sporadically over intervals of 100 My or more with changing
tectonic styles (e.g., Cloncurry, Duncan et al., 2011; Williams
and Pollard, 2001; central Andes, Chen, 2010; Sillitoe, 2003;
southwestern North America, Barton et al., 2011a).

13.20.2.2.1 Association with igneous rocks (or lack thereof)


Most, but not all, IOCG districts contain abundant broadly
contemporaneous magmatism even though few deposits can
be linked to particular intrusive units (Williams et al., 2005).
Also lacking in IOCG systems is a link to characteristic igneous
textures, for example, to the distinctive crowded porphyry
texture that characterizes the source intrusions in porphyry
CuMoAu deposits (Seedorff et al., 2005). Compositions
range over much of the igneous spectrum from mafic to felsic
and subalkaline to, rarely, alkaline (Figure 3). Most associated
igneous rocks are magnetite-bearing, consistent with the correlation with relatively oxidized terrains (Haynes, 2000). Trace
element compositions indicate both arc and intraplate
magmas can be involved; these observations are compatible
with known or inferred tectonic settings. Mafic magmatism is
implicated in many areas (Barton and Johnson, 1996) including Australia where mafic magmas may have been key at
Olympic Dam (Skirrow et al., 2007) and are inferred to be an
important contributor to some of the Cloncurry systems
(Oliver et al., 2008). A few regions with magnetite-dominated
deposits have only basaltic compositions (e.g., Siberian trap
systems, Soloviev, 2010a; Mesozoic eastern United States,
Robinson, 1988). Elsewhere, intermediate compositions
dominate in many IOCG provinces (e.g., Andes, N Baltic,
Turkey, Yangtze), whereas felsic compositions dominate in
others (e.g., South Australia, Cloncurry, SE Missouri). These
generalizations aside, in many regions and even within
restricted areas, it is clear that multiple broadly coeval IOCGclan systems form with widely different magma types (e.g.,
northern Chile; western United States, Barton et al., 2011a,b).
It is noteworthy that many of the temporally and spatially
related intrusions lack early crystallizing (or any) amphibole
and hence were relatively water-poor. Conversely, a distinctive
connection with alkaline magmatism (e.g., Groves et al., 2010;
Hauck, 1990) appears to be the case only in a few districts with
definitive IOCG characteristics (e.g., Kuscu et al., 2011).
Interpretation of whole-rock data requires caution because
even nominally immobile elements (e.g., Al, Ti, HFSE) can move
during hydrothermal alteration. The unrecognized presence of
pervasive Na or K metasomatism has led to the interpretation of
many igneous rocks as having alkaline affinities (Figure 4). Such
patterns have been long recognized, for example, in metamorphosed iron oxide mineralized terrains (e.g., Lindgren, 1913,
p. 713), but they have been most clearly documented in young
IOCG-bearing volcanic terrains (e.g., Barton and Johnson,
2000; Marschik and Fontbote, 2001; Roddy et al., 1988).
Some hematite-dominated districts lack appreciable coeval
magmatism; thus, the implication is that other heat and material sources were dominant (e.g., Wernecke Breccias, Canada,

519

Iron Oxide(CuAuREEPAgUCo) Systems

12

Carajas
granites

Cloncurry
Great
Bear

Syn-collisional
granitoids

1000

e
alin
alk kaline
al
sub

8
Siberia

Rb (ppm)

Na2O + K2O (wt%)

10

Stuart Shelf

Sweden

Within plate
granitoids

100
volcanic arc
granitoids

Great Bear
fresh /
Na(Ca) alt

10
Western US IOCG
fresh / Na(Ca) alt

Copiapo (Chile)

Ocean ridge
granitoids

Humboldt

0
45
(a)

1
50

55

60

65

70

SiO2 (wt%)

75
(b)

10

100

1000

Y+Nb (ppm)

Figure 3 Selected igneous compositions temporally and spatially associated with IOCG systems (compiled for areas summarized in Barton and
Johnson, 1996; Pollard, 2006). (a) Total alkali versus silica diagram for nominally unaltered igneous rocks. (b) Tectonic discrimination diagram (Pearce
et al., 1984) showing fresh and selected altered igneous rocks.

Hunt et al., 2007; southwestern United States detachment


deposits, Wilkins et al., 1986). Similarly, a number of
metamorphic-hosted, magnetite-dominated systems have an
enigmatic relationship to regional magmatism, such as some
of those in Rajasthan (Knight et al., 2002), N Baltic (Billstrom
et al., 2010; Niiranen et al., 2007), Namibia (Moore, 2010),
Cloncurry (Duncan et al., 2011), and possibly Tennant Creek
(Skirrow and Walshe, 2002).

13.20.2.2.2 Framework lithologies and paleoclimate


Just as with magmas, rocks hosting IOCG systems range widely
from felsic to mafic igneous rocks, clastic to chemical sedimentary rocks, and their metamorphic equivalents. Of major rock
groups, only strongly reduced (carbonaceous) host rocks are
scarce (Haynes, 2000) and even they may be significant in a
few areas, such as Cloncurry (e.g., Baker, 1998; Williams,
2010b). In some terrains, particularly in arc settings, the vast
majority of rocks are igneous-affiliated (plutonic, volcanic, and
volcaniclastic). In others, there is an abundance of (meta)
sedimentary materials with fewer igneous components (Cloncurry, Tennant Creek).
Evidence for evaporitic environments and related materials,
whether contemporaneous or older, is documented in most
IOCG regions (Barton and Johnson, 1996), although the
genetic significance of evaporitic materials remains contentious. Much of this evidence is geologic, but it is supported
by geochemistry in a number of systems (discussed in the
succeeding text). Supporting this correlation, most Phanerozoic IOCG provinces are formed at low mid-latitudes similar to
the distribution of broadly contemporaneous evaporitic settings (Figure 5(a); Barton and Johnson, 1996). The few higher
latitude districts, such as the PermoTriassic Siberian magnetite breccia pipes, are formed where older Paleozoic evaporites
are abundant (Soloviev, 2010a). Such latitudinal correlations

are similar to what is reported for Phanerozoic sediment and


volcanic-hosted Cu(AgCo  U) systems (Kirkham, 1989)
and sediment-hosted PbZn deposits (Leach et al., 2010;
Figure 5(b); see also Wilkinson, Chapter 13.9).

13.20.3

Synopsis of Deposit Features

Deposits in the IOCG family share features including abundant


hypogene magnetite and/or hematite, a scarcity of sulfides, and
a distinctive suite of minor element enrichments. However, they
vary considerably in form, size, and degrees of enrichment (Cu,
Au, U, etc.), as well as in their paragenesis and zoning. This
poses the question: are these differences fundamental or part of
a continuum? Furthermore, the mineralized zones themselves
constitute only small parts of larger geologic systems that are
typified by voluminous, zoned NaCaK-dominated metasomatism and which also show considerable variability in
form, host rocks, and mineralogy (Barton and Johnson, 1996;
Hitzman et al., 1992; Porter, 2010a; Williams et al., 2005).
The deposits encompass a range of magnetite to hematite
ratios, degrees of enrichment in the minor elements (Cu, Au, P,
U, REE; rich vs. anomalous), hydrothermal alteration assemblages, and geometries. Notably, it has been long observed that
many IOCG clan occurrences are barren in that they lack
economic Cu, even though they may occur in the same district
with well-endowed deposits (Haynes, 2000; Hitzman, 2000;
Williams et al., 2005).

13.20.3.1

Deposit Types

A simple classification of economic deposits in the IOCG clan


can be made on the basis of hematite or magnetite dominance,
Cu-rich or Cu-poor, and the abundance of other minor elements (Williams, 2010a; Williams et al., 2005). Although such

Iron Oxide(CuAuREEPAgUCo) Systems

Western USA
Cortez Mtns
Humboldt Complex
Yerington district
fr / alt

14

K2O (wt%)

12
10

14
12

S. Basin and Range (Tert.)

K-feldspar = albite
(70% total)

Global unaltered
igneous rocks

Na2O + K2O (wt%)

520

6
8
10
Na2O (wt%)

12

14

Coastal belt
Great Bear

K-feldspar = albite
(70% total)

Global unaltered
igneous rocks

45

50

55 60 65
SiO2 (wt%)

70

75

80

70

75

80

70

75

80

12
Na2O + K2O (wt%)

10

40
14

Canada
K2O (wt%)

(b)

Chile

12

10
8
6
4

Alkaline
Subalkaline

0
0

14

6
8
10
Na2O (wt%)

12

14

Na2O + K2O (wt%)

Kiruna area
K-feldspar = albite
(70% total)

6
4

Global unaltered
igneous rocks

50

55 60 65
SiO2 (wt%)

12

Sweden

45

14

Cloncurry
S. Australia

10

40

(d)

Australia

12
K2O (wt%)

4 Alkaline

0
0

14

10
8
6
4

Alkaline
Subalkaline

2
0

0
(e)

Subalkaline

(c)

(a)

10

6
8
10
Na2O (wt%)

12

14

(f)

40

45

50

55 60 65
SiO2 (wt%)

Figure 4 K2O versus Na2O and K2O Na2O versus SiO2 (TAS) plots for igneous rocks from major IOCG regions showing the prevalence of alkali
exchange including the shift to higher total alkali contents. These patterns are repeated worldwide (cf. Figures 2, 3, and 11). Data compiled by Barton
and Johnson (1996, 2000).

distinctions are real, many deposits, districts, and regions contain examples that span the hematite/magnetite and Cu/Fe
spectrum (Table 1; Figure 2).

13.20.3.1.1 Magnetite- and/or hematite-dominated


deposits
Some districts contain principally Cu-poor magnetite
(actinoliteapatite) (Kiruna-type) ores; these can have nearby,

shallower stratiform hematitic ores (Table 1; e.g., Daliran et al.,


2010; Johnson and Barton, 2000; Parak, 1975). Magnetite
deposits form discordant breccias, veins, and massive bodies or
concordant replacement zones and, in some cases, stratiform
bodies of syngenetic or shallow replacement origin. Magnetitedominated deposits have variable amounts (150%) of accessory apatite, calcic amphibole, and/or pyroxene. (Some workers
(e.g., Williams, 2010a) have called Kiruna-type ores iron

Iron Oxide(CuAuREEPAgUCo) Systems

521

Mean age (Ma)


of mineralizaition

700
Provinces with
600 IOCG or IOCG-like
Siberian platform
500 systems

Abundance
(%/10 latitude)

Adelaidean

Iran
Saudi Arabia

evaporites

400

Altai (Devonian)

Turkey

300

S Urals

Siberian platform
deposits
Yangtze, E US

200
100

(a)

Lufilian

0
20
10

SW North America
Turkey

Central Andes
Clastic dominated
Pb+Zn deposits
(Leach et al., 2010)
Theoretical random
distribution

0
90

60

30

30

60

90

Paleolatitude

(b)

Figure 5 Paleolatitudes of Phanerozoic IOCG provinces (a) and clastic-host PbZn provinces (b). Determined using the continental reconstructions
of Scotese and Denham (1988). The lowmid-latitude environments are consistent with the availability of saline nonmagmatic fluids (Barton and
Johnson, 2000). The exception of the Siberian trap deposits reflects their formation coincident with voluminous older (Cambrian and Devonian)
evaporites for which the inferred position is shown by the arrow. Also shown in (b) is the theoretical distribution if independent of latitude (modified
from Leach DL, Bradley DC, Huston D, et al. (2010) Sediment-hosted leadzinc deposits in Earth history. Economic Geology 105: 593626).
Kirkham (1989) documented similar patterns for Phanerozoic sediment and volcanic-hosted Cu(AgCo  U) deposits.

Table 1

Division of IOCG deposit types with selected district (deposit) examples (cf. Figure 1)

Key metals

Magnetite-rich (hematite)

CuAu
(Ag  UCo)

Andes (Raul-Condestable, Mina Justaa),


Andes (Mantoverdea, Teresa de Colmo Punta
Carajas (Salobo, Alemaoa), Cloncurry (Ernest
del Cobrea), Carajas (Sossegoa), Gawler
Henry, Starra, Mt Elliot), Missouri (Boss
(Olympic Dam, Carrapateena, Prominent
Bixby), Tennant Creek (Juno, Gecko, West
Hilla), Great Bear (Sue Dianne)
Uncertain igneous association: Canada
Peko) W United States (Yeringtona, Lights
Creek)
(Wernecke), W United States (detachment)
Andes (Pampa Larga), Cloncurry (Valhalla?),
Singhbhum (Turamdih, Jadugudaa), Great
Bear (Echo Bay), Cloncurry
Gawler (Mt Painter a), Great Bear (Echo Bay)
Sweden (Kirunaa), Iran (Bafqa), Andes
Sweden (Kirunaa), Iran (Bafqa), Missouri (Pilot
(Marcona, El Romeral, El Laco), Missouri
Knoba), W US (Cortez Mtnsa, Humboldta,
Iron Springsa), Mexico (La Perla)
(Pea Ridgea), Siberian traps (Vetka,
Korshunovsk), Kazakhstan (Turgai), W United
States (Humboldt a, Eagle Mtn, Iron Springs)

UAgAuCoCu
Fe(CuCo)

Hematite-rich (magnetite)

Others
Igneous association: N Baltic
(Pahtohavre, Finnmark), India
(Khetri), Cloncurry
(Greenmount, Eloisea)

Cloncurry (Mt Cobalt), W United


States (Humboldta)
Magnetite-rich but uncertain
igneous association: barren
ironstones at Tennant Creek,
Cloncurry

Deposit with mixed characteristics (e.g., Mt Hm) within single deposits. In a broader context, most Fe-oxide-rich districts show mixed characteristics.
Source: Derived from Williams PJ, Barton MD, Fontbote L, et al. (2005) Iron-oxide-copper-gold deposits: Geology, spacetime distribution, and possible modes of origin. Economic
Geology 100th Anniversary Volume: 371406; Williams PJ (2010b) Magnetite-Group IOCGs with Special Reference to Cloncurry (NW Queensland) and Northern Sweden: Settings,
Alteration, Deposit Characteristics, Fluid Sources and Their Relationship to Apatite-Rich Iron Ores. Short Course Notes, vol. 20. Canada: Geological Association of Canada.

oxideapatite deposits a variant on the older term magnetite


apatite. While a valid descriptive distinction, this term logically
includes titanium iron oxide ores (e.g., nelsonites, Kolker, 1982),
which differ in many ways other than the Ti contents, notably in
their lack of hydrothermal features.) In many districts, these
bodies are transitional with actinolite- and/or apatite-rich
veins. Sulfides, including CuFe sulfides, are commonly present
in minor quantities. These deposits are also enriched in REE
(Johnson and Barton, in press). For the most part, magnetite
apatiteactinolite deposits make Fe ores and only rarely
contain economically significant Cu. They have been routinely

distinguished from magnetite-dominated, Cu-rich varieties (e.g.,


Williams, 2010b).
The magnetite-only types appear to form a continuum with
systems where magnetite is important and may be dominant,
but hematite, either as specularite or mushketovite (magnetite
replacing hematite), is present and can be common. These
magnetite-rich, hematite-bearing deposits more commonly
have Cu mineralization and include many of the major
Andean (Sillitoe, 2003), Brazilian (Xavier et al., 2010), and
Australian (Williams, 2010b) districts. Although many magnetite  hematite deposits contain abundant Cu, this is far from

522

Iron Oxide(CuAuREEPAgUCo) Systems

universal (e.g., SE Missouri, Seeger, 2000; S Australia,


Bastrakov et al., 2007; Siberia, Soloviev, 2010a).
Hematite-dominated deposits are apparently less common
(Skirrow, 2010; Skirrow et al., 2007; Williams, 2010a). The preeminent Cu-rich examples are those of the Stuart Shelf in South
Australia (Olympic Dam, Prominent Hill, and Carapateena;
Skirrow et al., 2007) where hematite is both primary and secondary. Hematite-rich systems, including mushketovite-rich examples, are common in the Andes as well. The copper-rich types
are mainly discordant, structurally controlled breccias and veins.
Discordant to stratabound sulfide-poor hematite(REE-bearing)
deposits occur as well (e.g., Sweden; Parak, 1991; Mt. Painter,
Lambert et al., 1982; Bafq, Iran, Daliran et al., 2010; Jami
et al., 2007; Great Basin, United States, Johnson, 2000; Tobey,
1976). In some areas, hematitic ironstones may be substantially
older than contained CuAu mineralization, though this remains
unsettled (e.g., Tennant Creek, Skirrow and Walshe, 2002). Even
in these areas, hematite is commonly associated with magnetitebearing rocks or portions of deposits.
Although magnetite and hematite types may occur independently, sufficient control is lacking in most districts to be
confident of system-scale zoning. Where the third dimension is
well exposed, both lateral and vertical zoning occurs including
from magnetite to hematite types and with corresponding
changes in alteration and accessory minerals (e.g., Gawler
province, Skirrow et al., 2007; Sossego-Sequeirinho, Carajas,
Monteiro et al., 2008a; Manto Verde, Chile, Benavides et al.,
2007, Rieger et al., 2010; Humboldt mafic complex, Nevada,
Johnson and Barton, 2000). The best Cu(Au)-mineralized
systems are discordant; conversely those stratiform deposits
that may have formed at or near the earths surface lack abundant Cu (e.g., Kiruna, Sweden, Parak, 1975; El Laco, Chile,
Rhodes et al., 1999; Bafq, Iraq, Daliran et al., 2010).

13.20.3.1.2 Fe oxide-poor Cu(Au/Ag) deposits of proposed


affinity to IOCG systems
A source of ambiguity in the IOCG clan stems from inclusion
of other deposit types that are proposed variations on the Fe
oxide-dominated theme (Table 1; Haynes, 2000; Williams,
2010a). Variations include (1) ferroan carbonate- or quartzdominated veins with accessory Cu(Au) with or without
hematite  magnetite, (2) pyrite- and/or pyrrhotite-bearing
Cu(Au) deposits with sparse or absent Fe oxides, and (3)
volcanic-hosted Cu(Ag) deposits with minor (<5%) hematite. Briefly noted later are magnetite-bearing porphyry Cu and
alkaline magmatic systems, neither of which shows a close
spatial nor temporal relationship to IOCG systems.
The quartz/carbonate veins are of two types. Some represent
the fringes of Fe oxide-rich zones and have similar hematitebearing mineralogies (e.g., in Chile; Kreiner, 2011; Sillitoe,
2003; Great Bear Magmatic Zone, Mumin et al., 2010). These
can extend up to 10 km in strike and be tens of meters in width
(Mumin et al., 2010). In some of these districts, the carbonate
quartz vein systems include distal AgAsNiCoU deposits
(see Chapter 13.16). Others lack this spatial association and
are more common in metamorphic terranes where they are synor late metamorphic, commonly cutting higher temperature
fabrics and rocks (e.g., in N Baltic: Finnmark, Ettner et al.,
1993; Pahtohavare, Billstrom et al., 2010). In many respects,
this latter group resembles orogenic Au veins (Chapter 13.15).

Both types have well-developed Na and/or K alteration. An


intriguing parallel exists with vein and replacement siderite
(CuAgPb) deposits derived from basinal or orogenic fluids
(e.g., Palinkas et al., 2009).
CuAu deposits with pyrite  pyrrhotite and variable
amounts of magnetite occur in several metamorphic terranes
including Khetri (Knight et al., 2002), Tennant Creek (Skirrow
and Walshe, 2002), Singhbhum (Changkakoti et al., 1987; Pal
et al., 2009), and Cloncurry (Baker, 1998; Mark et al., 2006a)
where they are accompanied by NaCa alteration and tend to
occur in areas of reduced host rocks. This group overlaps with
variably magnetite-bearing metasediment-hosted CuCoAu
deposits (Slack et al., 2011) and with Udokan, Siberia, a huge,
enigmatic sediment-hosted magnetite-bearing Cu(Ag  Au)
deposit (Soloviev, 2010b).
The third group comprises hematite-bearing volcanic- and
sediment-hosted Cu(Ag) deposits. These zone from hematite
(minor, with little or no Fe addition), though chalcocite to
bornite to chalcopyrite to pyrite. The premier examples occur
in the Mesozoic volcanic rocks of northern Chile where they
overlap the time and region containing the well-defined IOCG
systems (Maksaev and Zentilli, 2002). None of the Cu(Ag)
deposits are IOCGs by the definition used here in that they lack
abundant Fe oxides; however, they share a number of distinctive features including Cu(Ag  Au) enrichments, alkali
(calcic) alteration, an oxidized but sulfur-poor geochemistry,
and a genetic link with brines. The associated metals, zoning,
and aspects of alteration in these deposits overlap with features
in many sediment-hosted Cu (and some U) systems with which
they share similar paleogeographic settings (Kirkham, 1989).

