You are on page 1of 16

©2020 Society of Economic Geologists, Inc.

Economic Geology, v. XXX, no. XX, pp. X–X

SCIENTIFIC COMMUNICATIONS

OPENING THE MAGMATIC-HYDROTHERMAL WINDOW: HIGH-PRECISION U-Pb GEOCHRONOLOGY


OF THE MESOPROTEROZOIC OLYMPIC DAM Cu-U-Au-Ag DEPOSIT, SOUTH AUSTRALIA

Liam Courtney-Davies,1,† Cristiana L. Ciobanu,1 Simon R. Tapster,2 Nigel J. Cook,1 Kathy Ehrig,3 James L. Crowley,4
Max R. Verdugo-Ihl,1 Benjamin P. Wade,5 and Daniel J. Condon2
1School of Chemical Engineering and Advanced Materials, The University of Adelaide, Adelaide, South Australia 5005, Australia
2NERC Isotope Geoscience Laboratory (NIGL), British Geological Survey, Keyworth, Nottinghamshire, NG12 5GG, United Kingdom
3BHP Olympic Dam, Adelaide, South Australia 5000, Australia
4Department of Geosciences, Boise State University, Boise, Idaho 83725, USA
5Adelaide Microscopy, The University of Adelaide, Adelaide, South Australia 5005, Australia

Abstract
Establishing timescales for iron oxide copper-gold (IOCG) deposit formation and the temporal relationships
between ores and the magmatic rocks from which hydrothermal, metal-rich fluids are sourced is often depen-
dent on low-precision data, particularly for deposits that formed during the Proterozoic. Unlike accessory
minerals routinely used to track hydrothermal mineralization, iron oxides are dominant components of IOCG
systems and are therefore pivotal to understanding deposit evolution. The presence of ubiquitous, magmatic-
hydrothermal U-(Pb)-W-Sn-Mo–bearing zoned hematite resolves a range of geochronological issues concerning
formation of the ~1.6 Ga Olympic Dam IOCG deposit, South Australia, at up to ~0.05% precision (207Pb/206Pb
weighted mean; 2σ) using isotope dilution-thermal ionization mass spectrometry (ID-TIMS). Coupled with
chemical abrasion-ID-TIMS zircon dates from host granite and volcanic rocks within and enclosing the ore-
body, a confident magmatic-hydrothermal chronology is defined. The youngest zircon date from the granite
intrusion hosting Olympic Dam indicates magmatism was occurring up until 1593.28 ± 0.26 Ma. The orebody
was principally formed during a major mineralizing event following granite uplift and during cupola collapse,
whereby the hematite with the oldest age is recorded in the outer shell of the deposit at 1591.27 ± 0.89 Ma,
~2 m.y. later than the youngest documented magmatic zircon. Hematite dates captured throughout major
lithologies, different ore zones, and the ~2-km vertical extent of the deposit support ~2 m.y. of hydrothermal
activity. New age constraints on the spatial-temporal evolution of the formation of Olympic Dam are considered
with respect to a mantle to crustal continuum model. Cyclical tapping of magma reservoirs to maintain crystal
mushes for extended time periods and incremental building of batholiths on the million-year scale prior to
main mineralization pulses can explain the ~2-m.y. temporal window temporal window inferred from the data.
Despite the challenge of reconciling such an extended window with contemporary models for porphyry depos-
its (≤1 m.y.), formation of Proterozoic ore deposits has been addressed at high-precision and supports the case
that giant IOCG deposits may form over millions of years.

Introduction hydrothermal phases, proving invaluable for dating deposits


Constraining the onset and life span of mineralizing events is formed in Phanerozoic volcanic arcs in subduction or collision
critical for understanding the genesis of ore deposits and for zones (Chiaradia et al., 2013, and references therein). For ex-
the development of prospecting strategies within metallogen- ample, the combination of U-Pb zircon and Re-Os molybde-
ic provinces. However, identifying cases where mineralization nite data has enabled mineralizing intervals to be temporally
can be accurately dated and bracketed by causative intrusions resolved at both high analytical precision and spatial resolu-
or events can prove challenging. The duration and periodicity tion, in the range of tens of thousands to millions of years for
of intrusion-related mineralization events are commonly con- porphyry systems, where distinct ore-causative intrusions and
trolled by the tectono-magmatic settings in which they occur. alteration-mineralization stages can be bracketed by crosscut-
Timescales required for such processes have been proposed ting vein networks (e.g., von Quadt et al., 2011; Chiaradia et
to range from near instantaneous (Weatherley and Henley, al., 2014; Zimmerman et al., 2014; Li et al., 2017).
2013) to millions of years (Sillitoe and Mortenson, 2010). Ad- Difficulties occur when dating Precambrian deposits due
vances in radioisotopic dating techniques, notably chemical to increasingly larger absolute uncertainties with increasing
abrasion-isotope dilution-(negative)-thermal ionization mass absolute ages. Such uncertainties may conceal discrete min-
spectrometry (CA-ID-TIMS and ID-NTIMS), has permitted eralizing events or overestimate the duration of individual
routine high-precision dating of discrete magmatic pulses and events. Furthermore, minerals that can confidently record
a mineralization date relative to magmatic activity are of-
ten elusive, either due to modification of primary isotopic
†Corresponding author: e-mail, liam.courtney-davies@adelaide.edu.au systematics or simply because they do not form at the con-
ISSN 0361-0128; doi:10.5382/econgeo.4772; 16 p.
Digital appendices are available in the online Supplements section. 1 Submitted: September 30, 2019 / Accepted: April 21, 2020

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
2 SCIENTIFIC COMMUNICATIONS