13.20.3.1.3 Possible modern analogues


Several modern geothermal systems share features with IOCG
systems (Barton and Johnson, 2000). The most compelling
modern analogue is the Salton Sea geothermal system
(McKibben and Hardie, 1997). This basalt-driven geothermal
system circulates evaporite-derived low total S brines that precipitate hematitechalcopyritecobaltian pyriteallanitic epidote veins. Flashing of these brines deposits magnetite with
high Au and U contents. The most abundant metals in the
brine Mn, Zn, Pb do not precipitate for want of sufficient
sulfide. Highly saline geothermal systems are known in the
Afar region of Ethiopia and adjacent Djibouti where modern
basaltic magmatism circulates fluids derived in part from evaporated seawater and evaporites in the modern rift (Bonatti
et al., 1972; Bosch et al., 1977).

13.20.3.2

Grade, Size, and Form

Few multielement resource estimates exist for IOCG ore bodies. Copper is rarely reported in deposits mined for Fe.
Conversely, P, Ti, and other elements are rarely presented for
Cu-rich deposits. Values are skewed even when they are available. For example, S is deleterious in Fe ores and thus sulfide(and Cu-)rich zones tend to be excluded from Fe resource
calculations. Compilation of published data (Williams et al.,
2005) shows that deposits mined for Fe including the magnetite(hematite  apatite) ores typically contain 3555 wt% Fe,
1 wt% Ti, 0.012 wt% P, and 02 wt% S. Where both are
reported, Fe and Cu grades span a range of metal contents for

523

Iron Oxide(CuAuREEPAgUCo) Systems


both magnetite- and hematite-dominated types (Figure 6). Ore
bodies range in size from a few million to >1 billion tons. Iron
concentrations vary from about 15 to 55 wt%, whereas Cu
contents vary from a trace to about 1.5 wt%. From the
deposit-type definition, it follows that all Cu-rich IOCG ores
contain abundant Fe oxides; however, a number of the Feonly ore bodies contain >100 000 tons Cu (Figure 6(a)).
Average grades reflect the principal economic commodity,
with high Cu (and lower Fe) in those mined for Cu, and the
converse in the Fe deposits (Figure 6(b)). Data are limited to a
subset of deposits given that only one of the two commodities
is of economic interest. A few deposits have both Fe and
Cu have economic importance as byproducts or coproducts
(e.g., Candelaria and Santo Domingo Sur, Chile, Daroch
and Barton, 2011; Cornwall, Pennsylvania, Lapham, 1968;
Pumpkin Hollow, Nevada, Dilles et al., 2000). This continuum
in metal contents makes restricting IOCG deposits to the Curich group a function of economics (cf. Williams et al., 2005).
Overall, Cu tends to be more abundant in hematite/mushketovite-dominated deposits. In some regions, only a few
deposits contain appreciable Cu (N Baltic, SW, mid-continent
North America, and Iran). In other regions (Chile, Gawler),
many hematite magnetite-bearing examples are Cu mineralized, whereas the magnetite-dominated deposits are more varied in their Cu contents and many have only minor Cu. Some
older, variably metamorphosed regions (Cloncurry, Carajas)
have both hematite-dominant and magnetite-dominant Curich deposits. In Cu-rich systems, Au averages 0.10.5 ppm
(Williams et al., 2005). Analogous patterns can be deduced
for other elements including U, REE, and Co.

13.20.3.3

Ore Mineralogy and Paragenesis

In addition to the iron oxides, the ores contain <15% total


sulfides, plus accessory silicates, carbonates (commonly ferroan;
siderite is locally abundant), and phosphates (Hitzman et al.,
1992; Williams et al., 2005). Chalcopyrite and pyrite are the
most common sulfides with scattered bornite, chalcocite, pyrrhotite, sphalerite, and NiCo sulfides. Barite and anhydrite
occur in some deposits; the latter, along with rare halite, is
abundant in a few skarn systems (e.g., Soloviev, 2010a; Zhang,
1986). Weathering effects have not been discussed in most areas
with the exception of deposits in South Australia where supergene processes may have redistributed Cu, U, and Au (Oreskes
and Einaudi, 1990; Skirrow, 2010).
Hematite-dominated assemblages commonly have higher
sulfide contents than related magnetite ores. Hematite typically
occurs as specularite, which may be replaced by magnetite, but
hematite also occurs in more massive steely form and as finegrained replacement of other minerals including silicates and,
especially, magnetite. Only a few deposits, including Olympic
Dam and Salobo, contain abundant bornite  digenite
chalcocite that occur with hematite or magnetite (Williams
et al., 2005). Common gangue phases in the hematitedominated assemblages include quartz, Ca(FeMg) carbonates,
chlorite, sericite, and sparse apatite. Rarer elements, including
REE and U, commonly occur as discrete phases. Related alteration minerals reflect more acid conditions and include sericite,
chlorite, quartz, and rarely, pyrophyllite, kaolinite, and alunite.
Magnetite-dominated assemblages contain apatite and/or
actinolite(hastingsite), minor to absent sulfides (typically

10

10
Main commodity

Typical recoverable
Fe grades

Dominant
oxide

Cu Fe
Mt
Hm

Porphyry
Cu deposits

10x

10

u/Fe

tal C

crus

Cu (wt%)

Cu (wt%)

Typical recoverable Cu grades

0.1

0.1
C

in

co
nt
a

0.

u/Fe

tal C

Crus

75% Mt
75% Hm

Porphyry Cu/Au limits


Cu(AuMo)
calcalkaline
CuAu alkaline
AuCu dioritic

ed

0.01

0.01
10
(a)

100

1000

Reported resource (Mt)

10 000 0
(b)

10

20

30

40

50

60

Fe (wt%)

Figure 6 Size and grade of IOCG deposits for which Cu with or without Fe contents is available (based on a compilation by Barton and Johnson used in
Williams et al., 2005). (a) Copper grades and tonnages for magnetite-dominated and hematite-dominated deposits, indicating principal commodity.
Copper contents are not commonly reported for Fe-only deposits, though many mention Cu minerals. Fields for global porphyry Cu deposits are
shown for comparison. (b) Comparison of Cu and Fe contents for the same deposits as in (a) where Fe contents are also available. Also shown are
approximate cutoff (lowest) grades for ores of Fe and Cu, a representative field for the bulk composition of magnetite-rich porphyry systems, and a
band indicating representative crustal Cu/Fe. In both figures, note the approximate continuum in grade and size. See text for discussion.

524

Iron Oxide(CuAuREEPAgUCo) Systems

pyrite > chalcopyrite), and a variety of accessory minerals


including titanite. Associated alteration phases typically
include combinations of sodic plagioclase, marialitic scapolite,
calcic amphiboles or pyroxenes, epidote, titanite, and in some
places (notably, but not exclusively, in carbonate host rocks),
calcic garnets. Biotite and/or K-feldspar are also common alteration products in transitional magnetitehematite systems that
are hosted by intermediate to felsic rocks. Other Fe2 minerals
are commonly present and, in some cases, may be volumetrically important. Accessory ferroan carbonates including siderite occur in a number of deposits (e.g., at Olympic Dam;
Oreskes and Einaudi, 1990) where they may be either early
or late. Silicates have variable Fe2 contents, and locally, they
may be important as in fayalite-bearing mineralization at Vergenoeg, RSA (Borrok et al., 1998), or metamorphosed (?)
fayalitegrunerite-bearing assemblages as at Salobo (Requia
et al., 2003). Ilmenite is rare; nearly all Ti occurs in titanite.
In deposits where both Fe oxides occur, magnetite-rich
assemblages typically form earlier, at higher temperatures, and
at greater depths than hematite-dominated, variably sulfidebearing assemblages. Although magnetite commonly oxidizes
to hematite, the reverse appears to be more common in many
systems, even in deposits where there is little evidence for new
growth of magnetite (e.g., Marschik and Fontbote, 2001). Cubearing sulfides are typically late in the paragenesis and may
postdate Fe introduction. In detail, the parageneses can be
quite complicated and exhibit multiple reversals of sequence
(e.g., Kreiner, 2011) including clearly prograde sequences (e.g.,
Daroch and Barton, 2011; de Haller and Fontbote, 2009).

13.20.3.4 Minor Element Contents and Mineralogy: UTh,


REE, CoNiV, ClFBr, B
A distinctive and broad suite of minor elements occur both in the
principal phases and in accessory minerals. A few elements U,
REE, and Co have been of economic interest. Others show informative patterns including Th, NiV, FClBr, B, and AsBiTe.
Uranium, Th, and REE contents commonly, though not
universally, exceed 10  crustal abundance: 5500 ppm for U
(Hitzman and Valenta, 2005) and 0.23%, rarely >10%, for
P
REE (Johnson and Barton, in press). Abundances correlate
with those in the host rocks (Barton and Johnson, 1996); for
example, deposits associated with felsic and alkaline igneous
rocks generally have higher concentrations of the more lithophile elements, such as the REE, than those deposits associated
with more mafic rocks. This correlation has been used as
evidence for either a specialized alkaline source (e.g., Groves
et al., 2010) or a host-rock control independent of magmatic
processes (e.g., Barton and Johnson, 1996, 2000).
Common U and Th minerals include uraninite, davidite,
brannerite, thorite, thorianite, and allanite with appreciable
quantities in apatite, titanite, xenotime, and epidote
(Hitzman and Valenta, 2005; Mazdab et al., 2008). Although
fairly low grade, the worlds greatest single U resource is Olympic Dam (9.2 Gt at 270 ppm U, BHP Billiton Ann. Rept.,
2009). Smaller U resources occur in other IOCG settings
including magnetite-hosted deposits (e.g., Singhbhum district,
India, Pandey et al., 1994) and distal AgCoNiAsU veins
around the magnetiteapatiteactinolite deposits (e.g., Great
Bear Magmatic Zone, Canada, Mumin et al., 2007).

REE-bearing minerals are ubiquitous in IOCG systems. In


magnetite-dominated ores, REE are located primarily in rockforming minerals including apatite, titanite, and epidote
allanite (Mazdab et al., 2008). In hematite-rich assemblages,
REE occur in apatite and in discrete REE, phases occur including
phosphates (monazite, xenotime, britholite), carbonates (bastnaesite, synchysite), and silicates (allanite, stillwellite) (Johnson
and Barton, in press; Oreskes and Einaudi, 1992). Monazite and
other REE-rich minerals postdate apatite and may reflect additional enrichment or breakdown of earlier REE-bearing apatite
and titanite (e.g., Andersson et al., 2002; Torab and Lehmann,
2007). Although REE are not currently recovered from any IOCG
system, some deposits contain large and locally high-grade
resources (Johnson and Barton, in press).
Among the siderophile elements, Co and Ni are documented in small amounts, typically in pyrite (0.12 wt%), in pyrrhotite (when present), and locally, in discrete Ni and Co
sulfides and arsenides (e.g., Mazdab, 2001; Monteiro et al.,
2008b; Pal et al., 2009; Rusk et al., 2010). By far, the highest
enrichments are in the distal CoAgAs(NiU) deposits that
occur in some districts (e.g., Great Bear, Mumin et al., 2007).
Few data are available on overall Co enrichments; estimates of
bulk ore grade based on the bulk pyrite content indicate that
Co (and Ni) rarely exceeds a few 100 ppm. Co Ni enrichment
in the iron sulfides more likely reflects the overall scarcity of
sulfur rather than an abundance of these elements. Only a
handful of IOCG-related magnetite-rich deposits are also rich
in cobalt (e.g., some CuCo ores, Slack et al., 2011); however,
for many years, principal cobalt supply in the United States
came either from magnetitebiotitesulfide ores from the
Idaho Cobalt Belt (Nash and Hahn, 1989) or the Co-rich pyrite
of the Pennsylvania deposits (Lapham, 1968).
Apart from minor V and Ti, other siderophile elements are
rarely enriched. Magnetite is purer than those from igneous or
FeTiV systems. Magnetite generally contains minor Ti (2 wt
%) and V (1 wt%) and trace amounts of Cr, Al, Ga, Zn, Mn,
Co, and Ni. Dupuis and Beaudoin (2011) argue that IOCG
magnetites have distinctive compositions, whereas others (e.g.,
Monteiro et al., 2008b; Rusk et al., 2009; Xavier et al., 2011) have
shown that magnetite compositions can be highly varied even in
the same deposit. Other Fe-bearing phases can also be
V-enriched in IOCG deposits including epidote, hematite, and
coulsonite (F.K. Mazdab, written comm., 2011). Most Ti occurs
in titanite or rutile, a reflection of the moderate temperatures
and relatively oxidized, commonly Ca-rich, conditions.
Halogens occur not only widely in OH sites in silicates
and apatite (Cl, F) but also commonly in marialitic scapolite
(Cl, Br), locally in fluorite, and rarely in chlorosilicates, such as
pyrosmalite (Dreher et al., 2008; Mazdab, 2001). With the
exception of Cl in marialitic scapolite, halogens content are
low with Cl exceeding 1 wt% only in hastingitic amphiboles,
apatites, and some biotites (e.g., Mazdab, 2001, 2003; Monteiro
et al., 2008b; Treloar and Colley, 1996). Fluorine contents are
generally low except as a constituent of apatite and as fluorite in
deposits that are hosted in F-rich, low-Ca felsic rocks (e.g., in
South Australia and SE Missouri, Einaudi and Oreskes, 1990;
Bushveld granites, RSA, Crocker et al., 1988). Br and I, are
known from fluid inclusion analyses, and Br is present in scapolite at levels (0.0X%) where evidence indicates that Br/Cl
ratios reflect fluid ratios (Pan and Dong, 2003).

Iron Oxide(CuAuREEPAgUCo) Systems

Boron, though rarely abundant, occurs in accessory phases


and has captured attention particularly for use as an isotopic
tracer (Collins, 2010; Tornos et al., 2011; Xavier et al., 2008).
Nearly all IOCG regions contain tourmaline, whereas other
boron minerals, such as dumortierite, axinite, danburite, datolite,
and ludwigite, are less common. Dumortierite occurs in IOCGrelated advanced argillic assemblages in Chile, the western United
States, and elsewhere (Kreiner, 2011; Ray and Dick, 2002).
Although most IOCG tourmaline compositions lie near the
schorldravite join, they exhibit a wide range of compositions
that correlate with the alteration assemblage (Collins, 2010).
Accessory minerals containing As, Bi, and Te are reported
from many deposits, yet only a handful are known to contain
appreciable amounts. These take the forms of sulfides (arsenopyrite, cobaltite, carrollite, bismuthinite), many arsenides, and
trace tellurides. Such minerals are typically late and sparse if
recognized at all yet are common in a few regions, such as
Tennant Creek (Huston et al., 1993; Skirrow and Walshe,
2002) and the Great Bear magmatic zone (Mumin et al., 2010).

13.20.4
Zoning

clastic rocks, KNa(Ca) exchange is manifested mainly in the


changes in framework silicate mineralogy, with K-feldspar
(and/or biotite) developing in K-altered zones and variable
sodic plagioclase or scapolite in sodiccalcic (NaCa, sensu
lato) altered volumes. Whole-rock compositions in IOCG districts and regions show these changes through large shifts in
bulk composition especially Na for K exchange (Figure 4) and
minor element contents (Figure 7). In many of these areas, the
feldspars take on a characteristic maroon to red color due to
tiny inclusions of hematite and known colloquially as red
rock alteration (Wenner and Taylor, 1987; Williams, 1994a).
Carbonate hosts are far less common; where present, they
contain similar calcic minerals but reflect a fundamentally
different type of metasomatism.

13.20.4.1.1 Sodic to calcic alteration types


Sodic to calcic alteration types (individually, Na, Na(Ca), and
Ca; collectively, NaCa) encompass a spectrum reflecting
removal of K and addition of Na and(or) Ca (Figures 4 and
7(a)). It is manifested in formation of new framework silicates
(or, less commonly, epidote or garnet) as the aluminous
phase and a suite of increasingly Ca-rich mafic phases. Silica
may be added or subtracted; in many areas, quartz contents
decrease through dissolution (as on a fluid heating path;
Fournier, 1987) or by reactions to make other phases
(e.g., CaAl2Si2O8 4SiO2 2Na 2NaAlSi3O8 Ca2; Carten,
1986). Textures can be preserved or obliterated. In extreme
cases, essentially monomineralic rocks are produced including
albitites, scapolitites and epidosites. Metals including Fe and Cu
are commonly removed when the original Fe-bearing phases
(including trace magmatic sulfides) are converted to other
minerals (Figure 7(b)). Metal loss is widespread (Barton and
Johnson, 1996, 2000; de Jong et al., 1998; Johnson, 2000;
Johnson and Barton, 2000; Menard, 1995; Oliver et al.,
2004, 2008; Williams, 1994b); however, it is not universal, for
instance, NaCa alteration in many areas in Cloncurry shows
consistent depletion in Zn and Pb but irregular changes in Cu
(Oliver et al., 2004). Locally, metals including Fe and Cu are

Hydrothermal Alteration and System-scale

This section summarizes IOCG-related alteration types and


zoning, aspects of mass transfer, and a comparison of these
features with other kinds of hydrothermal systems. Complex
in detail (reflecting geologic context), alteration is remarkably
voluminous and it exhibits large gains and losses in multiple
components. Zoning also varies in detail, yet the broad patterns
resemble those illustrated in Figure 1. The combined intensity
and style of hydrothermal alteration further distinguishes the
IOCG clan from other families of ore-forming systems.

13.20.4.1

Types of Hydrothermal Alteration

All IOCG systems exhibit extensive NaCaK(Fe) exchange


coupled with gains and losses in Fe, Si, and minor elements.
Hydrolytic (acid) alteration is present in many. In igneous and

+500

Ranges and geometric means of mass changes


for IOCG-related Na(Ca) alteration

(a)

525

50 mm

(b)

(igneous protoliths; constant AI2O3 normalization)


3 of 14 > 0

Mt

Change in mass (%)

2 of 13 > 0

+100

+42
+8

2 of 14 > 0

2 of 14 > 0

+3

II

-47

-9

Ru
-24

-30

-50

-5

-9

-11

-42

-43

-53

-47

-43

-57

-90
-95

Ab

Ti
14 data sets from Australia, Chile, USA

Na2O K2O

Rb

Ba

CaO

Sr

MgO MnO SFe

Co

Cu

Zn

SiO2 TiO2 P2O5

Figure 7 Gains and losses during Na(Ca) alteration. (a) Changes in mass compared to protolith normalized to constant Al2O3. (b) Backscattered
electron image illustrating Fe removal during Na(-Ca) alteration; igneous ilmenite (Il) and magnetite (Mt) converting to rutile (Ru) then titanite (Ti) in
albitic plagioclase (Ab) (Cortez Mountains, Nevada; Johnson, 2000). See text for discussion.

526

Iron Oxide(CuAuREEPAgUCo) Systems

added with NaCa alteration types near or within some deposits


(e.g., Corriveau and Mumin, 2010; Johnson and Barton, 2000;
Kreiner, 2011).
Mineralogy reflects the specific alteration type and change
in bulk composition (e.g., Dilles et al., 1995). In sodic (Na)
alteration, K and Ca are removed and albite, chlorite, and rutile
are the main minerals; they replace, respectively, the earlier
feldspars (white micas), mafic minerals, and oxides/titanite.
In sodic(calcic) (Na(Ca)) alteration, Ca-bearing minerals
including calcic amphiboles (actinolitic to hastingsitic) and
diopsidic pyroxenes replace the mafic minerals, while oligoclase, marialitic scapolite, or epidote replace the feldspars, and
titanite is the typical Ti phase. In calcic (Ca) alteration, calcic
plagioclase, garnet, or clinozoisite/epidote replace the felsic
minerals, calcic pyroxene replaces the mafic minerals, and
titanite is the Ti phase.
As implied in Figure 1, NaCa assemblages tend to be deep
and peripheral. In such cases, they tend to be metal-depleted
with loss of iron and base metals (Figure 7). However, in some
districts, NaCa assemblages accompany magnetite-rich (typically Cu-poor) styles of mineralization (e.g., Johnson and
Barton, 2000; Perring et al., 2000). NaCa alteration can be
transitional into more Fe-rich and, locally, K-bearing assemblages with hastingsitic amphibole and, locally biotite, or
when conditions become more acid, sericite, or chlorite (e.g.,
Dilles et al., 2000; Kreiner, 2011). In some IOCG systems, the
Na(Ca) minerals locally take on weakly peralkaline compositions manifested as acmitic and riebeckitic components in
pyroxenes and amphiboles (e.g., Barton and Johnson, 1996;
Vapnik et al., 2007; Wang and Williams, 2001). Significantly,
none of these areas has any evidence for peralkaline
magmatism; thus, a different source of alkalinity is required.
Sodiccalcic alteration by itself or even with some of the
other alteration types does not define an IOCG system. The
classic example is seafloor alteration, where deep sodic to calcic
alteration assemblages underlie low-temperature K alteration
and focused acid alteration in upwelling zones (e.g., Alt, 1999;
Franklin et al., 2005). NaCa mineral assemblages can also
reflect meta-evaporites or regional-scale metasomatism related
to metamorphic reactions with evaporitic fluids. These are well
documented in a number of areas (that typically also contain
IOCG systems) including Australia (Oliver et al., 2004), N
Baltic (Frietsch et al., 1997), and India (Golani et al., 2002).

13.20.4.1.2 Carbonate-hosted alteration: Skarn and Fe


oxide replacement
Where present in IOCG systems, carbonate rocks host distinctive calcsilicate and Fe oxide alteration. In contrast to the calcic
alteration (Ca addition) in aluminous (igneous or sedimentary) protoliths, formation of calcsilicate and Fe oxide minerals
in carbonate rocks requires a fundamentally different kind of
metasomatism addition of Fe  Si. Most carbonate-hosted
alteration is dominated by magnetite  hematite with subordinate andraditic garnet, diopsidic pyroxene, actinolitic amphibole, chlorite, serpentine, carbonates, and other nonsilicates
(rarely including anhydrite, e.g., Zhang et al., 2011, or halite,
Soloviev, 2010a). Many deposits consist largely or entirely of
Fe oxides with little or no calcsilicate minerals. Because these
types require carbonate protoliths, they are most common in
the Phanerozoic, for example, in the eastern United States

(Robinson, 1988), southwestern North America (Barton


et al., 2011a), Siberia (Soloviev, 2010a), Kazakhstan
(Hawkins et al., 2010), and the lower Yangtze basin, China
(Pan and Dong, 1999). In most of these districts, carbonatehosted magnetite  skarn is but one variety of a spectrum of
deposits ranging from igneous-hosted magnetiteapatite
actinolite veins and NaCa alteration to distal hematite
magnetite  chalcopyrite veins and replacement bodies in
volcanic rocks (e.g., Iron Springs, Tobey, 1976; Yerington,
Dilles et al., 2000; Yangtze region, Pan and Dong, 1999).
By no means are all Fe oxide-rich skarns IOCG type. Magnetite and hematite are common accessory minerals in many
skarn types, particularly in dolomitic hosts (Einaudi et al.,
1981). Many magnesian magnetite skarns form with other
types of igneous-related deposits (e.g., Sn greisens, porphyry
CuMo) and lack the characteristic minor element suites and
the distinctive associated alteration of the IOCG clan.

13.20.4.1.3 K-rich alteration: High-temperature and


low-temperature types
K-rich alteration in IOCG environments belongs to two separate
groups with distinctly different characteristics (Figure 1; Barton
and Johnson, 2000). The first type is proximal to, and associated
with, mineralization. It consists of discordant, moderate- to
high-temperature mineral assemblages. A second, lowtemperature type is widespread, commonly regionally extensive
and typically stratabound alteration that is commonly metal
deficient and does not coincide with orebodies. Both types can
be intense with K2O contents exceeding 8 wt% (Figure 4), yet
they represent fundamentally different parts of the systems.
The proximal variety contains K-feldspar and/or biotite,
typically in association with abundant magnetite and with
calcic or magnesian minerals, such as actinolite, epidote, titanite, or chlorite (Hitzman et al., 1992; Williams et al., 2005).
Where calcic phases are abundant, this type has been termed
calcpotassic (KCa) by analogy with other alteration types.
These assemblages are widely interpreted to be part of the
upwelling and therefore cooling part of the hydrothermal system (Figure 1; e.g., Andes, Sillitoe, 2003; South Australia,
Skirrow et al., 2007; Carajas, Xavier et al., 2010). The abundance of Fe2-bearing silicates and the scarcity of hematite
imply moderate redox conditions. Quartz may be added but,
in most cases, is minor compared to the Fe oxides. These
characteristics contrast with Ksilicate alteration in most porphyry and greisen systems which generally contain far less
magnetite, considerably more quartz, and few, if any, calcsilicate minerals (Barton and Johnson, 2000; Proffett, 2003).
Only alkalic porphyry/epithermal systems have similar assemblages though much less Fe oxide (Cooke et al., 2007; Jensen
and Barton, 2000; Lang et al., 1995, Chapter 13.14).
In contrast to proximal, discordant magnetite-bearing Krich alteration, distal K-enriched (meta)volcanic rocks with
abundant K-feldspar and/or biotite are documented in many
IOCG districts (e.g., N Sweden, Parak, 1975; SE Missouri,
Kisvarsanyi, 1972; N Chile, Marschik and Fontbote, 2001;
Ray and Dick, 2002; Turkey, Helvaci and Griffin, 1983; Great
Bear, Gandhi, 1988; SW N Am, Barton et al., 2011a) and have
been noted for over a century (e.g., Lindgren, 1913). K2O
contents may exceed 10 wt% and locally reach 14 wt% in
intermediate and mafic compositions. The presence of

Iron Oxide(CuAuREEPAgUCo) Systems

hematite and available whole-rock analyses demonstrate that


many are highly oxidized (molar Fe2O3/FeO  1). They are
largely stratabound and extend for many kilometers. They
occur in the upper parts of the geologic column, which hosts
the ore-bearing systems. Where unmetamorphosed, this alteration forms hematitic, K-feldspar-dominated rocks that may
show evidence of base metal depletion (Hollocher et al., 1994;
Roddy et al., 1988). They are inferred to be shallow alteration
products of saline surficial fluids (e.g., Ennis et al., 2000; Kerrich
and Rehrig, 1987; Roddy et al., 1988; Wenner and Taylor,
1987). Such K-rich metasomatism reflects a briny diagenetic
environment but can form in the absence of the energy source
and focused high-T fluid flow needed for IOCG mineralization.
A possible variation of the low-T alteration is the widespread scapolitebiotite-bearing rock that occurs in a number
of IOCG-bearing metamorphic terranes (Serdyuchenko, 1975;
N Baltic: Frietsch et al., 1997; Australia: Oliver et al., 2004). The
origin of these NaK-rich rocks is uncertain; they might represent metamorphosed low-temperature alkali-rich alteration or
they might be metamorphosed and/or remobilized evaporitic
materials.

13.20.4.1.4 Hydrolytic (acid) alteration


Hydrolytic (acid) alteration forms in the upper and proximal
portions of many systems, mostly with hematite-dominated or
mixed hematitemagnetite ores (Figure 1; Hitzman et al.,
1992). Acid assemblages include sericitic and chloritic types
and, in some areas, distinctive sulfur-poor advanced argillic
assemblages (Barton and Johnson, 1996, 2000; Kreiner, 2011).
Overall, acid alteration is considerably less voluminous than the
NaCaK(Fe) alteration. In well-developed cases, feldspars are
fully converted to sericite or more rarely chlorite, and mafic
minerals alter to chlorite and other iron minerals (hematite pyrite). Iron is added as hematite or magnetite and may
replace host minerals, locally preserving delicate textures
(e.g., Olympic Dam, Oreskes and Einaudi, 1990). Sericite
hematite quartz forms in many felsic-hosted systems (e.g.,
Olympic Dam, Pea Ridge; Einaudi and Oreskes, 1990), whereas
chlorite  sericite is more common in intermediate- to mafichosted systems (e.g., in many Chilean examples; Kreiner and
Barton, 2009; Rieger et al., 2010; Sillitoe, 2003; in Carajas,
Torresi et al., 2012). In many cases, feldspar hydrolysis is
incomplete, yielding mixed sericite or chlorite plus K-feldspar
or albite assemblages. Sericitic/chloritic alteration in IOCG systems contrasts with other settings, notably in porphyry and
seafloor systems in that the others are typically pyrite-rich, lack
abundant hematite and/or magnetite, tend to exhibit more
intense acid leaching, and are commonly more voluminous.
Advanced argillic alteration has been noted in about 15
IOCG systems (Kreiner, 2011) where it is evidenced by diagnostic AlSi phases including pyrophyllite, andalusite, diaspore,
corundum, kaolinite, and dumortierite. Distinctive features of
IOCG-related advanced argillic alteration include common
specular hematite, a near absence of sulfides, and typically a
dearth of sulfates. Alunite-rich alteration at El Laco, Chile, is
unusual and may represent a steam-heated zone or it could be
unrelated to the iron oxide bodies (Sillitoe and Burrows, 2002).
Advanced argillic zones occur at high structural levels, mainly in
contemporaneous volcanic rocks (Kreiner, 2011; Ray and Dick,
2002). Although the combined structural and stratigraphic

527

control of IOCG-related advanced argillic alteration parallels


that in other hypogene advanced argillic alteration (e.g., in
epithermal AuAg systems), IOCG systems have far less sulfur
and lower sulfidation states.

13.20.4.2
Patterns

System- to Regional-Scale Spatial and Temporal

The general architecture of IOCG systems (Figure 1) derives


from observations in many regions (Barton and Johnson,
1996; Hitzman et al., 1992; Williams et al., 2005; papers in
Porter, 2000, 2002, 2010b). Superposition of KFeCaH
alteration types is the rule within the vicinity (a few km maximum) of deposits; generally, more acid and lower temperature varieties are superimposed upon and/or overlie higher-T,
less acid types. Conversely, NaCa and low-temperature K
metasomatism can extend over large areas (tens to hundreds
of km2). Although these latter types are broadly linked in time
and space, it can be difficult to relate them to individual
deposits. At regional scales, alkali-rich alteration reflects overlapping systems and, in some regions, cumulatively extends
over distances up to several hundred kilometer (Barton et al.,
2011a; de Jong and Williams, 1995; Menard, 1995; Oliver
et al., 2008; Parak, 1975). Controlling features at the regional
scale include major fault/shear zones and contacts of intrusive
complexes (e.g. Cloncurry, Oliver et al., 2008; Carajas, Xavier
et al., 2010; Chile, Menard, 1995; Sillitoe, 2003).
The vertical extent of IOCG systems is not widely documented, yet it is clear that vertical zoning is the rule. In some areas,
tilted systems or multiple levels of exhumation demonstrate
the vertical zoning summarized in Figure 1. Where such
system-scale observations are available, the vertical extent of
systems are 36 km from hematite  K-rich tops to CuFepoor Na(Ca) roots (e.g., western United States, Dilles et al.,
2000; Johnson, 2000; Johnson and Barton, 2000; Andes,
Kreiner and Barton, 2009). Many ore bodies show a more
limited vertical extent ( 1 km) of metals and alteration assemblages (e.g., many districts in Chile, Kreiner and Barton, 2009;
Marschik and Fontbote, 2001; Rieger et al., 2010, SW North
America, Johnson and Barton, 2000; Barton et al., 2011a;
Carajas, Dreher et al., 2008; Monteiro et al., 2008a; Siberia,
Soloviev, 2010a; Gawler, Bastrakov et al., 2007).
Although there are many paragenetic and zoning studies,
integrated timespace relationships of alteration and magmatism have been addressed only in a few districts (e.g., Barton
et al., 2011b; Dilles et al., 2000; Johnson and Barton, 2000;
Mark et al., 2006b). In some areas, widespread alkali-rich alteration is demonstrably coeval with mineralization as evidenced
by crosscutting relationships and geochronology even in cases
where there are multiple independent hydrothermal systems
(e.g., Marcona area, Peru; Chen et al., 2010b; Copiapo area,
Chile; Barton et al., 2011b). However, in other areas, at least
some of the NaCa and K assemblages may represent independent alteration episodes, alkali-rich diagenetic changes, or metamorphism of evaporitic materials (e.g., Cloncurry, Oliver et al.,
2008; N Baltic, Frietsch et al., 1997; western United States,
Barton et al., 2011b). The common superposition of lower
temperature on higher temperature assemblages can be interpreted as cooling or telescoping of systems, potentially with
significant time lags (e.g., Skirrow, 2010).

Iron Oxide(CuAuREEPAgUCo) Systems

13.20.4.3 Extent of Metasomatism and Comparison with


Porphyry/Alkaline Cu Systems
The intensity and volume of IOCG-related metasomatism
stands out when compared to that seen in most other hydrothermal environments (Figure 8). Alkali exchange alteration
and related metal redistribution commonly exceed 5% of the
mass of the original rock and are documented over tens to
hundreds of kilometer cube in single districts and much more
at regional scales. Mass transfer associated with NaCa alteration typically results in loss of Fe, K, base metals, Rb, and Ba
and addition of Na and/or Ca (Figure 7); comparable masses of
alkalis and some metal exchanges occur during lowtemperature K alteration (Figure 4). The amount of Fe, Cu,
and other metals that can be mobilized in these processes
commonly greatly exceeds the amounts reported in associated
deposits (Barton and Johnson, 2000; de Jong et al., 1998;
Roddy et al., 1988; Williams, 1994b). Volumes of K(CaNa
Fe) and H(Fe) alteration are smaller than NaCa or lowtemperature K types. In the larger systems, they range from 1
to 10 km3, which is roughly comparable to the volumes of
metal-bearing K(NaCa) and acid alteration in porphyry and
arc-related seafloor systems (Figure 8). Globally, only marine
spreading centers have more extensive metasomatism by virtue
of alteration that accompanies ongoing creation and cooling of

TH dilute Magmatic
Marine arcs
1011

(Johnson, 2000)
10 kg
12

TH saline
Ridges

1010
100

10
H+
NaK
MgCa

109
108

1
1

Isotopic
exchange

10

100

1000

10
10 000

(K2O+Na2O+CaO+MgO+FeO*)

1000

100

Absolute mass change (g kg-1 rock)

Durations of IOCG episodes are similarly diverse. Deposits


may be fairly restricted in time (e.g., 16001570 Ma in South
Australia, Skirrow et al., 2007); however, in most areas, temporal
development is protracted. For example, in the Andes (Chen,
2010; Sillitoe, 2003) and western North America (Barton et al.,
2011a), IOCG-style systems range in age from Jurassic through
the Neogene. Even in the same district, hydrothermal activity
may take place intermittently over tens of millions of years (e.g.,
Marcona, Peru, Chen et al., 2010b; Copiapo, Chile, Barton et al.,
2011b). The same holds in older, metamorphosed terranes, such
as the Cloncurry region, where there were at least two major
epochs of metasomatism and mineralization, one pre- or synmetamorphic (>1580 Ma), the other contemporaneous with
1550 Ma batholiths (Mark et al., 2006a,b; Oliver et al., 2008).
Although Hitzman (2000) and other workers have proposed
that Cu-rich systems may form relatively late in any given region,
the accumulating results on regional patterns call into question
any particular age progression; Cu-rich deposits clearly form
earlier, later, or contemporaneously with barren deposits.
Some members of the IOCG clan lack the full range of
hydrothermal features, particularly in the case of the nominally
amagmatic variants that are related to extensional or basinfluid expulsion processes (Barton and Johnson, 2000; e.g.,
SW Unite States detachment deposits, Wilkins et al., 1986;
Wernecke Breccias, Hunt et al., 2007). Such systems contain
relatively little magnetite or high-temperature alteration, yet
they share the lower temperature features notably hematiterich CuAu mineralization. Similar limitations may hold for
the oxide-poor CuAu systems that appear to have a fundamentally metamorphic origin in those cases, shallow or
regionally extensive cogenetic metasomatic features may be
absent (Figure 1; e.g., some deposits in Cloncurry, Oliver
et al., 2008; N Baltic, Billstrom et al., 2010; and Khetri,
Knight et al., 2002; see Haynes, 2000).

% Mass change from protolith

528

Volume of alteration (km3)

Figure 8 Magnitude and volume of mass changes for different families


of hydrothermal systems illustrating the size and intensity of
metasomatism in IOCG systems (Johnson, 2000). IOCG-related systems
are larger than nearly all the others, except for spreading center alteration
or the isotopic exchange associated with shallow magmatism in the
terrestrial crust (line bottom right center; Criss and Taylor, 1986). See
text for discussion.

oceanic crust. In the terrestrial environment, only H and O


isotopic exchange  volatile addition are comparable in volume
though not in mass (Figure 8; e.g., Criss and Taylor, 1986).
Salient metasomatic features of IOCG systems distinctly differ from well-established magmatic-hydrothermal systems, notably porphyry Cu and skarn Cu deposits (Figure 8; Barton and
Johnson, 2000; Williams et al., 2005). IOCGs have tens of cubic
kilometers of NaCa(K) alteration, plus massive concentrations of hypogene Fe oxides, a relative scarcity of quartz, and
the distinctive suite of accessory REE, Co, and U. In contrast,
porphyry Cu systems, even alkaline varieties, lack the large volumes (in many examples, any) of Na-rich alteration. Where
present, NaCa metasomatism is either superimposed and
related to external fluids (Dilles et al., 1995) or forms limited
volumes (generally 1 km3) of various Na(KCaFeH)
types that are localized on the intrusive centers (Seedorff et al.,
2005). In contrast to IOCGs, in calcalkaline porphyry systems,
quartz veins are abundant, CuFe sulfides commonly form early,
and later acid assemblages are voluminous (110 km3) and
sulfur-rich (Seedorff et al., 2005). Magnetite-bearing veins
(quartz  amphibole), though widespread, rarely occur over
large volumes (e.g., >1 km3) nor do magnetite additions average
more than a few wt% over these volumes (Clark and Arancibia,
1995; Muntean and Einaudi, 2000; Proffett, 2003). Magnetite
(and lesser hematite) can be relatively abundant in skarns associated with porphyry systems, but with the exception of dolomitic hosts, oxides are subordinate to iron-rich calcsilicates, such
as andraditic garnet (Einaudi et al., 1981) thus, these too differ
from most IOCG carbonate-hosted mineralization. These differences are subdued in alkaline porphyry systems that, due to their
distinctive igneous chemistries, commonly have less quartz, less
acid alteration, and more pronounced K(CaNa) alteration
(Cooke et al., 2007; Jensen and Barton, 2000; Lang et al.,
1995). The extreme cases are carbonatite-related deposits that
have been grouped by some with IOCGs (e.g., Groves et al.,
2010; Hauck, 1990; Hitzman et al., 1992); carbonatites have
different trace element enrichments (e.g., high Nb), have

Iron Oxide(CuAuREEPAgUCo) Systems

strongly peralkaline alteration (fenites), and with the exception


of Bayan Obo, China, lack abundant massive Fe oxides.
Of course, completely independent origins are not
required. Hybrid districts reflect spatial overlap of porphyrytype and IOCG-type systems with mixed fluid sources (Barton
and Johnson, 2000). The Yerington district, Nevada, is the
preeminent example (Dilles and Proffett, 1995; Dilles et al.,
2000) and other districts show similar superimposed features including a number of Andean occurrences (e.g., Barton
et al., 2011b; Dow and Hitzman, 2002; Tornos et al., 2010)
and the deformed AitikNautanen trend deposits in Sweden
(Wanhainen and Martinsson, 2010).