ditions of mineralization. Developments in analysis of the (e.g., Kirchenbaur et al., 2016; Schlegel et al., 2017; Court-
hematite U-Pb mineral geochronometer have allowed direct ney-Davies et al., 2019c; Dmitrijeva et al., 2019a).
dating of ores derived from iron oxide copper-gold (IOCG), The Olympic Dam deposit is entirely contained within the
porphyry-related Au, unconformity-related, and banded Olympic Dam breccia complex (Reeve et al., 1990). Breccias
iron formation deposits (Ciobanu et al., 2013; Courtney- are hosted and largely derived from the Roxby Downs Gran-
Davies et al., 2016, 2019a, b; Zhou et al., 2017; Walter et al., ite, a Hiltaba Suite pluton of the Burgoyne batholith (Fig.
2018; Keyser et al., 2019). Such a tool is significant due to 2). Emplacement of the Roxby Downs Granite occurred at
the conspicuous lack of dateable crosscutting features within depths of 6 to 8 km (plagioclase-amphibole thermobarometry;
breccia-hosted IOCG systems. Kontonikas-Charos et al., 2017), with equivalent estimates
The archetypal example of an IOCG system sensu stricto is obtained for a quartz monzonite of the Burgoyne batholith
the Mesoproterozoic Olympic Dam deposit, South Australia distal from Olympic Dam (Creaser, 1996). A temporal range
(Ehrig et al., 2012), containing a total resource of 10,892 mil- in Hiltaba Suite plutons (up to ~30 m.y.) has been reported
lion tonnes (Mt) at 0.73% Cu, 0.24 kg/t U3O8, 0.31 g/t Au, across the Gawler craton (Fanning et al., 2007). Five samples
and 1.2 g/t Ag (BHP, 2019). Located within one of the rich- of shallow (<681 m) Roxby Downs Granite from three drill
est Precambrian metallogenic provinces on Earth, Olympic holes ~2 km south of the Olympic Dam breccia complex were
Dam hosts U-bearing, oscillatory-zoned hematite as a ubiqui- previously dated by zircon CA-ID-TIMS (Boise State Univer-
tous component of ores throughout the 6-km strike and 2-km sity; Cherry et al., 2018), yielding a pooled weighted mean of
depth of the deposit (Verdugo-Ihl et al., 2017). Moreover, U- 1593.03 ± 0.21 Ma (n = 36, mean square of weighted deviates
bearing hematite is found within all major lithologies, includ- [MSWD] = 1.08), suggesting the Roxby Downs Granite was
ing weakly mineralized granite forming the outer shell of the not emplaced incrementally. However, Hiltaba Suite granite
deposit (Verdugo-Ihl et al., 2020). at the Acropolis IOCG prospect, 25 km south of Olympic
The presence of an abundant U-(Pb)–bearing phase, which Dam, dated at 1594.88 ± 0.50 Ma (n = 6, MSWD = 0.7) by
can be radioisotopically robust (Courtney-Davies et al., CA-ID-TIMS (McPhie et al., 2020), indicates that the larger
2019a), makes Olympic Dam particularly well suited for tem- Burgoyne batholith was built incrementally.
porally constraining the magmatic-hydrothermal window (i.e., Several issues of debate remain concerning formation of the
the relative ages of the two systems), by integrating hydro- Olympic Dam breccia complex and, implicitly, the contained
thermal hematite and magmatic zircon (CA-)ID-TIMS U-Pb mineralization. Concentric zoning of sulfides (deeper, outer
geochronology. chalcopyrite-pyrite, to bornite-chalcopyrite and shallow, inner
Data presented below are interpreted with respect to the bornite-chalcocite; Fig. 3) relative to a barren, high-Fe breccia
evolving genetic model for the deposit. Early concepts involv- center suggests fluids were supplied from a central diatreme
ing shallow granite emplacement with diatreme-maar devel- (Reeve et al., 1990). However, this model was redefined with
opment (Reeve et al., 1990; Haynes et al., 1995; Johnson and the discovery of mineralization across a ~2-km-thick interval
Cross, 1995) have been superseded by deep granite crystal- in the deposit’s southeastern lobe (Fig. 3; Ehrig et al., 2012).
lization (Kontonikas-Charos et al., 2017), whereby hydrother- This suggests that there could have been deeper centers offset
mal fluids accumulate and form ores during uplift, breccia- to the zoning or, alternatively, that the concentric ore distri-
tion, and cupola collapse of the host granite (Verdugo-Ihl et bution results from superposition of primary (upward zona-
al., 2020). The absolute and inferred timing of such processes tion preserves high-temperature species in Cu-(Fe) sulfides;
are investigated. Ciobanu et al., 2017) and secondary (replacement) processes.
Defining a temporal timeframe for the deposit is complicat-
Background ed by the extensive, multistage brecciation within a volumi-
The eastern Gawler craton hosts the Mesoproterozoic Olym- nous and heterogeneous breccia complex (Reeve et al., 1990;
pic Cu-Au province (Skirrow et al., 2007; Reid, 2019), a 6 × 8 km in plan; Oreskes and Einaudi, 1990; and at least
metallogenic belt that strikes roughly north-south for 700 km ~1.9 km in vertical extent, Ehrig et al., 2012), which under-
and encompasses several dozen IOCG-style deposits and went pervasive, hematite-sericite alteration (Ehrig et al.,
prospects (Fig. 1). At the center of the province (the Olympic 2012). The deposit also straddles intersections between re-
Dam district), geology includes a sequence of gneissic rocks gional fault zones (Hayward and Skirrow, 2010), where long-
(the Mulgathing Metamorphic Complex, ~2.5 Ga), intruded lived reactivated faults have dismembered large segments
by Donington Suite granitoids (~1.85 Ga), all unconformably of the orebody (Clark et al., 2018). High-precision dating of
overlain by metavolcanic-sedimentary rocks of the Wallaroo magmatic zircon distal to the orebody (drill hole RD2488;
Group (~1.75 Ga; Fig. 1) and concealed under thick Neo- Fig. 2A) provides a maximum age constraint for hydrothermal
proterozoic cover. Breakup of the Columbia/Nuna super- activity. Historically, a major, broad mineralizing event occur-
continent triggered a craton-scale tectono-magmatic event at ring at ~1.6 Ga has been recognized at Olympic Dam using a
~1.6 Ga, forming the >25,000 km2 (Blissett et al., 1993) Gawl- variety of minerals and dating techniques (Reeve et al., 1990;
er silicic large igneous province (SLIP; Allen et al., 2008). The Creaser and Cooper, 1993; Johnson and Cross, 1995).
SLIP is composed of the intrusive/extrusive Gawler Range There is now a growing body of laser ablation-inductive-
Volcanics and comagmatic Hiltaba Suite granitoids. Although ly coupled plasma-mass spectrometry (LA-ICP-MS) U-Pb
highly debated, the source of IOCG mineralizing fluids has data, which broadly constrain the hydrothermal system to
been largely attributed to Hiltaba Suite granitoids across the ~1590 Ma. Hematite dates were first acquired by Ciobanu et
province through a combination of studies applying geochem- al. (2013) and corroborated by samples from different areas
ical modeling, geochronology, and mineralogical observations of the deposit generating comparable data (Courtney-Davies

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
SCIENTIFIC COMMUNICATIONS 3

et al., 2016; Apukhtina et al., 2017). Additional constraints at of U-bearing zoned hematite and the secondary hematite tex-
~1590 Ma were obtained from fluorapatite and uraninite in tures, which can be tied to replacement of the primary zoned
altered Roxby Downs Granite and from volcanic rocks outside type (Verdugo-Ihl et al., 2017, 2020), clearly indicate that U was
the orebody, respectively (Apukhtina et al., 2017). Irrespec- an integral component of the magmatic-hydrothermal IOCG
tive of the accuracy of LA-ICP-MS methods used, data gen- system. Therefore, high-precision dating of primary, zoned he-
erated thus far supports hydrothermal activity commencing matite from the aforementioned lithologies and ore zones may
broadly coeval with Roxby Downs Granite crystallization as provide a temporal framework for the hydrothermal system.
determined via ID-TIMS, with analytical uncertainties of ~5
to 25 m.y. for individual hydrothermal mineral dates. Rationale and Samples
Primary, oscillatory-zoned, U-(Pb)-W-Sn-Mo–bearing he- Although there is widespread evidence for multiple tectono-
matite is present throughout all ore zones and all major li- thermal events impacting on Olympic Dam, likely contribut-
thologies at Olympic Dam (Verdugo-Ihl et al., 2017). The ing to cycles of U and Cu remobilization and new mineral
majority of hematite in the orebody occurs either as replace- growth (e.g., 820 Ma dike swarms; Huang et al., 2015), this
ments or as overprinted variants of the zoned type via cou- does not necessarily imply multiple, major episodes of metal
pled-dissolution-(re)precipitation reactions, which can allow introduction. Previous studies of U, Pb, and Sm isotopes
partial preservation of preexisting trace element signatures have suggested that post ~1590 Ma events contributed to
(Verdugo-Ihl et al., 2017). and upgraded the large U resource within the deposit (e.g.,
There is an intimate association between Fe metasomatism Johnson, 1993, Meffre et al., 2010; Maas et al., 2011; Kirch-
and Cu mineralization, displayed by the outline of the Cu re- enbaur et al., 2016; Cherry et al., 2017). At present, how-
source mirroring the >5 wt % Fe boundary within the breccia ever, there are no geochronological constraints to confirm
complex (Fig. 2; Ehrig et al., 2012). The ubiquitous presence or refute this.

A South Australia Olympic Dam district

Gawler craton
Vulcan

Adelaide
Olympic Dam
IOCG system
Archean to early-
Mesoproterozoic faults Acropolis
B Wirrda Well Islan
Island
1595-1585 Ma tectonothermal event

Felsic GRV Dam

Hiltaba Suite

Mafic GRV
Oak Dam West
Emmie Bluff

Kimban orogeny
1.79 - 1.75 Ga

BIFs
Wallaroo Group
metavolcanic Carrapateena Fig. 1. (A) Map of South Australia overlain by the
sedimentary units outline of the Gawler craton, with the location of
the Olympic Dam district boxed, and enlarged,
displaying a geologic map of the region (right).
2.5 Ga 1.85 Ga

Donington Suite
granitoids The location of selected IOCG systems and major
faults are marked. (B) Generalized diagram depict-
ing relationships between the different country
Mulgathing rocks and IOCG host lithologies within the Olym-
Complex 15 km pic Dam district. GRV = Gawler Range Volcanics.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
4 SCIENTIFIC COMMUNICATIONS

A altered,
1 km B

RD
weakly brecciated
RDG RD2765
NW-arm Outer shell

RU41-9882
RU65-7976
6632000N RU36 - 9867 B
RD647
6630000N

RD1988
RD2366
Cu - resource
SE-Lobe A
RD2786A
6628000N
Masher’s fault
biotite ‘out’
RDG

678000E RD2488 682000E 1 km


Hematite breccia
Vertical drill hole
Fe (<5 wt%) Granite (RDG) Hm-Qz sediments
Subvertical drill hole
Fe (5-20 wt%)
Volcanic breccia Mafic-sediments Cross-section line
Fe (>20 wt%)
Fig. 2. (A) Sketch of the Olympic Dam deposit including major lithologies and sampled/discussed drill holes and drill hole
projections. (B) Expanded map of the southeastern lobe, marking the cross section A-B, shown in Figure 3. Hm = hematite,
Qz = quartz, RDG = Roxby Downs Granite.

The absence of evidence for extensive mineralizing events con located within ~50 m of sampled hematite were dated by
post-1590 Ma prompted this study to define and refine the CA-ID-TIMS (Fig. 3; Table 1).
timing and extent of the major, initial ore-forming event, us- We assess whether hematite can capture discrete ore-form-
ing a single mineral with a traceable geochemical signature. ing stages from geochemically diverse locations across the
Spanning all ore zones and major lithologies, zoned hematite deposit. For example, zoned hematite contains distinct sig-
from the deposit center to outer shell (nine samples) was se- natures with respect to high field strength elements (HFSEs;
lected for ID-TIMS dating (Fig. 2A, B; Table 1). To provide e.g., Ti, Th) in specimens from the deposits outer shell but
further constraints on the maximum timing of mineralization not in the orebody (Verdugo-Ihl et al., 2020; Fig. 4). Such a
from samples spatially associated with zoned hematite, vol- signature is indicative of an early hydrothermal origin, since
canic zircons within the breccia complex and magmatic zir- these elements are attributable to the magmatic stage. Chang-

Table 1. Details of Studied Samples

Relative level
Sample Lithology depth (m) Location Drill hole Ore zone
Hematite
7295BW Hematite-rich breccia 1,720 Southeastern lobe RD1988 Pyrite-chalcopyrite
LCD5 1,578 RD2786A Chalcopyrite-bornite
MV026 1,583 Pyrite-chalcopyrite
LCD6/MV027 Deep-bedded rocks 1,633 Chalcopyrite-bornite
LCD35 Volcanic breccia 821 RD647 Sulfide-poor
MV101 Hematite-quartz breccia 650 Center RD2765 Sulfide-poor
CLC201 High-grade bornite ore 425 Northwest arm RU41-9882 (Chalcopyrite)-bornite
CLC10 Hematite breccia 750 Outer shell RU65-7976 Pyrite-chalcopyrite

Zircon
LCD1 Granite (Roxby Downs 1,516 RD2786A
LCD11 Granite) 1,810 Southeastern lobe
LCD12 1,830
LCD13 1,850
LCD30A Volcanic breccia (Gawler 604 RD647
Range Volcanics)

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
SCIENTIFIC COMMUNICATIONS 5

es in the rare earth element (REE) signatures of fluorapatite extraction method, extensive hematite ID-TIMS methodol-
(Krneta et al., 2017, 2018) and previous dating of a younger ogy, and assessment of hematite U-Pb isotope systematics can
hematite in high-grade bornite ores (Ciobanu et al., 2013; Fig. be found in Appendix 1 and Courtney-Davies et al. (2019a).
4A, B) suggest protracted fluid introduction, which may be Hematite fractions were spiked with the NIGL-2 tracer solu-
reflected in high-precision dating. Packages of bedded he- tion. Zircon separated from core samples was selected for CA-
matite and quartz are present at depth (up to ~1.6 km) and ID-TIMS at both the BGS and the Isotope Geology Labora-
can be used to constrain a geodynamic model for the deposit, tory, Boise State University. Grains were imaged at Boise in
whereby the timing of their introduction into the Olympic cathodoluminescence (CL) mode using a scanning electron
Dam breccia complex can be understood through dating of microscope. Analyses were composed of single zircon grains
in situ hydrothermal hematite (Figs. 3, 4C, D). More broadly, or grain fragments, using a method modified after Mattinson
the resolution at which Proterozoic IOCG ore deposits can be (2005). All zircon analyzed at the BGS was spiked with the
accurately dated is interrogated. ET235 tracer solution, while zircon at Boise was spiked with
either BSU1B or ET2535 (App. 1, Table A1). Broader labo-
Methodology ratory and analytical details for hematite and zircon CA-ID-
One-inch-diameter polished blocks were prepared from drill TIMS can be found in Appendix 1.
cores. Zoned hematite grains were identified by scanning All hematite and zircon U-Pb data are presented at the 2σ
electron microscopy using an FEI Quanta 450 equipped with level and reduced using the U decay constants recommend-
a backscattered electron (BSE) detector and an energy dis- ed by Jaffey et al. (1971), with a preferred 238U/235U value
persive X-ray spectrometer at Adelaide Microscopy, The Uni- of 137.8185 ± 0.045 (2σ; Hiess et al., 2012). This value was
versity of Adelaide. selected due to its close comparability with direct U isotope
Hematite dating was conducted at the Natural Environ- studies of Olympic Dam, where a value of 137.801 was ob-
ment Research Council, Isotope Geoscience Laboratory, Brit- tained from whole-rock samples (Kirchenbaur et al., 2016).
ish Geological Survey (BGS). Full details of the in situ grain It should be noted that previous studies dating the Roxby

A RD1988 SE-lobe RD647 B

A B
faults

-400

No sulfide LCD30A
Bn-Cc
RL depth (m)

-800
LCD35
Cp-Bn

Py-Cp
-1200
RD2786A

LCD1
LCD4, 5 & MV026
-1600
LCD6 & MV027 Deep
LCD11-13 bedded
RX7295 500m rocks RDG

Hematite breccia Granite (RDG) Chloritized mafic sediments Drill hole


Fe (<20 wt%) Volcanic breccia Hm-Qz sedimentary rocks Drill hole off section
Fe (>20 wt%) Hm-Qz breccia Cover sequences
Sample

Fig. 3. (A) Cross section (A-B; marked on Fig. 2) through the southeastern lobe, with lithologies, drill holes, sample locations,
and ore zones marked. (B) Photographs of sampled bedded packages (left) and Roxby Downs Granite (RDG; right). Abbrevi-
ations: Bn-Cc = bornite-chalcocite, Cp-Bn = chalcopyrite-bornite, Hm = hematite, Py-Cp = pyrite-chalcopyrite, Qz = quartz.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
6 SCIENTIFIC COMMUNICATIONS

1000 μm
Deep bedded rocks

1000 μm

1000 μm

Fig. 4. (A-F) Representative coarse-scale photomicrographs of zoned hematite-bearing samples (left) and magnified BSE
images (right) of dated zoned hematite grains. (B) The white arrow and line marks the corroded grain boundary, populated
by a dense field of uraninite inclusions. (G) Isotopic LA-ICP-MS maps of a zoned hematite grain in sample CLC10, displaying
strong zonation in both U and Th (scale: n · 105). Abbreviations: Bn = bornite, Cp = chalcopyrite, Hm= hematite.

Downs Granite (Cherry et al., 2018) and Olympic Dam he- Isotope measurements were processed using the TRIPOLI
matite (Courtney-Davies et al., 2019a), used a 238U/235U value software package (Bowring et al., 2011). Data reduction, er-
of 137.88 (Steiger and Jäger, 1977); therefore, when age data ror propagation, and date calculations were generated using
from previously published studies are cited, they have been the Excel spreadsheet of Schmitz and Schoene (2007) for he-
modified using the same 238U/235U value applied in this study. matite and zircon at Boise; ET REDUX was used for zircon

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
SCIENTIFIC COMMUNICATIONS 7

at the BGS. Concordia and weighted average diagrams were extracted at 1595.9 ± 1.10 Ma (n = 2) and 1594.50 ± 1 Ma
constructed using the Excel add-in ISOPLOT 3.75 (Ludwig, (n = 4, MSWD = 0.27). The two older grains, of likely ante-
2012). Only analytical uncertainties are presented; uncertain- crystic origin, are plotted in yellow (Fig. 5C).
ties arising from the 238U/235U value uncertainty, different From the shallower sample (LCD1), six zircons with varying
tracer solution systematics, comparing hematite to zircon, degrees of Pb loss define a lower intercept at 546 Ma (Dela-
and data reduction methods are not included because of the merian orogeny), and upper intercept at 1595.97 ± 0.94 Ma.
negligible additional uncertainty and low relative uncertainty (Fig. 5D).
compared to the decay constant uncertainty for a sample of Two populations can be extracted in 207Pb/206Pb space at
~1.6 Ga (207Pb/206Pb uncertainty is ~5.6 Ma). 1595.10 ± 1.10 Ma (n = 2,) and 1593.28 ± 0.26 Ma (n = 4,
Isotope maps of hematite were acquired using a RESOlu- MSWD = 1.20), with the former likely representing antecrys-
tion-LR 193-nm ArF excimer laser microprobe (Australian tic material, exerting influence on the upper intercept date.
Scientific Instruments) coupled to an Agilent 7900cx quad- Two further samples of Roxby Downs Granite from the
rupole ICP-MS. The maps, scaled in counts per second (cps), same drill hole (sampled within 40 m) were prepared for
were compiled and processed into 2-D images displaying the CA-ID-TIMS at Boise University (Fig. 5E, F). Five grains
combined background-subtracted intensities using IOLITE were analyzed from sample LCD11 (using either ET2535 or
(Paton et al., 2011; see Verdugo-Ihl et al., 2017, for further BSUB1 tracer solutions; App. 1, Table A2), with one grain
details). broken into two fragments that were analyzed separately
(fraction z4a, b). The poorly defined intercept line forms a
Results: Petrography and Geochronology date at 1594.2 ± 1.5 Ma (n = 6, MSWD = 4). Two dates from
Sample locations, drill hole identifications and depths relative the broken grain yield a 207Pb/206Pb weighted mean of 1595.10
to sea level (RL) are summarized in Table 1 and displayed ± 1.10 Ma (MSWD = 1.6; BSU1B tracer solution). This grain
in Figures 2 and 3, with corresponding geochronological data has much lower U and higher Th concentrations than the oth-
(isotopic ratios, dates, and U-Th-Pb contents) provided in er grains, inferring a different fractionation history and poten-
Appendix 1, Tables A1 and A2. tial antecrystic origin (App. 1, Table A2). The dates from the
other four grains yield a weighted mean of 1593.28 ± 0.26 Ma
Zircon (MSWD = 1.2; ET2535 tracer solution).
Concordia diagrams and weighted average dates for zircon are From sample LCD12, 11 grains spiked with BSU1B or
presented in Figure 5. CL images of whole and fragmented ET2535 (Fig. 5F) generated poorly defined intercept dates
magmatic Roxby Downs Granite zircon highlight typical mag- with high degrees of Pb loss. This inhibits calculation of an
matic oscillatory zonation patterns (Fig. 5A; sample LCD12). acceptable weighted mean when including all data. When
These are invariably crosscut by swelling cracks induced by considering only the most precise data points, which have not
time and actinide-dependent metamictization. Grains aver- been (likely) leveraged by ancient Pb loss or contaminated by
age ~150 μm in length, smaller than zircon found within the antecrystic material, a 207Pb/206Pb date of 1593.06 ± 2.0 Ma
Olympic Dam breccia complex (Ciobanu et al., 2013). Ex- (n = 2; ET2535 tracer solution) is defined.
tended CL images are displayed in Appendix 1, Figure A1.
Gawler Range Volcanics: Polymictic volcanic clast-rich Hematite
breccia in a fine-grained matrix was sampled from the south- Concordia diagrams are presented in Figure 6, with a selec-
eastern lobe of the deposit (604 m), interpreted as belong- tion of BSE images of grains microsampled for ID-TIMS pre-
ing to the Gawler Range Volcanics (sample LCD30A; Fig. 3; sented in Figure 6A. Plots of 207Pb/206Pb age versus Th/U and
App. 1, Fig. A2). One large porphyritic clast was isolated from 206Pb/204Pb are used to discriminate data, defining the most