13.20.5

Petrologic and Geochemical Characteristics

Study of IOCG systems over the last 15 years provides evidence


from fluid inclusions, isotopic analyses, whole-rock characteristics, trace elements, and petrology on the compositions of the
fluids, the degree and character of mass transfer, the conditions
of formation, and the possible sources of components. Recent
reviews are provided in Williams et al. (2005, 2010) and Porter
(2010a), and for selected regions including Cloncurry (Mark
et al., 2006a,b; Oliver et al., 2008), Gawler (Bastrakov et al.,
2007; Skirrow, 2010), N Baltic (Billstrom et al., 2010), Andes
(Chen, 2010), Wernecke (Hunt et al., 2007, 2011), and Carajas
(Torresi et al., 2012; Xavier et al., 2010). This section briefly
reviews the data and notes the principal inferences drawn from
each. Genetic linkages can be problematic and are discussed in
the following section.
Table 2

13.20.5.1

529

Conditions of Formation

Geologic and petrologic evidence indicates that most, though


not all, IOCG systems form in the upper crust, over a wide range
of temperatures, and from sulfur-poor, oxidized NaCaK(Fe)
brines with or without a second sulfur- or metal-bearing fluid.
Table 2 summarizes ranges in depth, temperature, and fluid
compositions for selected regions.

13.20.5.1.1 Depth
As originally summarized by Hitzman et al. (1992) and evaluated in many subsequent papers, most IOCG systems form in the
upper crust, with mineralization commonly within a few kilometer of the paleosurface but with alteration extending to greater
depths. Some systems, particularly those in metamorphic terranes, show evidence of formation in the 520 km range.
The evidence for shallow formation (see Table 2 for areas)
comes from direct evidence for proximity to the paleosurface,
stratigraphic and structural reconstructions, metamorphic and
igneous assemblages, and fluid inclusion observations including
evidence for boiling. Many IOCG systems are hosted in broadly
contemporaneous volcanic and sedimentary rocks. Stratigraphic
reconstructions preclude more than a few kilometer overburden
(e.g., Punta del Cobre, Marschik and Fontbote, 2001). In a number of districts, penecontemporaneous syngenetic iron formation
overlies discordant mineralization (Barton et al., 2011b; Hitzman
et al., 1992; Johnson and Barton, 2000; Seeger, 2000; Tobey,
1976). Some massive stratiform magnetite- or hematite-rich
bodies, such as Kiruna, Bafq, and El Laco, have been interpreted to form at, or very near the, contemporary surface (e.g.,
Jami et al., 2007; Naslund et al., 2002; Parak, 1975; Sillitoe and

Synopsis of ranges in P, T, and fluid compositions for selected IOCG regions

Region

P (kb)/T ( C)a

Principal fluid compositions (wt%


NaCleq, cations, CO2)

Comments

Key recent references

Andes, coastal
Chile and Peru

<0.51.5/150550
(>800b)

Australia, Gawler

<0.51.5?/150540

650% Na(Ca  Fe  K  Mg);


V-rich (boiling), late fluids dilute,
CO2 (melt inclusionsb)
223%, >25% Na(CaKFe, minor
Cu), rare CO2, lower-T fluids dilute

Commonly high-T saline,


low-T dilute (melt in
Mt-Ap-Actb)
Unmixing/boiling (Moonta,
Olympic Dam)

Australia,
Cloncurry

1.54/220550

540% Na(Ca); 3070% Na(Ca


KFe, variable Cu); CO2-rich

Australia, Tennant
Crk
Brazil, Carajas

2.55(?)/300400

1020% Na(Ca) [Fe]; 110/20


35% [Cu], variable CH4, CO2
Mostly 2550% Na(CaKFe) mix
with cooler dilute; minor CO2
1535% Na(KCa); minor CO2

Multiple populations reflect


complex history and
contrasting systems
Fe precedes CuAuBi,
mixing inferred
Mostly shallow; Salobo
deep/metamorphosed
Fluid data for five element
veins petrography

Also melt inclusions,


multiphase solid-rich
Mix of metamorphic and
premet systems
Jurassic and Tertiary
Fe  Cu

Broman et al. (1999), Chen


(2010), Kreiner (2011), and
Velasco and Tornos (2009)
Bastrakov et al. (2007),
Morales-Ruano et al. (2002),
and Oreskes and Einaudi (1992)
Baker et al. (2008), Fisher and
Kendrick (2008), and Rusk et al.
(2010)
Skirrow and Walshe (2002) and
Zaw et al. (1994)
Torresi et al. (2012) and Xavier
et al. (2009, 2010)
Changkakoti et al., 1986 and
Mumin et al., 2010
Hunt et al. (2011)
Soloviev (2010a)

Canada, Great Bear

<1? (Salobo, >2?)/


150570
0.51.5?/150450

Canada, Wernecke
Russia, Siberian
trap
N Baltic

0.42.4/185350
<0.51.5/130420
(rarely >600)
<13?/120500

United States,
western

<0.51.5/150550

2442% NaCa, no CO2


2747% Na(CaFe), no CO2 nor
V-rich (boiling?), solid-rich
1245% Na(CaK), minor
CO2  methane
2550% Na(Ca  Fe  K), no CO2

Billstrom et al., 2010 and Gleeson


and Smith, 2009
Barton et al. (2011a), Johnson
(2000), and Wilkins et al. (1986)

Pressure and temperature ranges are based on reported petrologic and geologic estimates for pressure and reported fluid inclusion homogenization temperatures, isotope
partitioning, and phase equilibria for temperature.
b
Data reported for selected MtApAct rocks.

530

Iron Oxide(CuAuREEPAgUCo) Systems

Burrows, 2002). Locally, very fine-grained (10 mm) hydrothermal minerals, such as actinolite and chalcedonic silica, point
to rapid formation and subsequent lack of annealing that
would be expected with depth (e.g., Punta del Cobre and Santo
Domingo Sur, Chile, Barton, unpublished observations). Incorporation of syngenetic mineralization in breccias also implies
very shallow formation, for instance, at Olympic Dam (McPhie
et al., 2011; Oreskes and Einaudi, 1990) and Punta del Cobre
(Barton et al., 2011b). Relatively little quantitative barometry
has been done; however, fluid inclusion data locally demonstrate boiling of CO2-poor fluids thus requiring pressures <1 kb
(Table 2). The Salton Sea geothermal system demonstrates that
IOCG-style mineralization can form <1 km below the modern
surface (McKibben and Hardie, 1997).
Deeper levels of formation are inferred in areas with protracted metamorphic histories, particularly where deformation
and metamorphism may have accompanied ore formation. At
Tennant Creek and in some Cloncurry deposits, CO2-rich fluid
inclusions and host mineral assemblages indicate pressures of
1.54 kb (e.g., Baker et al., 2008; Skirrow and Walshe, 2002).
These estimates are compatible with regional metamorphic
assemblages and are among the deepest reported for large
magnetite-rich systems. Among the clearest examples of metamorphic deposits are some in Rajasthan (Knight et al., 2002),
N Baltic (Billstrom et al., 2010; Ettner et al., 1993), and
Cloncurry (Oliver et al., 2008). Not all metamorphic-hosted
deposits necessarily formed at higher pressure as evidenced
by regions with both metamorphosed (contact and regional)
and unmetamorphosed examples in the same area (contact
overprint Candelaria, Chile, Ryan et al., 1995; regional overprint on Salobo, Carajas, Xavier et al., 2010; regional overprint on a number of the older Cloncurry deposits, Oliver
et al., 2008).

13.20.5.1.2 Temperature
Temperatures of formation range from >500  C in magnetite
apatiteactinolite(or clinopyroxene) associations down to
200300  C for hematite-dominated mineralization (Table 2).
Many deposits show superposition of lower temperature on
high-temperature assemblages. Temperatures exceeding
800  C are inferred in a few magnetiteapatiteactinolite
deposits based on fluid inclusion observations and isotopic
thermometry (e.g., Broman et al., 1999; Chen et al., 2010a;
Velasco and Tornos, 2009). Fluid inclusion homogenization
temperatures are most widely cited yet are challenging in IOCG
systems because of the paucity of suitable hosts and because
postentrapment changes and superimposed events can obscure
links to earlier features. Other mineralogical and isotopic
thermometers have been applied with mixed results. In many
cases, they yield similar ranges (e.g., 250400  C with chlorite
thermometry) and are generally compatible with limits on
mineral stabilities (e.g., scapolite >400  C; Pan, 1998). Brittle
ductile behavior, implying temperatures in the 300400  C
range, is observed in a number of deposits including some
not only in metamorphic terranes (Oliver et al., 2008) but
also in shallow veins (Kreiner, 2011). Isotopic thermometers
(mainly S and O) reported in a number of papers yield mixed
results, plausible in many cases but unrealistic in others,
presumably due to lack of cogenesis, original disequilibrium,
or resetting.

13.20.5.1.3 Fluid inclusion compositions


Fluid inclusion studies by numerous authors show that Na
CaK(Fe) brines predominate in IOCG systems, with or without vapor-rich inclusions (see Table 2, recent summaries in
Chen, 2010; Torresi et al., 2012; Williams et al., 2010). Total
salinities in aqueous inclusions vary from 1 to >80 wt%
NaCleq; most are in the 1050 wt% range. Fluid inclusions
commonly have not only multiple daughter minerals, generally halite, but also other K and Ca salts, various Fe minerals
(FeClx, hematite, ferropyrosmalite), and occasionally carbonates (calcite, nahcolite). Consistent with the low sulfur mineral
assemblages, chalcopyrite daughter crystals are rare in contrast
to brine inclusions from porphyry copper deposits. Also contrasting with the typical situation in porphyry copper deposits
(Chapter 13.5), salt-rich fluid inclusion assemblages in IOCGs
commonly lack vapor-rich inclusions. CO2 is a widespread
minor component (Table 2) as it is in other deposit types.
Liquid CO2-bearing inclusions have been noted in a number
of areas, particularly in deposits found in metamorphosed
terranes (e.g., Pollard, 2001). Methane is reported in some
metamorphic-hosted deposits, particularly those with reduced
mineral assemblages (e.g., Tennant Creek, Zaw et al., 1994;
Salobo, Torresi et al., 2012; Cloncurry, Oliver et al., 2008).
Although early studies suggested that high CO2 contents might
be characteristic of IOCG deposits (e.g., Pollard, 2000), many
systems lack CO2-rich inclusions including most of those that
are not metamorphosed (Table 2). It remains to be seen how
many other IOCG deposits have moderate CO2 contents and if
the contents are comparable, for instance, to those found in
epithermal precious metal deposits (Chapter 13.5). A handful
of magnetiteapatiteactinolite occurrences have nonhomogenizing (exceptionally high T?) hydrous saline/silicate-rich
inclusions, which have been interpreted as evidence for coexisting melts at the time of mineralization (Broman et al., 1999;
Hou et al., 2010; Velasco and Tornos, 2009).
Microthermometry and direct analysis of brine inclusions
using PIXE, crush-leach, and LA-ICP-MS reveal complex, commonly metal-rich fluids. Abundant Ca and other divalent cations are widely indicated by low first melting temperatures.
Both the heating/freezing experiments and direct analyses indicate that IOCG fluids have significantly higher Ca/Na and
lower K/Ca than porphyry copper fluids (Bastrakov et al.,
2007; Rusk et al., 2009; Xavier et al., 2009). In addition to
the evidence from daughter minerals, direct analyses show that
many fluids contain Na, Ca, K, and Fe well in excess of 1 wt%
(Fe can exceed 10 wt%) and contain Mn, Ba, Zn, Pb, As, and Sr
in the 0.11 wt% range. Analyses from both Cu-rich mineralized systems and Cu-poor deposits show appreciable Cu,
mostly in the 1005000 ppm range (Baker et al., 2008;
Niiranen et al., 2005; Rusk et al., 2009; Williams et al.,
2001). A handful of areas, generally associated with granitic
intrusions, yield higher values, in excess of 1% Cu (Baker et al.,
2008; Bastrakov et al., 2007; Perring et al., 2000).
Overall, brine compositions are similar between Cu-rich
and Cu-poor IOCG systems (Rusk et al., 2009; Williams
et al., 2010). What is distinct is that there is commonly evidence for later, lower temperature, lower salinity (115 wt%
NaCleq) fluid that mixes with or is superimposed upon more
saline brines (e.g., Bastrakov et al., 2007; Chen et al., 2010a,b;
Oreskes and Einaudi, 1992; Torresi et al., 2012). This fluid is

531

Iron Oxide(CuAuREEPAgUCo) Systems

often invoked to bring sulfur and/or metals into otherwise


barren iron-rich mineralization (e.g., Chen, 2010; Gow
et al., 1994; Haynes et al., 1995; Skirrow, 2010). Distinct
fluid inclusion populations are common and are widely interpreted to reflect multiple fluid sources (e.g., see Baker et al.,
2008; Chen, 2010).

aqueous sulfate must also be low based on the evidence for


substantial Ba mobility demonstrated by whole-rock data
(Figure 7) and supported by high Ba contents measured in fluid
inclusions (Williams et al., 2001). Furthermore, if sulfate were
particularly high, it would be limited on cooling to moderate to
low temperatures through sulfidation of magnetite by reactions,
such as 4H 2SO42 12Fe3O4 18Fe2O3 FeS2 2H2O
(Figure 9(c), succeeding text; Gow et al., 1994).
The exception to these generalizations is the reduced,
pyrrhotite-bearing, variably magnetite-poor deposits hosted
in metamorphosed terranes, such as a number of those (but
not all) in northern Australia (Tennant Creek, Cloncurry),
N Baltic, and India (Khetri, Singhbhum). In many districts,
pyrrhotite appears to be primary; however, in other cases,
pyrrhotite may be a metamorphic product as it is at Candelaria
where pyrrhotite occurs only in the contact metamorphosed
portion of a more extensive, pyrrhotite-free district (Marschik
and Fontbote, 2001; Ryan et al., 1995).

13.20.5.1.4 Oxidation state, sulfidation state, and


total sulfur
The mineral assemblages and temperatures in IOCG systems
require that mineralization occurred under conditions that
were relatively oxidized (SO42 > H2S) and low in total sulfur.
In comparison, most porphyry Cu systems are more sulfur-rich
and somewhat more reduced (SO42  H2S). The key indicators
of low sulfidation state coupled with relatively oxidized conditions are (1) the near absence of sulfides in high-temperature
magnetite-bearing assemblages, (2) relatively low Fe/Mg in
most hydrothermal silicates (Mazdab, 2001), (3) the presence
of bornite (or digenite) with hematite rather than magnetite,
(4) the overall scarcity of Fe(Cu) sulfides including in late-stage
intense hydrolytic alteration (Kreiner, 2011), (5) the rarity of Zn
and Pb sulfides in the deposits in spite of appreciable fluid
contents (Rusk et al., 2009), and (6) distal hematite-bearing
assemblages that in some cases have arsenides and other minerals indicative of unusually low sulfur fugacity.
The presence of sulfates (anhydrite, barite) in some deposits
indicates either moderate aqueous sulfate contents or, more likely,
mixing with a second sulfate-bearing fluid. In many systems,

Phanerozoic
marine

aq.

Considerable work has been done on IOCG systems using


tracers including stable isotopes (H, B, C, O, S, Cl), radiogenic
isotopes (RbSr, ReOs, SmNd, UThPb), and noble gas
(He, Ne, Ar) and halogen (Cl, Br, I) ratios (see recent synoptic
reviews in Williams et al., 2005, 2010). Collectively, the data
can be interpreted in terms of many possible sources, often two
or more, within the same system. This diversity contributes to

Proterozoic/Archean

sed

Sulfates
Sulfides

sed

Temperature range reported for most


sulfide deposition in IOCG systems

D = d 34Ssulfide - d 34SSSO4

oxidized
reduced

10

SSO4

IOCG veins, skarn

sed

PCD, Cu skarn

d 34Smin

15

Tracer Studies and Sources of Components

sed

lacustrine

20

minerals

25

13.20.5.2

ign

Fractionation from fluid


with uniform aqueous SSO42

-5
-10

Cpy

-15

Py

H2S

Sulfide partitioning from Ohmoto and


Rye (1979); sulfatesulfide partitioning
from Ohmoto and Lasaga (1982)

-20

(b)

-25
200

(c)

250

300
T(C)

350

400

D = d 34SH2S - d 34Sfluid

log SSO42/SH2S

M(P)/(C)

H/(C)
Wernecke Breccias, YT, Canada

MH/C

Hannuikainen, Rautuvaara, Finland

MH/C

Boss Bixby, MO, USA

H(M)/C

MH/C

Salobo, Sossego, Igarape Bahia, Carajas, Brazil

MH/C

Moonta-Wallaroo, SA Australia

Wirrda Well, Olympic Dam, Oak Dam, SA, Australia

MHP /C

Emmie Bluff, SA, QLD, Australia

MH/C
Ernest Henry, QLD, Australia

Starra, Osborne, QLD, Australia

MH/C
MP/C

M(H)/(C)

M(H)/(C)
Mt skarns, Urals/Turgai, Russia/Kazakhstan

Tennant Creek, NT, Australia

M(H)/(C)

Siberian trap Mt breccia pipes, Russia

M(H)/(C)

Mont-de-dAigle, QC, Canada

Divrigi, Turkey

HM/C

HM/C

Candelaria, Punta del Cobre, Ojancos Viejo, Chile

EI Laco, Cerro Negro Norte, Los Colorados, Chile

HM/C
Mantoverde, Santo Domingo, Teresa de Colmo, Chile

MH/C

M(H)/(C)

MH/(C)

MHP/C
Raul-Condestable, Peru

Marcona, Mina Justa, Peru

Longquiao, Yangtze region, China

MH/(C)
Humboldt, NV, USA

Cornwall, Grace, French Creek, PA, USA

HM/(C)
Cortez Mtns, NV, USA

M/C

Yerington, NV, USA

H/(C)
Salton Sea Geothermal System, CA, USA

M/C
Porphyry copper systems

Sedimentary sulfates

(a)

(individual deposits typically less)

-10

MH(P)/C

sed

-5

-25

-20

-15

Hm
Mt

-10

Increasing SS, H+

Mt
Py

Bn+
Py

-5
-1

Cpy

-0.1
Mineral equilibria modified after Ohmoto and Rye (1979)
for neutral pH, m + = 0.5, mCa2+ = 1.0, mMg2+ = 0.25
Na
Sulfatesulfide partitioning from Ohmoto and Lasaga (1982)

-5
200

250

300

Py
Po

350

400

T(C)

Figure 9 Sulfur isotopes in IOCG systems. (a) Ranges of d34S for sulfides and sulfates summarized from the many papers cited in the text for these
districts. Key FeSO minerals (bottom of range) are M, magnetite; H, hematite; P, pyrrhotite. (b) Fractionation between sulfide and sulfate over
temperature range typical of CuFe sulfide deposition in IOCG systems and compared with (c) showing calculated fractionation between sulfide and fluid
as a function of temperature and sulfate to sulfide and related mineral stabilities (reproduced from Ripley EM and Ohmoto H (1977) Mineralogic, sulphur
isotope, and fluid inclusion studies of the stratabound copper deposits at the Raul Mine, Peru. Economic Geology 72: 10171041). See text for
discussion.

532

Iron Oxide(CuAuREEPAgUCo) Systems

the range of interpretations. Although results are reviewed here


by tracer type, bear in mind that most deposit and district
studies combine data sets to draw more specific and often
more restrictive (but commonly nonunique) inferences (e.g.,
see Williams et al., 2010).