the breccia and crushed, with eight separated zircons ana- reliable data with fewest inclusions and highest radiogenic
lyzed at the BGS (Fig. 5B). Analyses are largely concordant versus common Pb at ~1590 Ma (Fig. 7).
(apart from one highly discordant data point) defining a zero- Southeastern lobe hematite-rich breccias: The deepest
age lower and upper intercept at 1594.55 ± 0.8 Ma, in good zoned hematite was sampled at 1,720 m (pyrite-chalcopyrite
agreement with the 207Pb/206Pb weighted average (1594.63 ± zone) from drill hole RD1988 (sample 7295BW). Seven ali-
0.71 Ma; n = 8, MSWD = 0.77). quots obtained from four grains returned a zero-age lower in-
Roxby Downs Granite: Granite hand specimens display tercept and upper intercept at 1591.7 ± 2.5 Ma (MSWD = 1.6)
macroscopic alteration at varying degrees but retain granit- (Fig. 6B). Individual 207Pb/206Pb dates range from 1606 to
oid textures and some relict magmatic minerals (Fig. 3B; see 1565 Ma, precluding calculation of an acceptable weighted
Kontonikas-Charos et al., 2017, for further details). Zircon mean date.
dated here underwent pervasive Fe-Cl metasomatism shortly Hematite-rich breccias from two locations in drill hole
after crystallization, during initiation of the hydrothermal sys- RD2786A show a similar spread in data, with highly vari-
tem at Olympic Dam (Courtney-Davies et al., 2019c), pos- able U concentrations, Pb loss, and common Pb components.
sibly influencing the U-Pb data presented. Sample LCD 5 forms a zero-age lower and upper intercept at
Zircon separated from two samples of Roxby Downs Gran- 1592.7 ± 2.2 Ma (n = 5, MSWD = 2.6) (Fig. 6C). Removal of
ite from the same drill hole (samples LCD13 and LCD1; Fig. the highly discordant and common-Pb-rich data point defines
3) were analyzed at the BGS (Fig. 5C, D). Six zircons were an upper intercept of 1588.8 ± 4.4 Ma (MSWD = 0.95; n = 4).
separated from the deeper sample (LCD13). The intercept Sample MV026 yields highly scattered analyses, with high
date is poorly constrained due to concordance of the data. As common Pb and Th/U returning an unconstrained lower in-
with the previous sample, two 207Pb/206Pb populations can be tercept and upper intercept at 1583 ± 14 Ma (n = 7, MSWD

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
8 SCIENTIFIC COMMUNICATIONS

A
0.2835

100 μm

RDG zircon (LCD12)


0.283

B Volcanic breccia
(LCD30A; BGS)
1600
C
0.2835
RDG
(LCD13; BGS) D
0.2835 RDG
(LCD1; BGS)
0.2825
0.281 1600
1590
0.2815

0.279 207
Pb/206Pb= 0.2805
1580 1594.63 ± 0.71 Ma 1590
n= 8, MSWD= 0.77 Antecrystic(?) zircon Non-zero age Pb-loss,
excluded no obtainable 207Pb/206Pb date
0.2795
0.277
Intercepts at -0.9 ± 3.9
207
Pb/206Pb=
and 1594.55 ± 0.81 Ma 1594.50 ± 1.0 Ma Intercepts at 546 ± 31
0.2785
n= 8, MSWD= 0.86 n= 4 MSWD =0.27 and 1595.97 ± 0.94 Ma
n= 6, MSWD= 1.6
Off diagram
3.72 3.76 3.80 3.84 3.76 3.78 3.80 3.82 3.84 3.86 3.88

E
0.2835 RDG
(LCD11; Boise)
1595 0.282 F
0.2835 RDG
(LCD12; Boise)
0.281
1590
0.280

0.280 0.278

1570
0.276
1585 Antecrystic (?) zircon
0.279
excluded
0.274
Pb/206Pb=
207
Non-zero age Pb-loss,
1593.28 ± 0.26 Ma no obtainable 207Pb/206Pb date
n= 4 MSWD =1.20 0.272
0.278

Intercepts at Intercepts at
2 ± 870 and 1594.2 ± 1.5 Ma 0.270 438 ± 77 and 1593.4 ± 1.1 Ma
n= 6, MSWD= 4.0 n= 11, MSWD= 26
0.277 0.268
3.75 3.77 3.81 3.83 3.60 3.64 3.68 3.72 3.76 3.80 3.84

Fig. 5. (A) Representative CL images of Roxby Downs Granite (RDG) zircon (sample LCD12). Grains are fragmented due
to the separation process and are invariably metamict and overprinted. (B-F) Wetherill concordia diagrams (CA-ID-TIMS)
for Gawler Range Volcanics zircon (B) and Roxby Downs Granite zircon (C-F). Intercept and weighted average (207Pb/206Pb)
dates are provided on the diagrams if obtainable. MSWD = mean square of weighted deviates.

= 63) (Fig. 6D). Individual 207Pb/206Pb dates taken alone can hematite grains, both displaying corrosion to grain margins
provide informative data; however, the spread precludes con- with nucleation of uraninite inclusions in marginal domains
fident weighted mean calculations. (Fig. 4B). Three concordant to reversely discordant data
High-grade bornite ore: Sample CLC201, from the north- points form an upper intercept at 1588.3 ± 1.6 Ma (MSWD
west arm of the deposit, represents an example of high-grade = 2.9; Fig. 6E), while the fourth is reversely discordant and
Cu mineralization (Fig. 4A), with abundant bornite and minor shifted to the left of all other data. The anomalous data point
chalcopyrite. Four aliquots were prepared from two zoned contains Th/U similar to other analyses (0.020), but it displays

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
SCIENTIFIC COMMUNICATIONS 9

7295BW LCD5 MV026 LCD35


A

300 μm 200 μm 400 μm 400 μm

0.4

B SE-Lobe breccias
(7295BW) C SE-Lobe breccias
(LCD5)
0.34
D SE-Lobe breccias
(MV026)
1800
0.32 Intercepts at
0.3 -838 ± 970 and 1583 ± 14 [±15] Ma
0.30
n= 7, MSWD= 63
1400
1650
0.28
0.2 1550
1000
0.26
1450

0.1
600 0.24
Intercepts at Intercepts at
-0.2 ± 2.6 and 1591.7 ± 2.5 [±6] Ma -2.03 ± 0.69 and 1592.7 ± 2.2 [±5.9] Ma 0.22
n= 7, MSWD= 1.6 n= 5, MSWD= 2.6
0.0 0.20
0 1 2 3 4 5 2.6 3.0 3.4 3.8 4.2

0.4

E High-grade ore
F G Volcanic breccia
5
0.285
Deep sedimentary rocks
(CLC201) 1610
0.4 (LCD6/MV027) Off diagram (LCD35)

Intercepts at 0.3 1800


0.283
-441 ± 350 and 1588.3 ± 1.6 [±5.0] Ma 1800
n= 3, MSWD= 2.9 0.3 Excl. 1400
1400
0.281
1590 0.2 Excl.
0.2 1000
1000
0.279

Excl. 0.1
600
0.1
600
0.277 Intercepts at Intercepts at
1570 0 ± 13 and 1587.6 ± 5.4 [±7.7] Ma -11 ± 37 and 1606 ± 14 [±15] Ma
n= 7, MSWD= 0.42 n= 4, MSWD= 0.42
0.275 0.0 0.0
3.55 3.65 3.75 3.85 0 1 2 3 4 0 1 2 3 4 5

H Outer shell
(CLC10) 1597 I box heights are 2s

5
1598
1
2
0.2815 1595 1

1594 3 4
1593

4
0.2805
1590 2 1591

1589
3
and Intercepts at
5 n=5, 277 ± 610 and 1591.8 ±1.3 [±6.5] Ma 1587
n =4, MSWD = 0.72
Mean=1591.27± 0.89Ma
n=5, MSWD = 0.52
3.785 3.795 3.805 3.815 3.825

Fig. 6. (A) Selection of BSE images showing zoned hematite grains, which were microsampled for ID-TIMS. (B-H) Wetherill
concordia diagrams for zoned hematite and corresponding intercept dates. Data marked Excl. is excluded from the intercept
calculations due to low U or high common Pb. Small error ellipses are arrowed. (I) 207Pb/206Pb weighted average diagram for
sample CLC10 from the deposit’s outer shell. MSWD = mean square of weighted deviates.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
10 SCIENTIFIC COMMUNICATIONS

206
Pb/204Pb dates spread from 1594 to 1581 Ma, precluding an acceptable
A weighted mean calculation.
(n= 50) Volcanic and barren deposit center breccias: The volcanic
1700
Roxby Downs Granite (Gawler Range Volcanics) breccia (sample LCD35) is he-
matite rich, with zoned grains up to 600 μm in length (Fig.
1600 6A). Hematite aliquots (five) sampled from two grains are
characterized by very low U contents (mean 15 ppm), low
206Pb/204Pb, and high Th/U. Despite this, with exclusion of
1500
one data point containing highest Th/U, a zero-age lower in-
tercept and an upper intercept at 1606 ± 14 Ma (MSWD =
1400 0.42) are obtainable (Fig. 6G). Hematite sampled from the
2σ uncertainty (Ma)
deposit center (sample MV101) is not shown in Figure 6, as an
1300 intercept cannot be formed from the scattered data.
Outer shell, northern deposit margin: Zoned hematite
Pb/206Pb (Ma)

grains are not pristine, with inclusion nucleation and porosity


100 1000 10000 developing on grain boundaries; however, the cores remain
clean and retain oscillatory zonation patterns. Compared to
B those within the orebody, grains are unusually rich in HFSEs,
(n= 50)
particularly Ti and Th, of which the latter mirrors U within
207

1700
zonation patterns (Fig. 4G). Further petrographic and geo-
chemical data for hematite in this sample and the outer shell
1600
was given by Verdugo-Ihl et al. (2020). Data are collected
from five aliquots (Fig. 6H) and microsampled in the center
1500 of two grains shown in Figure 4F and G. Uranium concentra-
tions ranged from 135 to 527 ppm (avg 331 ppm), with high
1400 radiogenic versus common Pb components for each aliquot
(avg 206Pb/204Pb = 8,505; Fig. 7). Of the five aliquots, four
are concordant, and one is highly discordant. The 207Pb/206Pb
1300
values for each aliquot are within uncertainty of one anoth-
er, with individual dates ranging from 1592.1 to 1590.6 Ma.
Including all data, a confident 207Pb/206Pb weighted average
.001 .01 .1 1 10
Th/U of 1591.27 ± 0.89 Ma (n = 5, MSWD = 0.52) is generated
(Fig. 6I). If the weighted average mean of the youngest zircon
SE-lobe breccias High-grade ore
Hematite population (LCD11B) and oldest hematite sample (CLC10)
Bedded rocks Outer shell
ID-TIMS are considered, a temporal interval of 2.01 m.y. is obtained.
Volcanic breccia Deposit center
When integrating the upper and lower limit of uncertainty for
Fig. 7. (A-B) Plots of 207Pb/206Pb (Ma) versus 206Pb/204Pb and Th/U, used to each weighted mean, the temporal interval ranges from 0.84
discriminate hematite data from different lithologies, in terms of its common to 3.16 m.y., decreasing in statistical probability with increas-
Pb content (A), and presence of inclusions (B). The most geochronologically ing limits of uncertainty.
reliable data clusters at ~1.59 Ga, with 2σ uncertainty bars increasing and
data points becoming more scattered when 206Pb/204Pb drops below ~1,000
and Th/U exceeds 0.1. Presented data can be found in Appendix 1, Table A1. Discussion
Formation of the Olympic Dam IOCG deposit: Constraints
on the magmatic-hydrothermal window
the lowest U and highest common Pb concentrations. Indi- Considering the new geochronological data in the context
vidual 207Pb/206Pb dates yield low relative uncertainties; how- of generic crustal continuum models for IOCG formation
ever, a weighted average with an acceptable MSWD cannot (Groves et al., 2010) and reconstructions of lithospheric archi-
be generated due to high scatter from ~1589 to ~1584 Ma. tecture for the eastern Gawler craton (Skirrow et al., 2018),
Deep-bedded lithologies: Bedded, fine-grained rocks are we suggest the following spatial-temporal evolution for for-
found at depths of at least 1.7 km (pyrite-chalcopyrite zone). mation of the Olympic Dam deposit (Fig. 8). Intracontinen-
The sampled rocks occur as a 44-m-thick interval, composed tal extension during the Gawler SLIP event induced crustal
of rhythmic bands of alternating hematite or quartz domi- thinning and emplacement of large, felsic bodies of magma
nance. Zoned hematite is found along, but also crosscutting, such as the Burgoyne batholith at upper crustal levels, below
bedding—i.e., along trails of sulfides, with orientation oblique active Gawler Range volcanoes (Fig. 8A). Bimodal magma-
to bedding. These packages were sampled in two locations 1 tism is evident from mafic lava flows and dikes/sills, including
m apart (LCD6 and MV027; Figs. 3B, 4G, H). Nine aliquots olivine-phyric basalts and dikes intersected throughout the
from four grains were prepared, with eight analyses forming Olympic Dam district (Huang et al., 2016), thus supporting
a zero-age lower intercept and an upper intercept at 1587.6 the hypothesis of deep-rooted magmatism derived from a
± 5.4 Ma (MSWD = 0.42; Fig. 6F). All analyses are either heterogeneous mantle source (Wade et al., 2019). Central to
discordant or reversely discordant, and individual 207Pb/206Pb magma fertility in the region is magma mingling at the sub-

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
SCIENTIFIC COMMUNICATIONS 11

Gawler SLIP ~1.6 Ga Cupola collapse and formation of


A B Magmatic-hydrothermal C
Intra-continental extension window the Olympic Dam Breccia Complex
GRV zircon
1594.63 Ma Burgoyne

Faults
Oldest hematite
~2-4 km 1591.27 Ma
GR
Hiltaba V

~2 Ma
RDG Uplift
Wallaro Suite
o Grou
Doning p Oute
ton Su r
ite shell
~6-8 km RDG zircon
Basem 1593.28 Ma
ent
~35 km
Vertical scale
exaggerated

Magma Youn
gest h y
reservoir e
1588 matite bod
SCLM ~20 km .95 M Ore
a
G
Felsic RD
ikes
g lsic d
eltin Mingling D
magmeep, late, n d fe
Zone lm ~35 km
fic a
Mafic of partia zone atic pu
lses(? Ma
)

Fig. 8. Schematic diagrams. (A) Mantle-crustal model for Olympic Dam district at ~1.6 Ga. (B) Assembly of the Roxby Downs
Granite (RDG) magmatic chamber and uplift/exhumation over ~2 m.y. (C) Cupola collapse resulting in formation of the
Olympic Dam orebody. GRV = Gawler Range Volcanics, SCLM = subcontinental lithospheric mantle, SLIP = silicic large
igneous province.