13.20.5.2.1 Light stable isotopes: H, O, S, C, and B


Light stable isotope data for IOCG systems have been summarized in a number of papers including Williams et al. (2005,
2010), Mark et al. (2006a,b), Bastrakov et al. (2007), Chen
(2010), and Torresi et al. (2012).
Oxygen and hydrogen isotopes have rock (or magma)dominated signatures with or without a later meteoric or
low-T sedimentary signature (e.g., Carajas, Torresi et al., 2012;
Xavier et al., 2010; Gawler, Bastrakov et al., 2007; Oreskes and
Einaudi, 1992; Andes, Chen, 2010; Rieger et al., 2012; Great
Bear, Reardon, 1992). The nonigneous signatures may correlate
with the second fluid observed in many fluid inclusion studies.
Calculated fluid d18OSMOW mainly ranges from 5 to 11%
but extend to light values (5 to 5%) particularly in the later
stages of mineralization. Similarly, dDSMOW values cluster
from 60 to 20%, but overall, they span 0% to 140%.
Most data are for mineralization and wall rock alteration in
the deposits themselves; however, extensive isotopic exchange
also accompanies the system-scale alteration, notably the
extensive metal-poor Na(Ca) and K metasomatism including
red rock alteration (SE Missouri, Wenner and Taylor, 1987;
W United States, Roddy et al., 1988; Solomon and Taylor,
1993). This exchange is similar in scale to other areas of
terrestrial O and H exchange (e.g., Criss and Taylor, 1986).
Overlapping compositions preclude unambiguous interpretation of the source of water from d18O and dD. Many
studies interpret the intermediate dD, rock-dominated d18O
values as representing magmatic sources; however, magmatic
and metamorphic water compositions overlap with compositions for mid-latitude surface fluids or basinal brines that
have exchanged with rock oxygen at elevated temperatures
(Sheppard, 1986). Most studies infer a nonmagmatic water
source for the more extreme dD values (<80%, >30%);
however, in an alternative interpretation, Mark et al. (2004)
proposed that large changes in dD at constant d18O in Cloncurry regional NaCa alteration reflect H isotope fractionation
during magma degassing.
Sulfur isotopes are of particular interest given sulfurs essential role in precipitation of Cu- and the S-limited nature of the
hydrothermal fluids. Sulfide d34SCDT ranges from 10 to 30%
(Figure 9; selected summaries in Bastrakov et al., 2007;
Benavides et al., 2007; Chen, 2010; Davidson and Dixon,
1992; Rose et al., 1985; Torresi et al., 2012; Vakhrushev et al.,
1981; Zhang, 1986). Although median values cluster near
0  5%, heavier values occur in many districts, and many
deposits show spreads of 520%. Late sulfides commonly
have higher d34S than early sulfides (e.g., Chen, 2010), and
in a few districts, there is evidence for zoning outward to
higher d34S (Raul-Condestable, de Haller and Fontbote, 2009;
Mantoverde, Rieger et al., 2010). A number of Cu-poor systems
have uniformly high d34S (>5%, commonly >10%; see
Figure 9(a)). Rarely, late sulfides have d34S <  20%, indicating involvement of biogenic sulfur (e.g., Raul-Condestable,
de Haller and Fontbote, 2009; Ripley and Ohmoto, 1977).

When present, hydrothermal sulfates (anhydrite, barite) are


820% heavier than the sulfides (Figure 9(a)); such sulfate
sulfide differences are broadly consistent with fractionation
at the inferred temperatures of 250400  C (cf. Figure 9(b)).
Published interpretations of bulk fluid d34S and hence sulfur sources vary; yet all must be considered in light of the redox
control on fractionation (Ohmoto and Rye, 1979). Many
investigators interpret values near 0% as indicative of an
igneous origin for sulfur (nominally magmatic), which is
plausible if conditions are relatively reduced and aqueous
sulfur is sulfide (H2S, HS)-dominated (Figure 9(c)). Sulfidedominant conditions are reasonable for the few IOCG-clan
deposits that have hydrothermal pyrrhotite and, possibly, for
a handful of other magnetite-bearing assemblages, particularly
at high temperatures (Figure 9(a); Chen, 2010); however, as
described earlier and noted on Figure 9(a), the majority of
IOCG systems contain hematite or mushketovite and other
indications of higher oxidation states. Under these conditions,
sulfate species (SO42, HSO4, NaSO4, etc.) will dominate,
particularly in brines where additional sulfate complexes
become important (Ohmoto and Rye, 1979; Ripley and
Ohmoto, 1977; see Figure 9(c)). Thus, it is more likely that
in many systems, the bulk fluid is sulfate-dominated and
515% heavier than coexisting sulfides (e.g., Bastrakov et al.,
2007). Consequently, typical d34Sfluid would be 10 to 20%
(e.g., modern Salton Sea fluids, McKibben and Hardie, 1997).
The shift from relatively light to heavier d34S that is observed in
many Andean deposits has been interpreted either as progressive evolution of a single source or as mixing of several sources
with an early igneous source and a later marine/evaporitic
component (Benavides et al., 2007; Chen, 2010; de Haller
and Fontbote, 2009; Rieger et al., 2012). Summing up, the
d34S results require multiple sources, some of which are isotopically heavy, others are near 0%, and a few may have quite
light values. This variability occurs within the same regions,
districts, and even some deposits (e.g., Andes, Chen, 2010;
Gawler, Bastrakov et al., 2007; Carajas, Torresi et al., 2012).
In contrast to sulfur, carbon isotopes in IOCG systems are
not particularly illuminating even when combined with other
isotopic data, such as d18O. Most hydrothermal carbonates
have d13CPDB between 0 and 10% and there is a broad
positive correlation of d13C with d18O. Values commonly
span large ranges within districts (e.g., 21 to 3, Copiapo,
Kreiner, 2011; 10 to 1, Wernecke, Hunt et al., 2011; 15 to
2, Carajas, Torresi et al., 2012; 8 to 11, Misi region,
Finland, Niiranen et al., 2005). Carbonate-hosted deposits
show systematic decreases in both d13C and d18O from marine
values toward lighter values. These changes reflect combinations of decarbonation and involvement of isotopically lighter,
permissibly magmatic fluids. In a handful of examples,
d13C <  10% implies involvement of reduced carbon, likely
from organic materials or graphite (e.g., Khetri, Deb and
Sarkar, 1990; Copiapo, Kreiner, 2011).
A few recent studies have examined boron isotopes that
have the potential to fingerprint certain fluid sources. Tourmaline d11B in IOCG systems ranges from 10 to 25%
(Figure 10). Ratios in some regions exhibit nearly as much
variation as has been reported globally, although individual
deposits commonly show less spread (Carajas, Xavier et al.,
2008; Singhbhum, Pal et al., 2010; Chile, Tornos et al., 2012;

Iron Oxide(CuAuREEPAgUCo) Systems

-20

-10

10

20

533

30

Carajas, Brazil
(Xavier et al., 2008)

Singhbhum, India
(Pal et al., 2010)

Chilean Iron Belt


(Collins, 2010; Tornos et al., 2011;
Mazdab and Barton, unpubl. data)

Jurassic W USA
IOCG and NaCa
(Johnson and Barton., 2010;
Mazdab and Barton, unpubl.)

Tourmalinefluid

f fractionation at 350C
Sedimentary and
metamorphic rocks
Igneous rocks and magmas

Marine evaporites and brines


Nonmarine evaporites and brines

-20

-10

10

20

30

d11B ()
Figure 10 Boron isotope ratios for tourmalines from worldwide IOCG districts, for representative reservoirs, and tourmaline-aqueous borate
fractionation at 350  C (Palmer et al., 1992; cf. Meyer et al., 2008). See text for discussion.

Collins, 2010). Many analyses cluster in an ambiguous range


around 0  10%, which is somewhat heavier than typical magmatic values and which overlaps with other crustal materials
(Kasemann et al., 2004; Palmer and Swihart, 1996). Given that
tourmaline is 38% heavier than aqueous boron (Meyer et al.,
2008; Palmer et al., 1992), most IOCG-related fluids had
d11B > 0% (Figure 10).

13.20.5.2.2 Radiogenic isotopes: Sr, Nd, Pb, and Os


Radiogenic isotopic signatures in IOCG systems have been
investigated as tracers in addition to their use as geochronometers (see Chapter 13.4). Many initial isotope ratios resemble those of associated igneous rocks; however, some notable
differences have been observed.
Nd isotopes reveal mixed sources. Most commonly Nd
isotope ratios resemble their host rocks as reported from
North America (Gleason et al., 2000) and South Australia
(Skirrow et al., 2007). The exception is Olympic Dam, where
eNd in the CuAu(REE) exceeds that of the host rocks and
magnetite mineralization and is interpreted to reflect an
additional contribution from mafic rocks (Johnson and
McCulloch, 1995). Nd isotopic evidence points to a syn- or
postmetamorphic source for REEU mineralization at Mary
Kathleen at Cloncurry (Maas et al., 1987), which although
not an IOCG deposit, has many features in common and is
broadly contemporaneous with nearby IOCG systems (Oliver
et al., 2008). Similarly, Nd and Sr data from Finnmark,

Norway, evidence a crustal source in this metamorphic deposit


and are interpreted to reflect crustal metal and fluid sources
(Bjorlykke et al., 1990).
In Chile, where there is a large isotopic contrast, Sr isotopes
demonstrate mixing between igneous and external Sr. Fresh
igneous rocks have initial 87Sr/86Sr of 0.7033  0.0002,
whereas altered igneous rocks and mineralization range from
0.7037 to 0.7062 (Barton et al., 2005; Chiaradia et al., 2006;
Kreiner, 2011). Similar evidence for introduction of nonigneous Sr occurs in intrusion-hosted IOCG-related mineralization in the Yerington and Iron Springs districts of the western
United States (F.K. Mazdab, written comm., 2012). Yu et al.
(2008) interpret initial 87Sr/86Sr of 0.70710.7073 in apatite
from Ningwu district magnetite ores to reflect an igneous
source from variably crustally contaminated gabbrodiorites
(0.70450.7077). Given the abundance of alteration in the
Yangtze IOCG-associated igneous rocks, a mixed hydrothermal
source seems equally plausible. Pb isotope ratios have been
determined for a number of South American deposits. The
results, though somewhat complex in detail, are all compatible
with igneous sources for Pb (de Haller et al., 2006; Galarza
et al., 2008; Marschik et al., 2003; Rieger et al., 2010).
Sparse Os data suggest several possible sources. In the most
extensive study to date, Mathur et al. (2002) found that there
was evidence for both igneous and nonigneous Os sources in
Chile. Magnetiteapatite deposits, such as El Romeral, yielded
elevated initial 187Os/188Os (1.48), whereas Cu-rich systems

534

Iron Oxide(CuAuREEPAgUCo) Systems

and igneous rocks at Candelaria and Manto Verde yielded


more primitive ratios (0.20.4). They interpreted the mantlelike ratios in the Cu(Au) deposits to indicate a magmatic
source. McInnes et al. (2008) reported sulfide and whole-rock
isochron ages, respectively, at Olympic Dam and Tennant
Creek that are considerably younger than the host rocks.
Although resetting seems likely at Olympic Dam, in both
cases, evidence points to relatively primitive sources for Os
and precludes major involvement of older radiogenic crust.
Conversely, Kontak et al. (2008) found elevated initial
187
Os/188Os of 1.0  0.4 in orogenic IOCG-like mineralization
in Nova Scotia, suggestive of a crustal source.

13.20.5.2.3 Halogens and noble gases: Ratios and isotopes


Halogen (Cl, Br, I) and noble gas (Ar, He, Ne, Xe) ratios and
isotopes can potentially fingerprint the source(s) of these
mainly conservative elements (i.e., with the exception of
scapolite  apatite, they are confined to the fluid). Halogen
ratios have been most widely applied as determined from
PIXE, crush-leach, and thermal decrepitation studies of fluid
inclusions. Observations, summarized in Figure 11, show that
the Br/Cl spans a broad range from ratios typical of residual
brines (bitterns), though nominally magmatic values to values
consistent with dissolution of halite evaporites (Cloncurry,
Kendrick et al., 2008a, 2011; Perring et al., 2000; Andes;
Chiaradia et al., 2006; Carajas, Xavier et al., 2009; Wernicke,
Kendrick et al., 2008b; Sweden, Gleeson and Smith, 2009). No
simple pattern emerges even within single regions, and even
single deposits can show a fairly wide spread in values (e.g.,
Williams et al., 2010). Although extreme values (exceptionally
high or low Br/Cl) may point to particular evaporitic sources,
most data overlap with ratios seen in basinal fluids and

magmas (Figure 11). Exchange reactions with rock govern


Na/Cl and similar nonconservative cation/anion ratios and
thus reflect paths rather than ultimate source. In addition to
halogen ratios, d37Cl has been measured in a number of South
American and Swedish districts (Figure 11(b)). In some cases,
d37Cl  0% and thus is consistent with either seawater (or
evaporite) or magmatic values. The negative values found in
some of the Swedish metamorphic-hosted deposits suggest
that other fractionation processes may be involved (Williams
et al., 2010).
Recently, I/Cl, 40Ar/36Ar, and other noble gas measurements have been undertaken by Kendrick and others on Cloncurry (Fisher and Kendrick, 2008; Kendrick et al., 2008a,
2011), Wernecke (Kendrick et al., 2008b), and Chilean
deposits (Marschik et al., 2011). 40Ar/36Ar varies from nearatmospheric values (296) in Chile (300420) to much less
radiogenic values in the Proterozoic deposits (<1000 to
>30 000). Measured I/Cl has higher values than would be
expected from simple evaporation of seawater or dissolution
of halite but overlap with fields for basinal brines (see
Chapter 7.14). The combined results, together with halogen
ratios, point to mixed sources of solutes sedimentary brines
plus deeper components. In the case of the noble gases, this
includes a deep magmatic or metamorphic signature in addition to a component of crustal or near-surface fluids.

13.20.6 Summary of the IOCG Clan, Likely Origins,


and the terrestrial Hydrothermal Environment
The recognition of IOCG systems as a distinct group reflects a
shared geochemistry and a common set of geologic features
that serves to distinguish them from other major classes of

100

10

Carajas - Sossego, Gameleira


Andes - Candelaria, Raul-Cond.
Andes - Romeral
Sweden - IOCG-low Cu
Sweden - IOCG-high Cu
Sweden - hybrid Cu(-Au)

High T processes
(dissolution, boiling etc.)

Scapolite-bearing
alteration, Copiap

Basinal brines

Br/Cl x 103 (wt)

Bittern brines

10

Basinal brine d 37Cl

Br/Cl x 103 (wt)

25 C halite
saturation

100

Mantle
Magmatic
1

1
Wernercke Bx - Hm-Cu
Cloncurry - Ernest Henry
Cloncurry - Lightning Crk
Sweden - IOCG-low Cu
Sweden - IOCG-high Cu
Sweden - hybrid Cu(-Au)
0.1
(a)

1.0

Salton Sea
geothermal
system

Metamorphic
brines?

Halite
10
CI (wt%)

-4.0

100
(b)

-3.0

-2.0

-1.0

0.0

1.0

2.0

0.1
3.0

d 37CI ()

Figure 11 Summaries of (a) Br/Cl versus Cl for IOCG fluids (see Williams, 2010b, for sources) and other reservoirs including basinal fluids (Kharaka
and Hanor, Chapter 7.14) and (b) Br/Cl versus d37Cl for Swedish (Gleeson and Smith, 2009) and Chilean (Chiaradia et al., 2006) IOCG systems.

Iron Oxide(CuAuREEPAgUCo) Systems

ore-forming systems. Key among these are concentrated hypogene iron oxides, low sulfide contents, a distinctive family of
trace elements (Cu, Au, P, REE, Co, U), and voluminous Na
CaKFeH hydrothermal alteration. Like the other major
groups, such as marine-hydrothermal systems or the porphyry/
greisen clan, IOCG systems exhibit considerable diversity in
detail but share system-scale patterns and geologic context.
Within the IOCG clan, alkalic alteration is ubiquitous and
voluminous, yet the assemblages and forms of deposits vary
considerably, notably with regard to the abundance of Cu, Au,
and U. Likewise, contemporaneous magmatism is commonly
present but is compositionally diverse precluding an essential
role for any particular type. Geologic settings span a broad range
of active tectonic regimes but widely share evidence for evaporitic materials. The principal common genetic factor is an
unequivocal association with a distinctive range of sulfur-poor,
high-salinity fluids across the IOCG spectrum.

13.20.6.1

Discussion of Genesis

Genetic interpretations span a range of origins (Figure 1)


including a possible role for immiscible melts. With the exception of those who advocate an immiscible melt contribution to
magnetiteapatiteactinolite deposits (e.g., Chen et al., 2010a;
Naslund et al., 2002; Velasco and Tornos, 2009), all agree on a
fundamentally hydrothermal origin of IOCG-clan deposits.
A role for oxide melts in the formation of some IOCG
deposits has been advocated on the basis of evidence for
immiscibility of Fe(Ti)-rich melts in nature (e.g., Clark and
Kontak, 2004; Hurai et al., 1998), from experiments (e.g.,
Charlier and Grove, 2012; Lledo, 2005), from high-T meltlike fluid inclusions (Broman et al., 1999; Velasco and Tornos,
2009), from textural evidence (e.g., Chen et al., 2010a;
Naslund et al., 2002), and from a parallel with FeTiP nelsonite deposits (Kolker, 1982). The immiscibility hypothesis has
been challenged on many geologic, textural, geochemical, and
petrologic grounds (e.g., Bookstrom, 1995; Parak, 1975;
Rhodes et al., 1999; Sillitoe and Burrows, 2002). At best, a
role for immiscibility in most IOCG systems seems problematic given the breadth of geologic evidence that is incompatible
with oxide melts and points to solely hydrothermal origins and
because associated magma types extend far beyond the dry,
reduced, and ferrobasaltic compositions that favor immiscibility (e.g., Charlier and Grove, 2012).
Most discussion centers on sources of fluids and metals,
depositional mechanisms, possible linkages among deposit
types (e.g., magnetite vs. hematite, Cu-rich vs. Cu-poor,
IOCG-clan vs. others), and the role of geologic setting. There
is general agreement that IOCG geochemistry reflects the high
solubility of iron (and other metals) in brines at elevated
temperatures and pressures, the consequences of cooling such
a fluid and its interaction along that path with different rock
types and other fluids (e.g., Barton and Johnson, 1996;
Williams et al., 2005). Saline, low-S fluids can carry many
metals at high temperature and will precipitate Fe and other
nonchalcophile elements on cooling, wall rock reaction or
mixing (e.g., Barton and Johnson, 1996; Chou and Eugster,
1977; Hemley et al., 1992). Consanguineous wall rock reaction
generates ubiquitous NaCaK alteration, and with localized
acid alteration, all are distributed in a regular manner in space

535

and time (Figure 1). These features parallel the patterns found
in many types of hydrothermal systems (Meyer and Hemley,
1967; Seedorff et al., 2005). The large physical and chemical
gradients in the upper few kilometer of the crust are reflected in
the telescoped nature of mineralization in many areas. Conversely, where deeper (metamorphic) saline fluids are key
actors, alteration patterns may be more subtle given the typically smaller physical gradients in metamorphic terrains.
Low aqueous sulfur contents inhibit sulfide precipitation;
therefore, a second sulfur source seems required in most cases
where chalcophile elements are abundant (Chen, 2010;
Haynes, 2000; Williams et al., 2005). However, in favorable
circumstances, sulfide could be generated by the reduction of
sulfate through the oxidation of magnetite to hematite (Gow
et al., 1994). The scarcity of sulfur precludes sulfate being
either the prime acid source (e.g., H2SO4) or the oxidant
required to make hematite from magnetite or aqueous Fe2
species. Possible acid species include HCl and, most likely,
FeCl2. The latter creates acidity via reactions, such as
3FeCl2 4H2O Fe3O4 6HCl H2. Note that in the absence
of sulfur species, water is likely the only oxidant that is sufficiently abundant to make voluminous hypogene magnetite or
hematite; even then, such a process would likely require selective loss of H2 perhaps via phase separation. Inefficient H2 loss
could lead to reduction of hematite to magnetite as could
diffusive reequilibration with most host rocks that are not
hematite stable. In such a scenario, a separate reduced fluid
pulse is not required.
The sources of metals, fluids, and sulfur remain contentious. Most IOCG systems are associated with magmatism,
and as reviewed earlier, many observations are compatible
with an igneous source though not necessarily a magmatic
source. Magmatic fluids can transport abundant Fe along
with other metals (Clark and Arancibia, 1995; Hemley et al.,
1992; Whitney et al., 1985), yet it is difficult to account for the
mass of Fe based on magmatic fluid contents alone just as it is
difficult to account for the extent of other hydrothermal features (Barton and Johnson, 2000). No viable mechanism exists
to create large volumes of NaCa alteration (Figure 8) from
magmatic fluids (the CO2-unmixing model of Pollard, 2001
fails due to the lack of acid to drive the decarbonation
reaction). Nevertheless, many geochemical and experimental
data are compatible with a possible magmatic-hydrothermal
sources (e.g., Groves et al., 2010; Pollard, 2000, 2006), and
such an origin has been argued for many deposits and regions
(e.g., de Haller and Fontbote, 2009; Marschik and Fontbote,
2001; Perring et al., 2000; Rusk et al., 2010). Magmatic heat is
clearly implicated especially at shallow levels, and magmatic
volatile components or phreatic processes undoubtedly play a
role in generating many IOCG-related breccias (e.g., Oliver
et al., 2008). Ultimately, a magmatic contribution of energy
is certain in most systems, even though indisputable evidence
is lacking for a direct magmatic (melt) source for any of the key
constituents in IOCG mineralization.
Conversely, external sources for fluids and nonmagmatic
(though perhaps igneous rock) sources are evidenced by the
volumes of, and types of, alteration (including the mass of
leached metals), the overall lack of correlation with igneous
compositions, the geochemical evidence for clearly nonigneous components, and a correlation with external sources of