continental lithospheric mantle and/or crustal boundary. Evi- ference suggests that average rates of exhumation/uplift must
dence for such phenomena has been inferred by the presence have been in the range of 1 to 2.5 mm/year (uplift rate from
of feldspar rapakivi textures and mafic enclaves in the Roxby 6–8 to 2–4 km over 2 m.y.). Although this scenario resembles
Downs Granite (Krneta et al., 2016; Kontonikas-Charos et al., the generic template for exsolution of magmatically derived
2017). fluids in a supercritical state to produce porphyry mineral-
The plutons that form the Burgoyne batholith (Creaser and ization (e.g., Candela, 1991; Candela and Blevin, 1995), it
Cooper, 1993) must have been fed by a deep undifferenti- cannot explain how a single intrusion can retain these fluids
ated magma reservoir lying at a crustal depth of ~20 km (Fig. over timescales >1 m.y., as it should solidify quickly, exsolv-
8B). Despite compositional differences among intrusive rocks ing fluids via isobaric crystallization (e.g., Chiaradia and Cari-
of the Burgoyne batholith, individual granitic subsuites and cchi, 2017). Latest numerical models for ore generation in
plutons yield zircon ages within uncertainty of one another a magmatic context postulate that a magma reservoir with
(Creaser and Cooper, 1993; Jagodzinski, 2005) or, in one case, abundant crystal mush and a tapping mush can maintain and
ages slightly older than the Roxby Downs Granite (McPhie produce magmatic volatile phases as metal-bearing bubbles
et al., 2020). A first magmatic crystallization stage in Roxby (immiscible phase) during periodic replenishment episodes
Downs Granite at a depth of ~6 to 8 km (Kontonikas-Cha- (Vigneresse and Truche, 2020). Increased fluid advection
ros et al., 2017) is constrained from zircon crystallization at leads to a switch in pressure from a supercritical, lithostatic to
~1593.2 Ma. However, coexisting antecrystic zircon with an hydrostatic regime when the crystal mush becomes too rigid
age of ~1595 Ma is likely recycled from the deeper magmatic to resist volatile buoyancy. We suggest that magma replen-
plumbing system (Fig. 8B). ishment within the Roxby Downs Granite reservoir at 6 to
The ~2 m.y. age difference between magmatic Roxby 8 km, after initiation of crystallization (Kontonikas-Charos et
Downs Granite zircon (1593.28 ± 0.23 Ma) and earliest hy- al., 2017), will support cycles of crystal mush and a tapping
drothermal hematite (1591.27 ± 0.89 Ma; Fig. 9) implies a mush (Vigneresse and Truche, 2020), thus maintaining a mol-
scenario analogous to that invoked by stochastic models for ten and/or mush state of the granite beyond the time frames
large porphyry system formation (Chiaradia and Caricchi, estimated for porphyry systems.
2017). In such models, magma that accumulated in the lower Early magnetite formation (preserved in the outer shell),
to middle crust (~20 km) during an interval of ~2 to 3 m.y. is predates a sudden depressurization that leads to cupola col-
transferred and degassed in the upper crust over time spans of lapse, hydraulic brecciation of both the granite and country
up to 2 m.y. High-precision dating of most porphyry systems, rocks, and deposition of hematite-dominant ore (Fig. 8C;
however, indicates time spans of no more than 1 m.y. between Verdugo-Ihl et al., 2020). The latter event is constrained by
magmatic zircon and hydrothermal minerals (e.g., Chiaradia the hematite age of 1591.27 ± 0.89 Ma. Assuming uplift rates
et al., 2013; Li et al., 2017). as above, this would correspond to a depth of 2 to 4 km and
If we assume that the Olympic Dam mineralization is could have been triggered by catastrophic, destabilizing tec-
sourced from hydrothermal fluids exsolved from the Roxby tonic events (e.g., megaearthquakes) as considered for large
Downs Granite during cupola collapse, the ~2 m.y. age dif- porphyry systems (Richards, 2018). Hematite from the de-

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
12 SCIENTIFIC COMMUNICATIONS

207
Pb/206Pb (Ma) probability density

Zircon
Roxby Downs Granite (RDG)
RDG zircon
Pb-loss Gawler Range Volcanics

Anticrystic (?) Hydrothermal hematite


RDG zircon
Sample/207Pb/206Pb date Lab/Tracer
LCD4= 1588.95 ± 0.93 Ma
ODBC
D C n= 3 MSWD= 0.45
SE-lobe
E b (Courtney-Davies et al., 2019a)
BGS

‘Outer
e sshell’ CLC10 =1591.27 ± 0.89 Ma (NIGL-2)
n= 5 MSWD= 0.52

LCD12 = no obtainable Boise


weighted mean
(BSU1B)

Olympic Dam Boise


SE-lobe LCD12= 1593.06 ± 2.0 Ma
n= 2 MSWD =1.19 (ET2535)

LCD11B= 1593.28 ± 0.26 Ma Boise


B Youngest n= 4 MSWD =1.20
RDG crystallization (ET2535)
A date LCD11A =1595.10 ± 1.10 Ma
n= 2 MSWD =1.60 (BSU1B)
B
LCD13B= 1594.50 ± 1.0 Ma
n= 4 MSWD =0.27 BGS
LCD13A=1595.9 ± 1.10 Ma
A
n= 2 MSWD =0.080 (ET235)

LCD30A =1594.63 ± 0.71 Ma BGS


ODBC n= 8 MSWD =0.77
SE-lobe
e (ET235)

1602 1598 1594 1590 1586 1582


207
Pb/206Pb (Ma)
Fig. 9. 207Pb/206Pb weighted average diagram for individual Gawler Range Volcanics/Roxby Downs Granite zircon and
highest-precision hematite samples (color marked in key). The corresponding 207Pb/206Pb probability density plot for
combined weighted averages is presented (upper) and grouped by Gawler Range Volcanics (one sample), Roxby Downs
Granite (four samples), and hematite (two samples; sample LCD4 included from Courtney-Davies et al. (2019a). The
207Pb/206Pb dates for the weighted averages are presented to the right, with the laboratory they were analyzed at and the

tracer samples were spiked with. BGS = British Geological Survey, MSWD = mean square of weighted deviates, ODBC
= Olympic Dam breccia complex.

posit center is characteristically fine grained and intensely lower level (425 m), yielded young but highly scattered indi-
brecciated, demonstrating highly variable U-Pb systematics, vidual 207Pb/206Pb dates. The high Cu grade of these samples
likely due to formation directly within the collapsing cupola may be the result of younger fluid pulses, introduced late into
(Figs. 8C, 10). the hydrothermal system (Krneta et al., 2017, 2018), rather
Zoned hematite at depth in the southeastern lobe (~1.6 km) than fault reactivation, which has not offset the northwest of
records a confident (207Pb/206Pb) population at 1588.95 ± the deposit as seen in the southwest. However, the poorly con-
0.93 Ma (sample LCD4), i.e., ~4 m.y. after Roxby Downs strained and nonzero lower intercepts likely leverage the date.
Granite zircon crystallization. Such younger ages could pos- Variably altered mafic (Huang et al., 2016) and felsic (Johnson
sibly result from fault (re)activation, augmenting the orebody and Cross, 1995) dikes tied to the Gawler Range Volcanics/
and (re)opening fluid pathways, or from episodic magmatic Hiltaba Suite event have been interpreted to intrude the Roxby
pulses that fed magmatic-hydrothermal fluids for an additional Downs Granite and Olympic Dam breccia complex, although
~2 m.y. interval after the Roxby Downs Granite uplift. High- mineral assemblages dated so far provide only low-precision
grade bornite ores from the northwest arm, sampled at a shal- constraints at ~1590 Ma. Bedded, silica- and hematite-rich

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
SCIENTIFIC COMMUNICATIONS 13

238
U/235U= 137.818 11
Box heights are 2σ

207
1650 15*

Pb/206Pb (Ma)
12 14
1600
1 2 3 4 5 6 7* 8
1550
9 10
1500 13
Hematite
Roxby Downs Granite (zircon)
Gawler Range Volcanics (zircon)

Fig. 10. 207Pb/206Pb weighted average diagram displaying zircon and zoned hematite ID-TIMS aliquot for each sample within
this study. Samples labeled with an asterisk are included from Courtney-Davies et al. (2019a). Sample numbers: 1 = LCD30A,
2 = LCD12A, 3 = LCD13, 4 = LCD11, 5 = LCD12B, 6 = CLC10, 7 = LCD4, 8 = LCD5, 9 = LCD6, 10 = MV101, 11 =
LCD35, 12 = CLC201, 13 = 7295BW, 14 = 7295GM, 15 = MV026.

rocks found at depths of ~1.6 km yield an ~1590 Ma hema- centage of 95% after crystallization of hydrous minerals (bio-
tite population. This indicates either that preexisting rocks tite, amphibole). This assumption is consistent with the cal-
(Wallaroo Group?) were entrapped within the Roxby Downs culated H2O requirement to produce 5% hydrous minerals
Granite during uplift/brecciation and subsequently altered or in Roxby Downs Granite (Ehrig et al., 2012), based on H2O
that these units were incorporated into the breccia complex values for the two minerals given by Cline and Bodnar (1991).
from above via faults, postdeposit formation. Entrapment of A fit is obtained between the resource and calculated Cu
Fe-rich lithologies from the Wallaroo Group in the outer shell produced by the model (App. 2, Table A1) if we assume a
at Olympic Dam was also documented from U-Pb dating of melt volume (15- × 10-km area × 5-km depth = 750 km3)
~1760 Ma magnetite (Courtney-Davies et al., 2020). that is of similar order of magnitude to a realistic estimated
The only published fluid inclusion and stable isotope study volume for the Roxby Downs Granite. For example, using
invokes a magmatic signature for magnetite (~400°C, 10‰ D fCu
/m
of 80, 4 wt % H2O, melt density of 2,250 kg/m3, and 50
δ18O) but suggests mixing with seawater or basinal ground- ppm Cu in the initial melt, we obtain 64.1 Mt Cu extracted to
waters during the hematite stage (~200°–400°C, 9‰ δ18O; fluid. The volume of the Roxby Downs Granite is, in reality,
Oreskes and Einaudi, 1992). This interpretation is, however, unconstrained and may be significantly larger than estimated
based upon an earlier scenario in which the deposit formed from available data.
close to the paleosurface rather than at depth, as presented
here. The fluid source for hematite can only be magmatic- Implications
hydrothermal in origin when considering the granitophile The measured magmatic-hydrothermal window can be un-
signature of dated, zoned hematite found throughout the derstood through incremental building of batholiths, which
deposit (Verdugo-Ihl et al., 2017). This signature, compris- undergo intermittent replenishment via magma tapping. The
ing a U-W-Sn-Mo element association, is typically enriched Burgoyne batholith can be compared to the Yerington batho-
in granite-derived fluids (e.g., Kozlov, 2011) and extends and lith, Nevada, which is similar in magma volume (i.e., an area
overlaps with mineralization throughout the Olympic Dam of 250 km2 at up to 6-km depth; Schöpa et al., 2017). A forma-
breccia complex, down to >2-km depth, based on principal tional interval of up to ~1 m.y. was modeled for building the
component analysis of large whole-rock data sets (Dmitrijeva latest mineralized granite of the Yerington batholith (Schöpa
et al., 2019a). et al., 2017). These authors suggest that magma batches were
We can test this hypothesis by simplified mass balance cal- emplaced at uncommonly high rates (>4 cm/yr) but with peri-
culations (App. 2), assuming fluid exsolution from a parent ods of dormancy between them, exceeding 100 k.y. Incremen-
melt to produce a mineralization mass comparable to that in tal building of the Burgoyne batholith is feasible when consid-
the Olympic Dam deposit. We select Cu rather than Fe since ering the geochemically heterogeneous nature of individual
the latter can be influenced by incorporation of older, preex- plutons, including Roxby Downs Granite itself; however, no
isting, Fe-rich lithologies (~1.75 Ga; Courtney-Davies et al., post-1593 Ma zircon age has yet been obtained. The Roxby
2020). The reported resource at Olympic Dam is 79.45 Mt Cu Downs Granite, which hosts Olympic Dam, may thus be a
(10,892 Mt at 0.73 wt % Cu; BHP, 2019). Choosing a range final pulse of the Burgoyne batholith, after periods of barren
of parameters for average granite composition, including Cu granite intrusion (e.g., Acropolis; Dmitrijeva et al., 2019b).
contents of 20 to 100 ppm measured in Roxby Downs Gran- The proposed model does not, however, explain why this
ite, water estimate for typical granitic melts (e.g., 2.5–4 wt % particular intrusion appears to have been so efficient in gen-
H2O; Clemens, 1984; Holtz et al., 2001), copper partitioning erating mineralization compared to others of similar compo-
coefficients between a granitic melt and fluid, and D fCu /m
of 2 sition elsewhere in the same region, including in the same
to 80 (Keppler and Wyllie, 1991; Bai and Koster van Groos, batholith. One intriguing possibility is the role of abundant
1999), we can calculate the amount of Cu produced from a volatiles within the hydrothermal fluids (as evidenced, for ex-
volume of parent melt, assuming a total fluid exsolution per- ample, by the presence of Cl in zircon; Courtney-Davies et al.,

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
14 SCIENTIFIC COMMUNICATIONS

2019c). Equally likely is that a confluence of events (tectonics, represent the highest-precision geochronological study of a
cupola collapse) led to the fluids being captured, precipitated, Proterozoic IOCG deposit, underpinning the growing value
and concentrated as an orebody rather than lost. Diagrams of hematite to accurately date ore-forming events at levels of
showing Olympic Dam as a maar type of IOCG deposit (e.g., precision equal to or higher than more conventional methods
Groves et al., 2010, in press) should be abandoned based on (e.g., Re-Os dating of molybdenite; Chiaradia et al., 2013).
present evidence.
The feasibility of a ~2 m.y. magmatic-hydrothermal time Conclusions
span for IOCG deposits contrasts with current models for Three key geochronological constraints on the magmatic-
porphyry deposits, for which measured formational life spans hydrothermal genesis of the Olympic Dam deposit are dem-
are ever shorter with increasing analytical precision. This may onstrated here with confidence for the first time. Integrat-
imply that giant Mesoproterozoic deposits are distinct, with ing zircon and hematite ages underpins a model involving a
timescales of formation far exceeding those recognized for single, major mineralizing event, hinting at a temporal evolu-
Cenozoic magmatic-hydrothermal systems. tion of zoned hematite within distinct locations and lithologies
Considering the maximum and minimum hematite dates throughout the deposit, ~2 m.y. after zircon crystallization. To
yielding statistically acceptable MSWD at the 95% confidence the best of our knowledge, Olympic Dam hematite (U-Pb)
interval, a single, major mineralizing event lasted at least ID-TIMS geochronology has generated mean ages with the
~2 m.y. (Fig. 9). When considering the size of the Olympic highest (current) analytical precision for a hydrothermal min-
Dam orebody and large magma flux during the Gawler SLIP eral reported within any IOCG system. Hematite and zircon
event, a long-lived magmatic-hydrothermal event is plausible, should, therefore, be considered a key mineral pair to con-
resulting in metallogenesis at the regional scale. However, strain the time frames of U-rich, intrusion-hosted ore systems
variation in hematite 238U/235U upon crystallization may result of any age.
in longer than expected mineralization time frames. The ac-
cepted 238U/235U value for zircon (137.818; Hiess et al., 2012) is 1. The magmatic-hydrothermal system at Olympic Dam
well constrained, although there are no direct measurements became active ~2 m.y. after the youngest recorded zircon
for this value in hematite, imposing a limiting factor on the in the granite hosting Olympic Dam at 1593.28 ± 0.26 Ma
accuracy of hematite U-Pb data. However, Kirchenbaur et al. (2σ).
(2016) observed negligible variation from an average 238U/235U 2. The orebody was principally formed during a single, major
value of 137.801 determined from analysis of 40 whole-rock mineralizing event, whereby the earliest magmatic-hydro-
samples in the Olympic Dam breccia complex, suggesting a thermal activity is recorded in hematite from the outer
common, narrow range of initial 238U/235U in all major litholo- shell of the deposit at 1591.27 ± 0.89 Ma (2σ).
gies and implicitly in constituent minerals. Therefore, con- 3. Hematite ages constrained throughout all lithologies and
sidering the highest-precision, concordant hematite data and ore zones within the deposit suggest the magmatic-hydro-
likely narrow range in the 238U/235U values for hematite, a thermal system was active over a period of ~2 m.y.
~2 m.y. mineralizing time frame is suggested (Fig. 9). If sam- 4. Mesoproterozoic (~1.6 Ga) magmatic-hydrothermal events
pled hematite does contain a 238U/235U value akin to that re- can be accurately constrained at up to ~0.05% uncer-
ported by Kirchenbaur et al. (2016), the difference in calcu- tainty (207Pb/206Pb) using the hematite U-Pb mineral
geochronometer.
lated dates compared to the zircon data reduced with the value
of Hiess et al. (2012) would be negligible at the presented Acknowledgments
resolution. Concordant data points confirm that some hema-
This work is a contribution to the FOX project (trace elements
tite grains have remained near closed system and are therefore
in iron oxides: deportment, distribution and application in ore
unaffected by postcrystallization isotopic modification.
genesis, geochronology, exploration, and mineral processing),
The timescales discussed here emphasize the analytical re-
supported by BHP Olympic Dam and the South Australian
solving power required to capture and distinguish between
Government Mining and Petroleum Services Centre of Ex-
magmatic and magmatic-hydrothermal events—a level of
cellence. N.J.C. and K.E. acknowledge support from the
precision that is currently beyond the capabilities of micro-
beam methods. Uranium-bearing hematite demonstrates ARC Research Hub for Australian Copper-Uranium (grant
IH130200033). We gratefully acknowledge highly construc-
that relative precisions of ±0.1 and ±0.05% (and interpreted
tive and detailed reviews from Massimo Chiaradia and Adam
ages) for individual crystal fragments and 207Pb/206Pb grouped
Simon, which helped us to improve this manuscript.
weighted means, respectively, is achievable using Mesopro-
terozoic samples. Despite this, temporal differences between REFERENCES
ore zones within the deposit have not been resolved and Allen, S., McPhie, J., Ferris, G., and Simpson, C., 2008, Evolution and archi-
may not exist (Fig. 10). The data also highlights how individ- tecture of a large felsic igneous province in western Laurentia: The 1.6 Ga
ual plutons and their associated hydrothermal fluids can be Gawler Range Volcanics, South Australia: Journal of Volcanology and Geo-
very difficult to date, especially when the orebody is hosted thermal Research, v. 172, p. 132–147.
Apukhtina, O., Kamenetsky, V.S., Ehrig, K., Kamenetsky, M.B., Maas, R.,
within a breccia complex and largely composed from the en- Thompson, J., McPhie, J., Ciobanu, C.L., and Cook, N.J., 2017, Early
closing host granite. Without access to a much wider set of magnetite-fluorapatite mineralization at the Olympic Dam Cu-U-Au-Ag
samples from multiple depths across the batholith, requiring deposit, South Australia: Economic Geology, v. 112, p. 1531–1542.
substantial drilling of holes almost certain to be barren, the Bai, T., and Koster van Groos, A., 1999, The distribution of Na, K, Rb, Sr, Al,
true extent of magmatic-hydrothermal activity may never be Ge, Cu, W, Mo, La, and Ce between granitic melts and coexisting aqueous
fluids: Geochimica et Cosmochimica Acta, v. 63, p. 1117–1131.
fully constrained. Overall, combined zircon-hematite ages