536

Iron Oxide(CuAuREEPAgUCo) Systems

Bn

log (Cl-)

log (Cl-)
Springs
Seawater
Alkaline lakes
Basinal brines

fresh

0.0

1.0

-2
-3.0

(b)

1 Zn

10 0.1

350 C 500 b
pH: Mu-Kf fo2: Mt/Hm

-1.0

log (SS)

-1

Py Mt

Solubilities in ppm
+
+
K /Na = 0.1

LS epithermal

Calculated using data from


Ohmoto et al. (1983)

-2.0

log (SS)

Ambient waters

-1.0

Py

SO4=/H2S

e-S
-F ion
Cu osit
p
de

-2.0

100 Zn

1 Fe

Py Hm

2
Cl
Au

-3.0

seawater

1 Cu

Sl

-4.0

10 000 Zn

solubility shifts
with increasing
SO4=/H2S

S) 2
(H
Au

-6.0
-5.0

Terrestrial
Terrestrial - saline
Marine
Marine - Red Sea
Volcanic gases

100 Fe

tic

Precipitation

(a)

Hm
/Mt

>
Fe

Dg

S
H2

a
gm

Geothermal fluids

-+
4
SO

100 Cu

-1

-4.0

Bn

ma

0
-2.0

Ccp

evaporitic

IOCG

log (Cl-)

0.0

SS = 0.001

10 000 Fe

10 000 Cu

SW

Modified from Davidson and Large (1994)

200

0.0

(c)

300

-2

400

T (C)

Figure 12 Diagrams illustrating the spectrum of natural fluids in terms of chloride and sulfur contents and implications for metal transport
and deposition in the terrestrial hydrothermal environment. Total Cl versus total S in natural fluids (a) and with mineral stability and solubility in (b).
(Adapted from Barton MD and Johnson DA (1996) Evaporitic source model for igneous-related Fe oxide(REECuAuU) mineralization. Geology 24:
259262 and Barton MD and Johnson DA (2000) Alternative brine sources for Fe-oxide (CuAu) systems: Implications for hydrothermal alteration
and metals. In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and Related Deposits: A Global Perspective, pp. 4360. Glenside, SA: Australian
Mineral Foundation.) (c) Transport of Au and Cu as a function of T and Cl content for contrasting fluid types illustrating parallels with other
ore-forming settings, such as high and low sulfidation epithermal deposits. See text for discussion.

salinity (Barton and Johnson, 1996, 2000; Haynes, 2000).


In particular, the volumes and types of alkali-rich alteration
(Figure 8) exceed in amount and differ in kind from what can
be produced from hydrous magmas (Barton and Johnson,
2000; Johnson, 2000). Magmatic aqueous fluids might contribute in some cases, but highly saline metamorphic, basinal, and
surface fluids (Figure 12(a)) are more widely available and are
consistent with known sources and settings (e.g., Figure 5).
Such fluids have high Cl/S because of the relatively low solubility of anhydrite (Barton and Johnson, 1996). Moreover, evaporitic or metamorphic saline fluids have the appropriate
compositions to transport large quantities of iron and other
metals in a sulfur-deficient fluid (Figure 12(b)) but only a
limited ability to precipitate chalcophile elements on cooling
this is the central feature of the spectrum of IOCG-clan deposits
thus the importance of having additional sulfur available for
Cu  Au enrichment. Availability of a shallow sulfur source will
reflect the setting; therefore, it stands to reason that different
regions may have widely different Cu endowments in otherwise similar Fe-rich hydrothermal systems. Enrichment in U
and Cu may also require a second near-surface source (Haynes
et al., 1995) that could either be contemporaneous or related
to a protracted history (Skirrow, 2010).

13.20.6.2 IOCG Systems and the Terrestrial Hydrothermal


Environment
In conclusion, these results reviewed here indicate that most
IOCG systems represent a high-salinity end-member of terrestrial hydrothermal systems ore-forming environments in
which surface or basinal fluids play a dominant, but not necessarily an exclusive, role. Low sulfur contents and high salinities are expressed in the oxide-dominated mineralogy, the
paucity of sulfides (with the most chalcophile, Cu, being predominant), and the voluminous system-scale alkali exchange
alteration together with the relative lack of acid alteration.
Variations on this theme reflect a diversity of fluid and solute
sources, expressed as differences in element enrichments, in intensity of alteration, and in some cases, in the abundance of key
components, such as the Fe oxides themselves. In spite of the

diversity (a common feature in many families of ore-forming


systems), the unifying themes tie these together in a way that
is comparable to other major groups of ore deposits.
Based on the evidence presented here, most IOCG systems
comprise the higher temperature analogues to various S-poor
basin-related lower temperature environments (e.g., stratabound Cu, PbZn(Ag)) and, as such, might be considered
igneous-powered red-bed copper systems (see Chapters 13.10
and 14.6). Ultimately, IOCG systems reflect a broad range of
fluidrock interactions starting with saline, broadly surfacederived terrestrial fluids but with likely metamorphic and
possible magmatic variants. As such, they are high-salinity
parallels to lower salinity terrestrial hydrothermal systems
Au(Ag)-bearing, quartz-rich, but base-metal-poor deposits
that form from an analogous array of lower salinity terrestrial
fluids and which have their own metamorphic (orogenic Au)
and magmatic (intrusion-related Au) analogues (Figure 12(c)).

Acknowledgments
Many students and postdocs have contributed to University of
Arizona IOCG studies, in particular David Johnson, Douglas
Kreiner, James Girardi, Frank Mazdab, and Eric Jensen. Financial support has come recently from the USGS MRERP
(08HQGR0060), the US National Science Foundation
(EAR08-38157, EAR98-15032), and the multiple companies
and Science Foundation Arizona who support this work
through the IMR. I thank Mike Porter for sharing selected preprints from his recent volumes. Thoughtful reviews by Steve
Scott, Roger Skirrow, Brian Rusk, and Hamid Mumin were
much appreciated.

References
Alt JC (1999) Hydrothermal alteration and mineralization of oceanic crust: Mineralogy,
geochemistry, and processes. Reviews in Economic Geology 8: 133155.
Andersson UB, Harlov DE, Forster HJ, Nystrom JO, Dulski P, and Broman C (2002) On
the origin of monazite inclusions in apatite from the Kiruna iron ores. GFF 124(4):
232233.

Iron Oxide(CuAuREEPAgUCo) Systems

Baker T (1998) Alteration, mineralization and fluid evolution at the Eloise CuAu
deposit, Cloncurry district, NW Queensland. Economic Geology 93: 12131236.
Baker T, Mustard R, Fu B, et al. (2008) Mixed messages in iron-oxidecoppergold
systems of the Cloncurry district, Australia: Insights from PIXE analysis of halogens
and copper in fluid inclusions. Mineralium Deposita 43: 599608.
Barton MD, Dilles JH, Girardi JD, et al. (2011a) Jurassic igneous-related metallogeny of
southwestern North America. In: Steinger RL and Pennell B (eds.) Great Basin
Evolution and Metallogeny, pp. 373396. Reno, NV: Geological Society of Nevada.
Barton MD, Jensen EP, and Ducea M (2005) Fluid sources for IOCG (Candelaria, Punta
del Cobre) and porphyry Cu-style mineralization, Copiapo batholith, Chile:
Geologic and Sr isotopic constraints. Geological Society of America Abstracts with
Programs 37(7): 316.
Barton MD and Johnson DA (1996) Evaporitic source model for igneous-related Fe
oxide(REECuAuU) mineralization. Geology 24: 259262.
Barton MD and Johnson DA (2000) Alternative brine sources for Fe-oxide (CuAu)
systems: Implications for hydrothermal alteration and metals. In: Porter TM (ed.)
Hydrothermal Iron Oxide CopperGold and Related Deposits: A Global Perspective,
pp. 4360. Glenside, SA: Australian Mineral Foundation.
Barton MD and Johnson DA (2004) Footprints of Fe-oxide (CuAu) Systems: SEG
2004 Predictive Mineral Discovery Under Cover. Centre for Global Metallogeny,
University of Western Australia, Special Publication 33: pp. 112116.
Barton MD, Kreiner DC, Jensen EP, and Girardi JD (2011b) Superimposed
hydrothermal systems and related IOCG and porphyry mineralization near Copiapo,
Chile. In: Proceedings of the 11th Biennial SGA Meeting, Society for Geology
Applied to Ore Deposits, Antofagasta, Chile, pp. 521523.
Bastrakov EN, Skirrow RG, and Davidson GJ (2007) Fluid evolution and origins of iron
oxide CuAu prospects in the Olympic Dam District, Gawler craton South Australia.
Economic Geology 102: 14151440.
Benavides J, Kyser TK, Clark AH, et al. (2007) The Mantoverde iron oxidecoppergold
district, III Region, Chile: The role of regionally-derived, non-magmatic fluid
contributions to chalcopyrite mineralization. Economic Geology 102: 415440.
BHP-Billiton Annual Report (2009) http://www.bhpbilliton.com/home/investors/
reports/Documents/annualReport2009.pdf (accessed December 1, 2011).
Billstrom K, Eilu P, Martinsson O, et al. (2010) IOCG and related mineral deposits of the
Northern Fennoscandian Shield. In: Porter TM (ed.) Hydrothermal Iron Oxide
CopperGold and Related Deposits: A Global Perspective, vol. 4, pp. 381414.
Adelaide: PGC Publishing.
Bjorlykke A, Cumming GL, and Krstic D (1990) New isotopic data from davidites and
sulfides in the Bidjovagge goldcopper deposit, Finnmark, northern Norway.
Mineralogy and Petrology 43: 121.
Bonatti E, Fisher DE, Joensuu O, Rydell HS, and Beyth M (1972)
Ironmanganesebarium deposit from the northern Afar Rift. Economic Geology
67: 717730.
Bookstrom AA (1995) Magmatic features of iron ores of the Kiruna type in Chile and
Sweden: Ore textures and magnetite geochemistry A discussion. Economic
Geology 90: 469475.
Borrok DM, Kesler SE, Boer RH, and Essene EJ (1998) The Vergenoeg
magnetitefluorite deposit, South Africa: Support for a hydrothermal model for
massive iron oxide deposits. Economic Geology 93: 564586.
Bosch B, Deschamps J, Leleu M, Lopoukhine M, Marce A, and Vilbert C (1977) The
geothermal zone of lake Assal (F. T. A. I.): Geochemical and experimental studies.
Geothermics 5: 165175.
Broman C, Nystrom JO, Henriquez F, and Elfman M (1999) Fluid inclusions in
magnetiteapatite ore from a cooling magmatic system at EI Laco, Chile. GFF
121: 253267.
Carten RB (1986) Sodiumcalcium metasomatism: Chemical, temporal, and spatial
relationships at the Yerington, Nevada, porphyry copper deposit. Economic Geology
81: 14951519.
Changkakoti A, Gray A, Morton RD, and Sarkar SN (1987) The Mosabani copper
deposit, India A preliminary study on the nature and genesis of ore fluids.
Economic Geology 82: 16191625.
Changkakoti A, Morton RD, and Gray J (1986) Hydrothermal environments during the
genesis of silver deposits in the Northwest Territories of Canada: Evidence from
fluid inclusions. Mineralium Deposita 21: 6369.
Charlier B and Grove TL (2012) Experiments on liquid immiscibility along tholeiitic
liquid lines of descent. Contributions to Mineralogy and Petrology 164: 2744.
Chen HY (2010) Mesozoic IOCG mineralization in the central Andes: An updated review.
In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and Related Deposits:
A Global Perspective Advances in the Understanding of IOCG Deposits, vol. 3,
pp. 259272. Adelaide: PGC Publishing.
Chen HY, Clark AH, and Kyser TK (2010a) The Marcona magnetite deposit, Ica, centralsouth Peru: A product of hydrous, iron oxide-rich melt. Economic Geology
105: 14411456.

537

Chen HY, Clark AH, Kyser TK, et al. (2010b) Evolution of the giant Marcona-Mina Justa
iron oxidecoppergold district, south-central Peru. Economic Geology
105: 155185.
Chiaradia M, Banks D, Cliff R, Marschik R, and de Haller A (2006) Origin of fluids in
iron oxidecoppergold deposits: Constraints from d37Cl, 87Sr/86Sr, and Cl/Br.
Mineralium Deposita 41: 565573.
Chou IM and Eugster HP (1977) Solubility of magnetite in supercritical chloride
solutions. American Journal of Science 277: 12961314.
Clark AH and Arancibia ON (1995) Occurrence, paragenesis and implications of
magnetite-rich alteration-mineralization in calc-alkaline porphyry copper deposits.
In: Clark AH (ed.) Proceedings: Giant Ore Deposits II. Giant Ore Deposits
Workshop, 2nd edn, pp. 511581. Kingston, ON: Queens University.
Clark AH and Kontak DJ (2004) FeTiP oxide melts generated through magma mixing
in the Antauta subvolcanic center, Peru: Implications for the origin of nelsonite and
iron oxide-dominated hydrothermal deposits. Economic Geology 99: 377395.
Collins AC (2010) Mineralogy and Geochemistry of Tourmaline in Contrasting
Hydrothermal Systems: Copiapo Area, Northern Chile, 225 p. Unpublished MS
Thesis, University of Arizona.
Cooke DR, Wilson AJ, House MJ, et al. (2007) Alkalic porphyry AuCu and associated
mineral deposits of the Ordovician to Early Silurian Macquarie Arc, New South
Wales. Australian Journal of Earth Sciences 54: 445463.
Corriveau L and Mumin H (eds.) (2010) Exploring for Iron Oxide Copper-Gold (Ag-BiCo-U) Deposits: Examples from Canada and Global Analogues. Short Course Notes,
vol. 20, pp. 185. Canada: Geological Association of Canada.
Cox DP and Singer DA (2007) Descriptive and grade-tonnage models and database for
iron oxide CuAu deposits. USGS Open-File Report 2007-1155, 13 p.
Criss RE and Taylor HP Jr (1986) Meteoric hydrothermal systems. Reviews in
Mineralogy and Geochemistry 16: 373424.
Crocker LT, Martini JE, and Sohnge APG (1988) The fluorspar deposits of the Republic
of South Africa and Bophuthatswana. South Africa Geological Survey Handbook
11: 1172.
Daliran F, Stosch HG, Williams PJ, Jamali H, and Dorri M-B (2010) Lower Cambrian
Iron OxideApatiteREE (U) Deposits of the Bafq District, East-Central Iran. Short
Course Notes, vol. 20, pp. 147159. Canada: Geological Association of Canada.
Daroch G and Barton MD (2011) Hydrothermal alteration and mineralization in Santo
Domingo Sur iron oxide (CuAu) deposit, Atacama Region, Chile. In: Proceedings
of the 11th Biennial SGA Meeting, pp. 488490. Antofagasta, Chile: Society for
Geology Applied to Ore Deposits.
Davidson GJ and Dixon GH (1992) Two sulphur isotope provinces deduced from ores
in the Mount Isa Eastern Succession, Australia. Mineralium Deposita 27: 3041.
Davidson GJ and Large RR (1994) Gold metallogeny and the coppergold association
of tile Australian Proterozoic. Mineralium Deposita 29: 208223.
de Haller A, Corfu F, Fontbote L, et al. (2006) Geology, geochronology, and Hf and Pb
isotope data of the Raul-Condestable iron oxidecoppergold deposit, central coast
of Peru. Economic Geology 101: 281310.
de Haller A and Fontbote L (2009) The Raul-Condestable iron oxide coppergold
deposit, central coast of Peru: Ore and related hydrothermal alteration, sulfur
isotopes, and thermodynamic constraints. Economic Geology 104: 365384.
de Jong G, Rotherham J, Phillips G, and Williams PJ (1998) Mobility of rare-earth
elements and copper during shear-zone-related retrograde metamorphism.
Geologie en Mijnbouw 76: 311319.
de Jong G and Williams PJ (1995) Giant metasomatic systems formed during
exhumation of mid-crustal Proterozoic rocks in the vicinity of the Cloncurry fault,
Northwest Queensland. Australian Journal of Earth Sciences 42: 281290.
Deb M and Sarkar SC (1990) Proterozoic tectonic evolution and metallogenesis in the
Aravalli Delhi orogenic complex, Northwestern India. Precambrian Research
46: 115137.
Dilles JH, Einaudi MT, Proffett JM, and Barton MD (2000) Overview of the Yerington
porphyry copper district: Magmatic to non-magmatic sources of hydrothermal
fluids, their flow paths, alteration affects on rocks, and CuMoFeAu ores. Society
of Economic Geologists Guide Book Series 32: 5566.
Dilles JH, Farmer GL, and Field CW (1995) Sodium-calcium alteration due to nonmagmatic saline fluids in porphyry copper deposits: Results for Yerington, Nevada.
Mineralogical Association of Canada Short Course 23: 309338.
Dilles JH and Proffett JM (1995) Metallogenesis of the Yerington batholith, Nevada.
Arizona Geological Society Digest 20: 306315.
Dow RJ and Hitzman MW (2002) GeoIogy of the Arizaro and Lildero Prospects, Salta
Province, Northwest Argentina: Mid Miocene hydrothermal Fe-Ox coppergold
mineralisation. In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold Deposits:
A Global Perspective, vol. 2, pp. 153161. Adelaide: PGC Publishing.
Dreher AM, Xavier RP, Taylor BE, and Martini S (2008) New geologic, fluid inclusion
and stable isotope studies on the controversial Igarape Bahia CuAu deposit,
Carajas Province, Brazil. Mineralium Deposita 43: 161184.

538

Iron Oxide(CuAuREEPAgUCo) Systems

Duncan R, Stein H, Evans K, et al. (2011) A new geochronological framework for


mineralization and alteration in the Selwyn Mount Dore corridor, Eastern Fold Belt,
Mount Isa Inlier, Australia: Genetic implications for iron oxide coppergold
deposits. Economic Geology 106: 169192.
Dupuis C and Beaudoin G (2011) Discriminant diagrams for iron oxide trace element
fingerprinting of mineral deposit types. Mineralium Deposita 46: 319335.
Einaudi MT, Meinert LD, and Newberry RJ (1981) Skarn deposits. Economic Geology
75th Anniversary Volume, 317391.
Einaudi MT and Oreskes N (1990) Progress towards an occurrence model for
Proterozoic iron-oxide (Cu, U, REE, Au) deposits A comparison between the ore
provinces of South Australia and SE Missouri. In: Pratt WP and Sims PK (eds.) The
Mid-continent; Permissive Terrain for an Olympic Dam Deposit, U.S. Geological
Survey Bulletin, vol. 1932, pp. 5869.
Ennis DJ, Dunbar NW, Campbell AR, and Chapin CE (2000) The effects of Kmetasomatism on the mineralogy and geochemistry of silicic ignimbrites near
Socorro, New Mexico. Chemical Geology 167: 285312.
Ettner DC, Bjorlykke A, and Andersen T (1993) Fluid evolution and AuCu genesis
along a shear zone: A regional fluid inclusion study of shear zone-hosted alteration
and gold and copper mineralization in the Kautokeino greenstone belt, Finnmark,
Norway. Journal of Geochemical Exploration 49: 233267.
Fisher LA and Kendrick MA (2008) Metamorphic fluid origins in the Osborne Fe oxide
CuAu deposit, Australia: Evidence from noble gases and halogens. Mineralium
Deposita 43: 483497.
Fournier RO (1987) Conceptual models of brine evolution. U.S. Geological Survey
Professional Paper 1350: 14871506.
Franklin JM, Gibson HL, Jonasson IR, and Galley AG (2005) Volcanogenic massive
sulfide deposits. Economic Geology 100th Anniversary Volume, 523560.
Frietsch R, Tuisku P, Martinsson O, and Perdahl JA (1997) CuAu and Fe ore deposits
associated with NaCl metasomatism in early Proterozoic rocks of northern
Fennoscandia: A new metallogenic province. Ore Geology Reviews 12: 134.
Galarza MA, Macambira MJB, and Villas RN (2008) Dating and isotopic characteristics
(Pb and S) of the Fe oxideCuAuUREE Igarape Bahia ore deposit, Carajas
mineral province, Para state, Brazil. Journal of South American Earth Sciences
25: 377397.
Gandhi SS (1988) Volcanoplutonic setting of UCu bearing magnetite veins of FAB
claims, southern Great Bear Magmatic Zone, Northwest Territories. Geological
Survey of Canada Paper 88-IC: 177187.
Gleason JD, Marikos MA, Barton MD, and Johnson DA (2000) Neodymium isotopic
study of rare earth element sources and mobility in hydrothermal Fe oxide
(FePREE) systems. Geochimica et Cosmochimica Acta 64: 10591068.
Gleeson SA and Smith MP (2009) The sources and evolution of mineralising fluids in
iron oxidecoppergold systems, Norrbotten, Sweden: Constraints from Br/CI
ratios and stable CI isotopes of fluid inclusion leachates. Geochimica et
Cosmochimica Acta 73: 56585672.
Golani PR, Pandit MK, Sial AN, et al. (2002) BNa rich Palaeoproterozoic Aravalli
metasediments of evaporitic association, NW India: A new repository of gold
mineralization. Precambrian Research 116: 183198.
Gow PA, Wall VJ, Oliver NHS, and Valenta RK (1994) Proterozoic iron oxide
(CuUAuREE) deposits: Further evidence of hydrothermal origin. Geology
22: 633636.
Groves DIP, Bierlein FP, Meinert LD, and Hitzman MW (2010) Iron oxide coppergold
(IOCG) deposits through Earth history; implications for origin, lithospheric setting,
and distinction from other epigenetic iron oxide deposits. Economic Geology
105: 641654.
Hauck SA (1990) Petrogenesis and tectonic setting of middle Proterozoic iron oxiderich ore deposits An ore deposit model for Olympic Dam-type mineralization.
USGS Bulletin 1932: 439.
Hawkins T, Herrington R, Smith M, Maslenikov V, and Boyce A (2010) The iron skarns
of the Turgai Belt, Northwestern Kazakhstan. In: Porter TM (ed.) Hydrothermal Iron
Oxide CopperGold and Related Deposits: A Global Perspective Advances in the
Understanding of IOCG Deposits, vol. 4, pp. 461474. Adelaide: PGC Publishing.
Haynes D (2000) Iron oxide copper(gold) deposits: Their position in the ore deposit
spectrum and modes of origin. In: Porter TM (ed.) Hydrothermal Iron Oxide
CopperGold and Related Deposits: A Global Perspective, vol. 1, pp. 7190.
Adelaide: PGC Publishing.
Haynes DW, Cross KC, Bills RT, and Reed MH (1995) Olympic Dam ore genesis:
A fluid-mixing model. Economic Geology 90: 281307.
Hayward N and Skirrow RG (2010) Geodynamic setting and controls on iron oxide
CuAu (U) Ore in the Gawler Craton, South Australia. In: Porter TM (ed.)
Hydrothermal Iron Oxide CopperGold and Related Deposits: A Global Perspective
Advances in the Understanding of IOCG Deposits, vol. 3, pp. 119146. Adelaide:
PGC Publishing.

Helvaci C and Griffin WL (1983) Metamorphic feldspathization of metavolcanics and


granitoids, Avnik area, Turkey. Contributions to Mineralogy and Petrology
83: 309319.
Hemley JJ, Cygan GL, Fein JB, Robinson GR, and DAngelo WM (1992) Hydrothermal
ore-forming processes in the light of studies of rock-buffered systems: 1.
Ironcopperzinclead sulfide solubility relations. Economic Geology 87: 122.
Hildebrand RS (1986) Kiruna-type deposits: Their origin and relationship to
intermediate subvolcanic plutons in the Great Bear Magmatic Zone, Northwest
Canada. Economic Geology 81: 640659.
Hitzman MC (2000) Iron oxideCuAu deposits: What, where, when, and why?
In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and Related Deposits:
A Global Perspective, vol. 1, pp. 925. Adelaide: PGC Publishing.
Hitzman MW, Oreskes N, and Einaudi MT (1992) Geological characteristics and tectonic
setting of Proterozoic iron oxide (CuUAuREE) deposits. Precambrian Research
58: 241287.
Hitzman MW and Valenta RK (2005) Uranium in iron oxide coppergold (IOCG)
systems. Economic Geology 100: 16571661.
Hollocher K, Spencer J, and Ruiz J (1994) Composition changes in an ash-flow cooling
unit during K-metasomatism, west-central Arizona. Economic Geology
89: 877888.
Hou T, Zhang Z, Encarnacion J, et al. (2010) Geochemistry of Late Mesozoic dioritic
porphyries associated with Kiruna-style and stratabound carbonate-hosted Zhonggu
iron ores, MiddleLower Yangtze Valley, Eastern China: Constraints on
petrogenesis and iron sources. Lithos 119: 330344.
Hunt JA, Baker T, Cleverly J, Davidson GJ, Fallick AE, and Thorkelson DJ (2011) Fluid
inclusion and stable isotope constraints on the origin of Wernecke Breccia and
associated iron oxidecoppergold mineralization, Yukon. Canadian Journal of
Earth Sciences 48: 14251445.
Hunt JA, Baker T, and Thorkelson DJ (2007) A review of iron oxide coppergold
deposits, with focus on the Wernecke Breccias, Yukon, Canada, as an example of a
non-magmatic end member and implications for IOCG genesis and classification.
Exploration and Mining Geology 16: 209232.
Hurai V, Simon K, Wiechert U, et al. (1998) Immiscible separation of metalliferous
Fe/Ti-oxide melts from fractionating alkali basalt: PTfO2 conditions and two
liquid elemental partitioning. Contributions to Mineralogy and Petrology
133: 1229.
Huston DL, Bolger C, and Cozens G (1993) A comparison of mineral deposits at the
Gecko and White Devil deposits: Implications for ore genesis in the Tennant Creek
district, Northern Territory, Australia. Economic Geology 88: 11981225.
Jami M, Dunlop AC, and Cohen DR (2007) Fluid inclusion and stable isotope study of
the Esfordi apatitemagnetite deposit, central Iran. Economic Geology
102: 11111128.
Jensen EP and Barton MD (2000) Gold deposits related to alkaline magmatism. Reviews
in Economic Geology 13: 279314.
Johnson DA (2000) Studies of Iron-Oxide (CuREEAuCoAgNiU) Mineralization
and Associated Sodic Alteration in the Great Basin, p 277. Unpublished PhD
Dissertation, University of Arizona.
Johnson DA and Barton MD (2000) Time-space development of an external
brine-dominated, igneous-driven hydrothermal system; Humboldt mafic complex,
western Nevada. Society of Economic Geologists Guidebook 32: 127143.
Johnson DA and Barton MD (in press) Iron-oxide  CuAuREE deposits and
their by-product potential. US Geological Survey Scientific Investigations Report
2012-x.
Johnson JP and McCulloch MT (1995) Sources of mineralizing fluids for the Olympic
Dam deposit, (South Australia): SmNd isotopic constraints. Chemical Geology
121: 177199.
Kasemann SA, Meixner A, Erzinger J, et al. (2004) Boron isotope composition of
geothermal fluids and borate minerals from salar deposits (central Andes/NW
Argentina). Journal of South American Earth Sciences 16: 685697.
Kendrick MA, Baker T, Fu B, Phillips D, and Williams PJ (2008a) Noble gas and
halogen constraints on regionally extensive mid-crustal NaCa metasomatism, the
Proterozoic Eastern Mount Isa Block, Australia. Precambrian Research
163: 131150.
Kendrick MA, Honda M, Gillen D, Baker T, and Phillips D (2008b) New constraints on
regional brecciation on the Wernecke Mountains, Canada, from He, Ne, Ar, Kr, Xe,
Cl, Br and I. Chemical Geology 255: 3346.
Kendrick MA, Honda M, Oliver NHS, and Phillips D (2011) The noble gas systematics of
late-orogenic H2OCO2 fluids, Mt Isa, Australia. Geochimica et Cosmochimica Acta
75: 14281450.
Kerrich R and Rehrig W (1987) Fluid motion associated with the Tertiary mylonitization
and detachment faulting: 18O/16O evidence from the Picacho metamorphic core
complex, Arizona. Geology 15: 5862.

Iron Oxide(CuAuREEPAgUCo) Systems

Kirkham RV (1989) Distribution, settings, and genesis of sediment-hosted stratiform


copper deposits. Geological Association of Canada Special Paper 36: 338.
Kisvarsanyi EB (1972) Petrochemistry of a Precambrian igneous province, St. Francois
Mountains, Missouri. Report of Investigations 51. Missouri Geological Survey, 103 p.
Knight J, Joy S, Cameron J, et al. (2002) The Khetri copper belt Rajasthan: Iron oxide
coppergold terrain in the Proterozoic of NW India. In: Porter TM (ed.)
Hydrothermal Iron Oxide CopperGold and Related Deposits: A Global Perspective,
vol. 2, pp. 321341. Adelaide: PGC Publishing.
Kolker A (1982) Mineralogy and geochemistry of FeTi oxide and apatite (nelsonite)
deposits and evolution of the liquid immiscibility hypothesis. Economic Geology
77: 11461158.
Kontak DJ, Archibald DA, Creaser RA, and Heaman LM (2008) Dating hydrothermal
alteration and IOCG mineralization along a terrane-bounding fault zone: The Copper
Lake deposit, Nova Scotia. Atlantic Geology 44: 146166.
Kreiner DC (2011) Epithermal Style Iron Oxide(CuAu) (IOCG) Vein Systems and
Related Alteration, p 659. Unpublished PhD Dissertation, University of Arizona.
Kreiner DC and Barton MD (2009) Hydrothermal alteration and mineralization zoning in
iron oxide (CuAu) vein deposits near Copiapo, Chile. In: Williams PJ (ed.) Smart
Science for Exploration and Mining, pp. 635637. Townsville: James Cook University.
Kuscu Y, Yilmazer E, Gulec N, et al. (2011) UPb and 40Ar39Ar geochronology and
isotopic constraints on the genesis of coppergold-bearing iron oxide deposits in
the Hasancelebi District, eastern Turkey. Economic Geology 106: 261288.
Lambert LB, Drexal IF, Donnelly TH, and Knutson I (1982) Origin of breccias in the
Mount Painter area, South Australia. Journal of the Geological Society Australia
29: 115125.
Lang JR, Stanley CR, and Thompson JFH (1995) Porphyry coppergold deposits
related to alkalic igneous rock in the TriassicJurassic arc terranes of British
Columbia. Arizona Geological Society Digest 20: 219236.
Lapham DDM (1968) Triassic magnetite and diabase at Cornwall, Pennsylvania.
In: Ridge JD (ed.) Ore Deposits of the United States 19331967, pp. 7294.
New York: American Institute of Mining Metallurgical and Petroleum Engineers.
Leach DL, Bradley DC, Huston D, et al. (2010) Sediment-hosted leadzinc deposits in
Earth history. Economic Geology 105: 593626.
Lindgren W (1913) Mineral Deposits, 883 p. New York: McGraw-Hill.
Lledo LH (2005) Experimental Studies on the Origin of Iron Deposits; and
Mineralization of Sierra La Bandera, Chile, 200 p. Unpublished PhD Thesis, State
University of New York, Binghamton.
Lund K, Tysdal RG, Evans KV, Kunk MJ, and Pillers RM (2011) Structural controls
and evolution of gold-, silver-, and REE-bearing coppercobalt ore deposits,
Blackbird District, east-central Idaho: Epigenetic origins. Economic Geology
106: 585618.
Maas R, McCulloch MT, and Campbell IH (1987) SmNd isotope systematics in
uranium-rare earth element mineralization at the Mary Kathleen uranium Mine,
Queensland. Economic Geology 82: 18051826.
Maksaev V and Zentilli M (2002) Chilean strata-bound CuAg deposits: An overview.
In: Porter TM (ed.) Hydrothermal Iron-Oxide CopperGold and Related Deposits:
A Global Perspective, vol. 2, pp. 185205. Adelaide: PGC Publishing.
Mark G, Foster DRW, Pollard PJ, et al. (2004) Stable isotope evidence for magmatic
fluid input during large-scale NaCa alteration in the Cloncurry Fe oxide CuAu
district, NW Queensland, Australia. Terra Nova 16: 5461.
Mark G, Oliver NHS, and Carew MJ (2006a) Insights into the genesis and diversity of
epigenetic CuAu mineralisation in the Cloncurry district, Mt. Isa Inlier, Northwest
Queensland. Australian Journal of Earth Sciences 53: 109124.
Mark G, Oliver NHS, and Williams PJ (2006b) Mineralogical and chemical evolution of
the Ernest Henry Fe oxideCuAu ore system, Cloncurry district, Northwest
Queensland, Australia. Mineralium Deposita 40: 769801.
Marschik R, Chiaradia M, and Fontbote L (2003) Implications of Pb isotope signatures
of rocks and iron oxide CuAu ores in the Candelaria-Punta del Cobre district,
Chile. Mineralium Deposita 38: 900912.
Marschik R and Fontbote L (2001) The Candelaria-Punta del Cobre iron oxide
CuAuZnAg deposits, Chile. Economic Geology 96: 17991826.
Marschik R, Rieger AA, Kendrick MA, and Diaz M (2011) Noble gas and halogen
constraints on the hypogene IOCG mineralization at Mantoverde, Chile.
In: Proceedings of the 11th SGA Biennial Meeting, pp. 452454. Antofagasta, Chile:
Society for Geology Applied to Ore Deposits.
Mathur RD, Marschik R, Ruiz J, et al. (2002) Age of mineralization of the Candelaria Fe
oxide CuAu deposit and the origin of the Chilean iron belt, based on ReOs
isotopes. Economic Geology 97: 5971.
Mazdab FK (2001) The Distribution of Trace Elements in Iron Sulfides and Associated
Chlorine-Bearing Silicates, 795 p. Unpublished PhD Thesis, University of Arizona.
Mazdab FK (2003) The diversity and occurrence of potassium-dominant amphiboles.
The Canadian Mineralogist 41: 13291344.

539

Mazdab FK, Johnson DA, and Barton MD (2008) Trace element characteristics of
hydrothermal titanite from iron-oxideCuAu (IOCG) mineralization. Geochimica et
Cosmochimica Acta 72: A609.
McInnes BIA, Keays RR, Lambert DD, Hellstrom J, and Allwood IS (2008) ReOs
geochronology and isotope systematics of the Tanami, Tennant Creek and Olympic
Dam CuAu deposits. Australian Journal of Earth Sciences 55: 967981.
McKibben MA and Hardie LA (1997) Ore-forming brines in active continental rifts.
In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, 3rd edn.,
pp. 875933. New York: Wiley.
McPhie J, Kamenetsky VS, Chambefort I, Ehrig K, and Green N (2011) Origin of the
supergiant Olympic Dam CuUAuAg deposit, South Australia: Was a
sedimentary basin involved? Geology 30: 795798.
Menard JJ (1995) Relationship between altered pyroxene diorite and the magnetite
mineralization in the Chilean Iron Belt, with emphasis on the EI Algarrobo iron
deposits (Atacama region, Chile). Mineralium Deposita 30: 268274.
Meyer C (1988) Ore deposits as guides to geologic history of the Earth. Annual Review
of Earth and Planetary Sciences 16: 147171.
Meyer C and Hemley JJ (1967) Wall rock alteration. In: Barnes HL (ed.) Geochemistry of
Hydrothermal Ore Deposits, pp. 166235. New York: Holt, Rinehart, and Winston.
Meyer C, Wunder B, Meixner A, Romer RL, and Heinrich W (2008) Boron-isotope
fractionation between tourmaline and fluid: An experimental re-investigation.
Contributions to Mineralogy and Petrology 156: 259267.
Monteiro LVS, Xavier RP, de Carvalho ER, et al. (2008a) Spatial and temporal zoning of
hydrothermal alteration and mineralization in the Sossego iron oxidecoppergold
deposit, Carajas mineral province, Brazil: Paragenesis and stable isotope
constraints. Mineralium Deposita 43: 129159.
Monteiro LVS, Xavier RP, Hitzman MW, Juliani C, de Souza Filho CR, and Carvalho E
(2008b) Mineral chemistry of ore and hydrothermal alteration at the Sossego iron
oxidecoppergold deposit, Carajas mineral province, Brazil. Ore Geology Reviews
34: 317336.
Moore JM (2010) Comparative study of the Onganja copper mine, Namibia: A link
between Neoproterozoic mesothermal CuAu mineralization in Namibia and
Zambia. South African Journal of Geology 113: 445460.
Morales-Ruano SM, Both RA, and Golding SD (2002) A fluid inclusion and stable
isotope study of the Moonta coppergold deposits, South Australia: Evidence for
fluid immiscibility in a magmatic hydrothermal system. Chemical Geology
192: 211226.
Mumin AH, Corriveau L, Somarin AK, and Ootes L (2007) Iron oxide coppergold-type
polymetallic mineralization in the Contact Lake Belt, Great Bear Magmatic Zone,
Northwest Territories, Canada. Exploration and Mining Geology 16: 187208.
Mumin AH, Somarin AK, Jones B, Corriveau L, Ootes L, and Camier J (2010) The
IOCG-Porphyry-Epithermal Continuum of Deposits Types in the Great Bear
Magmatic Zone, Northwest Territories, Canada. Short Course Notes, vol. 20,
pp. 5978. Canada: Geological Association of Canada.
Muntean JL and Einaudi MT (2000) Porphyry gold deposits of the Refugio district,
Maricunga belt, northern Chile. Economic Geology 95: 14451472.
Nash JT and Hahn GA (1989) Stratabound CoCu deposits and mafic volcaniclastic
rocks in the Blackbird mining district, Lemhi County, Idaho. Geological Association
of Canada Special Paper 36: 339356.
Naslund HR, Henriquez E, Nystrom JO, Vivallo, and Wand Dobbs FM (2002) Magmatic
iron ores and associated mineralization: Examples from the Chilean High Andes and
Coastal Cordillera. In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and
Related Deposits: A Global Perspective, vol. 2, pp. 207226. Adelaide: PGC
Publishing.
Niiranen T, Manttari I, Poutiainen M, Oliver NHS, and Miller JA (2005) Genesis of
Palaeoproterozoic iron skarns in the Misi region, northern Finland. Mineralium
Deposita 40: 192217.
Niiranen T, Poutiainen M, and Mantuiri I (2007) Geology, geochemistry, fluid inclusion
characteristics, and UPb age studies on iron oxideCuAu deposits in the Kolari
region, northern Finland. Ore Geology Reviews 30: 75105.
Ohmoto H and Rye RO (1979) Isotopes of sulfur and carbon. In: Barnes HL (ed.)
Geochemistry of Hydrothermal Ore Deposits, 2nd edn., pp. 509567. New York:
Wiley.
Oliver NHS, Butera KM, Rubenach MI, et al. (2008) The protracted hydrothermal
evolution of the Mount Isa Eastern Succession: A review and tectonic implications.
Precambrian Research 163: 108130.
Oliver NHS, Cleverley IS, Mark G, et al. (2004) Modeling the role of sodic alteration in
the genesis of iron oxidecoppergold deposits, Eastern Mount Isa block, Australia.
Economic Geology 99: 11451176.
Oreskes N and Einaudi MT (1990) Origin of REE-enriched hematite breccias at the
Olympic Dam CuUAuAg deposit, Roxby Downs, South Australia. Economic
Geology 85: 128.

540

Iron Oxide(CuAuREEPAgUCo) Systems

Oreskes N and Einaudi MT (1992) Origin of hydrothermal fluids at Olympic Dam:


Preliminary results from fluid inclusions and stable isotopes. Economic Geology
87: 6490.
Pal DC, Barton MD, and Sarangi AK (2009) Deciphering a multistage history affecting
UCuFe mineralization in the Singbhum Shear Zone, eastern India, using pyrite
textures and compositions in the Turamdih UCuFe deposit. Mineralium Deposita
44: 6180.
Pal DC, Trumbull RB, and Wiedenbeck M (2010) Chemical and boron isotope
compositions of tourmaline from the Jaduguda UCuFe deposit, Singhbhum
shear zone, India: Implications for the sources and evolution of mineralizing fluids.
Chemical Geology 277: 245260.
Palinkas SS, Spangenberg JE, and Palinkas LA (2009) Organic and inorganic
geochemistry of Ljubija siderite, deposits, NW Bosnia and Herzegovina. Mineralium
Deposita 44: 893913.
Palmer MR, London D, Morgan GB, and Babb HA (1992) Experimental determination of
fractionation of 11B/10B between tourmaline and aqueous vapor: A temperature- and
pressure-dependent isotopic system. Chemical Geology 101: 123129.
Palmer MR and Swihart GH (1996) Boron isotope geochemistry: An overview. Reviews
in Mineralogy and Geochemistry 33: 709744.
Pan Y (2010) Scapolite in skarn deposits: Petrogenetic and geochemical significance.
Mineralogical Association of Canada Short Course Series 26: 169209.
Pan Y and Dong P (1999) The Lower Changjiang (Yangzi/Yangtze River) metallogenic
belt, East Central China: Intrusion- and wallrock-hosted CuFeAu, Mo, Zn, Pb, Ag
deposits. Ore Geology Reviews 15: 177244.
Pan Y and Dong P (2003) Bromine in scapolite-group minerals and sodalite: XRF
microprobe analysis, exchange experiments, and application to skarn deposits. The
Canadian Mineralogist 41: 529540.
Pandey P, Kumar P, and Upadhyay LD (1994) Uranium deposits of TuramdihNandup
area, Singhbhum District, Bihar and their spatial relationship. Exploration and.
Research for Atomic Minerals 7: 113.
Parak T (1975) Kiruna iron ores are not intrusive magmatic ores of the Kiruna type.
Economic Geology 70: 12421258.
Parak T (1991) Volcanic sedimentary rock-related metallogenesis in the
KirunaSkellefte Belt of northern Sweden. Economic Geology Monographs
8: 2050.
Pearce JA, Harris NBW, and Tindle AG (1984) Trace element discrimination diagrams
for the tectonic interpretation of granitic rocks. Journal of Petrology 25: 956983.
Perring CS, Pollard PJ, Dong G, Nunn AJ, and Blake KL (2000) The Lightning Creek Sill
Complex, Cloncurry District, Northwestern Queensland: A source of fluids for the Fe
oxide CuAu mineralization and sodiccalcic alteration. Economic Geology
95: 10671089.
Pollard PJ (2000) Evidence of a magmatic fluid and metal source for Fe-oxide CuAu
mineralisation. In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and
Related Deposits: A Global Perspective, vol. 1, pp. 2746. Adelaide: Australian
Mineral Foundation.
Pollard PJ (2001) Sodic (calcic) alteration in Fe-oxideCuAu districts: An origin via
unmixing of magmatic H2OCO2NaCl  CaCI2KCl fluids. Mineralium Deposita
36: 93100.
Pollard PJ (2006) An intrusion-related origin for CuAu mineralization in iron oxide
coppergold (IOCG) provinces. Mineralium Deposita 41: 179187.
Porter TM (ed.) (2000) In: Hydrothermal Iron Oxide CopperGold and Related Deposits:
A Global Perspective, vol. 1, pp. 349. Adelaide: PGC Publishing.
Porter TM (ed.) (2002) In: Hydrothermal Iron Oxide CopperGold and Related Deposits:
A Global Perspective, vol. 2, pp. 378. Adelaide: PGC Publishing.
Porter TM (2010a) Current understanding of iron oxide associated-alkali altered
mineralised systems: Part II, a review. In: Porter TM (ed.) Hydrothermal Iron Oxide
CopperGold and Related Deposits: A Global Perspective, vol. 3, pp. 33106.
Adelaide: PGC Publishing.
Porter TM (ed.) (2010b) Hydrothermal Iron Oxide CopperGold and Related Deposits: A
Global Perspective Advances in the Understanding of IOCG Deposits, vols 3 and
4, 600 p. Adelaide: PGC Publishing.
Proffett JM (2003) Geology of the Bajo de la Alumbrera porphyry copper-gold deposit,
Argentina. Economic Geology 98(8): 15351574.
Ray GE and Dick LA (2002) The Productora prospect in north-central Chile: An example
of an intrusion-related Candelaria type FeCuAu hydrothermal system.
In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and Related Deposits: A
Global Perspective, vol. 2, pp. 131151. Adelaide: PGC Publishing.
Reardon NC (1992) Magmatic-Hydrothermal Systems and Associated Magnetite
ApatiteActinolite Deposits, Echo Bay, Northwest Territories, 154 p. Unpublished
MSc Thesis, University of Ottawa.
Requia K, Stein H, Fontbote L, and Chiaradia M (2003) ReOs and PbPb
geochronology of the Archean Salobo iron oxide coppergold deposit, Carajas
mineral province, northern Brazil. Mineralium Deposita 38: 727738.

Rhodes AL, Oreskes N, and Sheets S (1999) Geology and rare earth element
geochemistry of magnetic deposits at El Laco, Chile. In: Skinner BJ (ed.) Geology
and Ore Deposits of the Central Andes, vol. 7, pp. 299332. SEG Special
Publication.
Rieger A, Marschik R, and Daz M (2012) The evolution of the hydrothermal IOCG
system in the Mantoverde district, northern Chile: New evidence from
microthermometry and stable isotope geochemistry. Mineralium Deposita
47: 359369.
Rieger AA, Marschik R, Diaz M, et al. (2010) The hypogene iron oxide coppergold
mineralization in the Mantoverde District, Northern Chile. Economic Geology
105: 12711299.
Ripley EM and Ohmoto H (1977) Mineralogic, sulphur isotope, and fluid inclusion
studies of the stratabound copper deposits at the Raul Mine, Peru. Economic
Geology 72: 10171041.
Robinson GR Jr (1988) Base and precious metals associated with diabase in the
Newark, Gettysburg, and Culpeper basins of the eastern United States A review. U.
S. Geological Survey Professional Paper 1776: 303319.
Roddy MS, Reynolds SJ, Smith BM, and Ruiz J (1988) K metasomatism and
detachment-related mineralization, Harcuvar Mountains, Arizona. Geological
Society of America Bulletin 100: 16271639.
Rose AW, Herrick DC, and Deines P (1985) An oxygen and sulfur isotope study of
skarn-type magnetite deposits of the Cornwall type, southeastern Pennsylvania.
Economic Geology 80: 418443.
Rusk B, Oliver NHS, Brown A, Lilly R, and Jungmann D (2009) Barren magnetite
breccias in the Cloncurry region, Australia; comparisons to IOCG deposits.
In: Proceedings of the 10th Bien. SGA Meeting, pp. 656658. Townsville, Australia:
Society for Geology Applied to Ore Deposits.
Rusk BG, Oliver NHS, Cleverley JS, et al. (2010) Physical and chemical
characteristics of the Ernest Henry iron oxide copper gold deposit, Australia;
Implications for IOCG genesis. In: Porter TM (ed.) Hydrothermal Iron Oxide
CopperGold and Related Deposits: A Global Perspective, vol. 3, pp. 201218.
Adelaide: PGC Publishing.
Ryan PJ, Lawrence AL, Jenkins RA, et al. (1995) The Candelaria coppergold deposit,
Chile. Arizona Geological Society Digest 20: 625645.
Scotese CR and Denham DM (1988) Terra Mobilus: Austin, Texas, Earth in Motion
Technologies.
Seedorff E, Dilles JH, Proffett JM Jr, et al. (2005) Porphyry-related deposits:
Characteristics and origin of hypogene features. Economic Geology 100th
Anniversary Volume, 251298.
Seeger CM (2000) Southeast Missouri iron metallogenic province: Characteristics and
general chemistry. In: Porter TM (ed.) Hydrothermal Iron-Oxide CopperGold and
Related Deposits: A Global Perspective, vol. I, pp. 237248. Adelaide: PGC
Publishing.
Serdyuchenko DP (1975) Some Precambrian scapolite-bearing rocks evolved from
evaporites. Lithos 8: 17.
Sheppard SMF (1986) Characterization and isotopic variations in natural waters.
Reviews in Mineralogy and Geochemistry 16: 65184.
Sillitoe RH (2003) Iron oxidecoppergold deposits: An Andean view. Mineralium
Deposita 38: 787812.
Sillitoe RH and Burrows DR (2002) New field evidence bearing on the origin of the EI
Laco magnetite deposit, northern Chile. Economic Geology 97: 11011109.
Skirrow RG (2010) Hematite-group IOCG  U ore systems: Tectonic settings,
hydrothermal characteristics, and CuAu and U mineralizing processes.
In: Corriveau L and Mumin H (eds.) Exploring for Iron Oxide CopperGold
Deposits: Canada and Global Analogues. Short Course Notes, vol. 20, pp. 3959.
Canada: Geological Association of Canada.
Skirrow RG (2011) Controls on uranium in iron oxide coppergold systems: Insights
from Proterozoic and Paleozoic deposits in southern Australia. In: Proceedings of
the 11th Biennial SGA Conference, pp. 482484.
Skirrow RG, Bastrakov EN, Barovich K, et al. (2007) Timing of iron oxide CuAuU
hydrothermal activity and Nd isotopic constraints on metal sources in the Gawler
Craton, South Australia. Economic Geology 102: 14411470.
Skirrow RG and Davidson GJ (2007) A special issue devoted to Proterozoic iron oxide
CuAu(U) and gold mineral systems of the Gawler Craton: Preface. Economic
Geology 102: 13731375.
Skirrow RG and Walshe JL (2002) Reduced and oxidized AuCuBi iron oxide deposits
of the Tennant Creek inlier, Australia: An integrated geologic and chemical model.
Economic Geology 97: 11671202.
Slack JF, Johnson CA, Lund KI, Schulz KJ, and Causey JD (2011) A new model for
CoCuAu deposits in metasedimentary rocks: An IOCG connection?
In: Proceedings of the 11th Biennial Meeting of the Society for Geology Applied
to Ore Deposits, pp. 491493. Antofagasta, Chile: Society for Geology Applied to
Ore Deposits.

Iron Oxide(CuAuREEPAgUCo) Systems

Solomon GC and Taylor HP Jr (1993) Oxygen isotope studies of Jurassic fossil


hydrothermal systems, Mojave Desert, southeastern California. Geochemical
Society Special Publication 3: 449462.
Soloviev SG (2010a) Iron oxide coppergold and related mineralisation of the Siberian
Craton, Russia: I Iron oxide deposits in the Angara and Ilim river basins, southcentral Siberia. In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and
Related Deposits: A Global Perspective Advances in the Understanding of IOCG
Deposits, vol. 4, pp. 495514. Adelaide: PGC Publishing.
Soloviev SG (2010b) Iron oxide coppergold and related mineralisation of the Siberian
Craton, Russia: 2 Palaeoproterozoic and Mesozoic assemblages of iron oxide, Cu,
Au and U deposits of the Aldan Shield, Southeastern Siberia. In: Porter TM (ed.)
Hydrothermal Iron Oxide CopperGold and Related Deposits: A Global Perspective
Advances in the Understanding of IOCG Deposits, vol. 4, pp. 515534. Adelaide:
PGC Publishing.
Tobey EF (1976) Geology of the Bull Valley IntrusiveExtrusive Complex and Genesis of
the Associated Iron Deposits, 244 p. Unpublished PhD Dissertation, University of
Oregon.
Torab FM and Lehmann B (2007) Magnetiteapatite deposits of the Bafq district,
Central Iran: Monazite geochronology and ore formation. In: Proceedings of the 9th
Biennial SGA Conference, pp. 439442.
Tornos F, Velasco F, Barra F, and Morata D (2010) The Tropezon CuMoAu deposit,
Northern Chile: The missing link between IOCG and porphyry copper systems?
Mineralium Deposita 45: 313321.
Tornos F, Wiedenbeck M, and Velasco F (2012) The boron isotope geochemistry of
tourmaline-rich alteration in the IOCG systems of northern Chile: Implications for a
magmatic-hydrothermal origin. Mineralium Deposita 47: 483499.
Torresi I, Xavier RP, Bortholoto DFA, and Monteiro LVS (2012) Hydrothermal alteration,
fluid inclusions and stable isotope systematics of the Alvo 118 iron
oxidecoppergold deposit, Carajas mineral province (Brazil). Implications for ore
genesis. Mineralium Deposita 47: 299323.
Treloar PJ and Colley H (1996) Variation in F and Cl content in apatites from
magnetiteapatite ores in northern Chile, and their ore-genetic implications.
Mineralogical Magazine 60: 285301.
Vakhrushev VA, Ripp GS, and Kaviladze MS (1981) Isotope composition of sulfur from
iron ore deposits of east Siberia. Geologiya i Geofizika 22: 7481.
Vapnik Y, Bushmin S, Chattopadhyay A, and Dolivo-Dobrovolsky D (2007) Fluid
inclusion and mineralogical study of vein-type apatite ores in shear zones from the
Singhbhum metallogenetic province, West Bengal, India. Ore Geology Reviews
32: 412430.
Velasco F and Tornos F (2009) Pegmatite-like magnetiteapatite deposits of northern
Chile: A place in the evolution of immiscible iron oxide melts? In: Proceedings of
the 10th Biennial SGA Meeting, pp. 665667. Townsville, Australia: Society for
Geology Applied to Ore Deposits.
Wang S and Williams P (2001) Geochemistry and origin of Proterozoic skarns at the
Mount Elliott CuAu(CoNi) deposit, Cloncurry District, NW Queensland,
Australia. Mineralium Deposita 36: 109124.
Wanhainen C and Martinsson O (2010) The hybrid character of the Aitik deposit,
Norrbotten, Sweden: A porphyry CuAuAg(Mo) system overprinted by
iron-oxide CuAu hydrothermal fluids. In: Porter TM (ed.) Hydrothermal Iron Oxide
CopperGold and Related Deposits: A Global Perspective Advances in the
Understanding of IOCG Deposits, vol. 4, pp. 415426. Adelaide: PGC Publishing.
Wenner DB and Taylor HP Jr (1987) Oxygen and hydrogen isotope studies of a
Precambrian graniterhyolite terrane, St. Francois Mountains, southeastern
Missouri. Geological Society of America Bulletin 87: 15871598.
Whitney JA, Hemley JJ, and Simon FO (1985) The concentration of iron in chloride
solutions equilibrated with synthetic granite compositions: The sulfur-free system.
Economic Geology 80: 444460.

541

Wilkins JJ, Beane RE, and Heidrick TL (1986) Mineralization related to detachment
faults: A model. Arizona Geological Society Digest 16: 108117.
Williams PJ (1994a) Diverse nature of hematization (red rock alteration) in the
Cloncurry district, northwest Queensland. Australian Journal of Earth Sciences
41: 381382.
Williams PJ (1994b) Iron mobility during synmetamorphic alteration in the Selwyn
Range area, NW Queensland: Implications for the origin of ironstone-hosted AuCu
deposits. Mineralium Deposita 29: 250260.
Williams PJ (1998) An introduction to the metallogeny of the McArthur River-Mount
Isa-Cloncurry minerals province. Economic Geology 93: 250260.
Williams PJ (2010a) Classifying IOCG Deposits. Short Course Notes, vol. 20,
pp. 2338. Canada: Geological Association of Canada.
Williams PJ (2010b) Magnetite-Group IOCGs with Special Reference to Cloncurry (NW
Queensland) and Northern Sweden: Settings, Alteration, Deposit Characteristics,
Fluid Sources and Their Relationship to Apatite-Rich Iron Ores. Short Course Notes,
vol. 20, pp. 2338. Canada: Geological Association of Canada.
Williams PJ, Barton MD, Fontbote L, et al. (2005) Iron-oxidecoppergold deposits:
Geology, spacetime distribution, and possible modes of origin. Economic Geology
100th Anniversary Volume, 371406.
Williams PJ, Dong G, Ryan CG, et al. (2001) Geochemistry of hypersaline fluid
inclusions from the Starra (Fe oxide)-Au-Cu deposit. Cloncurry district,
Queensland: Economic Geology 96: 875883.
Williams PJ, Kendrick MA, and Xavier RP (2010) Sources of ore fluid components in
IOCG deposits. In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and
Related Deposits: A Global Perspective, pp. 107116. Adelaide: PGC Publishing.
Williams PJ and Pollard PJ (2001) Australian Proterozoic iron oxideCuAu deposits:
An overview with new metallogenic and exploration data from the Cloncurry District,
Northwest Queensland. Exploration and Mining Geology 10: 191213.
Xavier RP, Monteiro LYS, Rusk BG, and Torresi I (2011) Comparing iron oxide copper
gold systems of the Carajas mineral province (Brazil) using apatite and magnetite
trace elements: Implications on genesis and exploration. In: Proceedings of the 11th
Biennial SGA Conference, pp. 461463.
Xavier RP, Monteiro LVS, Souza Filho CR, et al. (2010) The iron oxide copper-gold
deposits of the Carajas mineral province, Brazil: An updated and critical review.
In: Porter TM (ed.) Hydrothermal Iron Oxide CopperGold and Related Deposits:
A Global Perspective, vol. 3, pp. 285306. Adelaide: PGC Publishing.
Xavier RP, Rusk BG, Emsbo P, and Lena YSM (2009) Composition and source of
salinity of ore-bearing fluids in CuAu systems of the Carajas mineral province,
Brazil. In: Proceedings of the 10th Biennial SGA Conference, pp. 272274.
Townsville, Australia: Society for Geology Applied to Ore Deposits.
Xavier RP, Wiedenbeck M, Trumbull RB, et al. (2008) Tourmaline B-isotopes fingerprint
marine evaporites as the source of high-salinity ore fluids in iron oxidecopper
gold deposits, Carajas mineral province (Brazil). Geology 36: 743746.
Yang X, Sun W, Zhao T, and Deng J (2011) The MiddleLower Yangtze Metallogenic
Belt. International Geology Review 53: 447448.
Yu J, Mao J, and Zhang C (2008) The possible contribution of a mantle-derived fluid to
the Ningwu porphyry iron deposits Evidence from carbon and strontium isotopes
of apatites. Progress in Natural Science 18: 167172.
Zaw K, Huston DL, Large RR, Memagh T, and Hoffmann CF (1994) Microthermometry
and geochemistry of fluid inclusions from the Tennant Creek goldcopper deposits:
Implications for ore deposition and exploration. Mineralium Deposita 29: 288300.
Zhang RH (1986) Sulfur isotopes and pyriteanhydrite equilibria in a volcanic basin
hydrothermal system of the Middle to Lower Yangtze Valley. Economic Geology
81: 3245.
Zhang R, Zhang X, and Hu S (2011) Pyriteanhydritemagnetitepyroxene-type
deposits and coexisting hydrothermal fluids in Mesozoic volcanic basins, Yangtze
River Valley, China. Ore Geology Reviews 43: 315332.

You might also like