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
SCIENTIFIC COMMUNICATIONS 15

BHP, 2019, BHP Annual Report 2019: 316 p., www.bhp.com/investor-centre/ for ore genesis and regional metallogeny: Ore Geology Reviews, v. 118,
annual-report-2019/. article 103337.
Blissett, A.H., Creaser, R.A., Daly, S.J., Flint, R.B., and Parker, A.J., 1993, Creaser, R.A., 1996, Petrogenesis of a Mesoproterozoic quartz latite-granitoid
Gawler Range Volcanics: Geological Survey of South Australia Bulletin, v. suite from the Roxby Downs area, South Australia: Precambrian Research,
54, p. 107–131. v. 79, p. 371–394.
Bowring, J.F., McLean, N.M., and Bowring, S.A., 2011, Engineering cyber Creaser, R.A., and Cooper, J.A., 1993, U-Pb geochronology of middle Pro-
infrastructure for U-geochronology: Tripoli and U-Pb redux: Geochemistry, terozoic felsic magmatism surrounding the Olympic Dam Cu-U-Au-Ag and
Geophysics, Geosystems, v. 12, p. Q0AA19. Moonta Cu-Au-Ag deposits, South Australia: Economic Geology, v. 88, p.
Candela, P.A., 1991, Physics of aqueous phase evolution in plutonic environ- 186–197.
ment: American Mineralogist, v. 76, p. 1081–1091. Dmitrijeva, M., Ciobanu, C.L., Ehrig, K., Cook, N.J., Metcalfe, A., Verdugo-
Candela, P.A., and Blevin, P.L., 1995, Do some miarolitic granites preserve Ihl, M.R., and McPhie, J., 2019a, Defining IOCG signatures through com-
evidence of magmatic volatile phase permeability?: Economic Geology, v. positional data analysis: A case study of lithogeochemical zoning from the
90, p. 2310–2316. Olympic Dam deposit, South Australia: Ore Geology Reviews, v. 105, p.
Cherry, A.R., McPhie, J., Kamenetsky, V.S., Ehrig, K., Keeling, J.L., 86–101.
Kamenetsky, M.B., Meffre, S., and Apukhtina, O.B., 2017, Linking Olym- Dmitrijeva, M., Ciobanu, C.L., Ehrig, K.J., Cook, N.J., Metcalfe, A., Ver-
pic Dam and the Cariewerloo basin: Was a sedimentary basin involved in dugo-Ihl, M.R., and McPhie, J., 2019b, Mineralization-alteration footprints
formation of the world's largest uranium deposit?: Precambrian Research, in the Olympic Dam IOCG district, South Australia: The Acropolis pros-
v. 300, p. 168–180. pect: Journal of Geochemical Exploration, v. 205, article 106333.
Cherry, A.R., Ehrig, K., Kamenetsky, V., McPhie, J., Crowley, J., and Ehrig, K., McPhie, J., and Kamenetsky, V., 2012, Geology and mineralogical
Kamenetsky, M., 2018, Precise geochronological constraints on the origin, zonation of the Olympic Dam iron oxide Cu-U-Au-Ag deposit, South Aus-
setting and incorporation of ca. 1.59 Ga surficial facies into the Olympic tralia: Society of Economic Geologists, Special Publication 16, p. 237–268.
Dam breccia complex, South Australia: Precambrian Research, v. 315, Fanning, C.M., Reid, A.J., and Teale, G.S., 2007, A geochronological frame-
p. 162–178. work for the Gawler craton, South Australia: Geological Society of South
Chiaradia, M., and Caricchi, L., 2017, Stochastic modelling of deep magmatic Australia Bulletin, v. 55.
controls on porphyry copper deposit endowment: Scientific Reports, v. 7, Groves, D.I., Bierlein, F.P., Meinert, L.D., and Hitzman, M.W., 2010, Iron
article 44523. oxide copper-gold (IOCG) deposits through Earth history: Implications for
Chiaradia, M., Schaltegger, U., Spikings, R., Wotzlaw, J.-F., and Ovtcharova, origin, lithospheric setting, and distinction from other epigenetic iron oxide
M., 2013, How accurately can we date the duration of magmatic-hydrother- deposits: Economic Geology, v. 105, p. 641–654.
mal events in porphyry systems?—an invited paper: Economic Geology, v. Groves, D.I., Zhang, L., and Santosh, M., in press, Subduction, mantle meta-
108, p. 565–584. somatism, and gold: A dynamic and genetic conjunction: Geological Society
Chiaradia, M., Schaltegger, U., and Spikings, R.A., 2014, Time scales of min- of America Bulletin, doi: 10.1130/B35379.1.
eral systems—advances in understanding over the past decade: Society of Haynes, D.W., Cross, K.C., Bills, R.T., and Reed, M.H., 1995, Olympic Dam
Economic Geologists, Special Publication 18, p. 37–58. ore genesis; a fluid-mixing model: Economic Geology, v. 90, p. 281–307.
Ciobanu, C.L., Wade, B.P., Cook, N.J., Schmidt Mumm, A., and Giles, D., Hayward, N., and Skirrow, R.G., 2010, Geodynamic setting and controls on
2013, Uranium-bearing hematite from the Olympic Dam Cu-U-Au deposit, iron oxide Cu-Au (±U) ore in the Gawler craton, South Australia, in Por-
South Australia: A geochemical tracer and reconnaissance Pb-Pb geochro- ter, T.M, ed., Hydrothermal iron oxide copper-gold and related deposits: A
nometer: Precambrian Research, v. 238, p. 129–147. global perspective, v. 3: Adelaide, PGC Publishing, p. 1–27.
Ciobanu, C.L., Cook, N.J., and Ehrig, K., 2017, Ore minerals down to Hiess, J., Condon, D.J., McLean, N., and Noble, S.R., 2012, 238U/235U system-
the nanoscale: Cu-(Fe)-sulphides from the iron oxide copper gold atics in terrestrial uranium-bearing minerals: Science, v. 335, p. 1610–1614.
deposit at Olympic Dam, South Australia: Ore Geology Reviews, v. 81, Holtz, F., Johannes, W., Tamic, N., and Behrens, H., 2001, Maximum and
p. 1218–1235. minimum water contents of granitic melts generated in the crust: A reevalu-
Clark, J.M., Poznik, N., Ehrig, K., Cherry, A., McPhie, J., and Kamenetsky, ation and implications: Lithos, v. 56, p. 1–14.
V.S., 2018, Syn- to post-mineralization structural dismemberment of the Huang, Q., Kamenetsky, V.S., McPhie, J., Ehrig, K., Meffre, S., Maas, R.,
Olympic Dam Fe-Cu-U-Au-Ag deposit, Gawler craton: Society of Eco- Thompson, J., Kamenetsky, M., Chambefort, I., Apukhtina, O., and Hu,
nomic Geologists, SEG 2018: Metals, Minerals, and Society, Keystone, Y., 2015, Neoproterozoic (ca. 820–830 Ma) mafic dykes at Olympic Dam,
Colorado, September 22–25, 2018, Conference Abstract. South Australia: Links with the Gairdner large igneous province: Precam-
Clemens, J.D., 1984, Water contents of silicic to intermediate magmas: brian Research, v. 271, p. 160–172.
Lithos, v. 17, p. 273–287. Huang, Q., Kamenetsky, V.S., Ehrig, K., McPhie, J., Kamenetsky, M., Cross,
Cline, J.S., and Bodnar, R.J., 1991, Can economic porphyry copper mineral- K., Meffre, S., Agangi, A., Chambefort, I., Direen, N.G., Maas, R., and
ization be generated by a typical calc-alkaline melt?: Journal of Geophysical Apukhtina, O., 2016, Olivine-phyric basalt in the Mesoproterozoic Gawler
Research, v. 96, p. 8113–8126. silicic large igneous province, South Australia: Examples at the Olympic
Courtney-Davies, L., Zhu, Z., Ciobanu, C.L., Wade, B.P., Cook, N.J., Ehrig, Dam iron oxide Cu-U-Au-Ag deposit and other localities: Precambrian
K., Cabral, A.R., and Kennedy, A., 2016, Matrix-matched iron-oxide laser Research, v. 281, p. 185–199.
ablation ICP-MS U-Pb geochronology using mixed solution standards: Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bentley, W.C., and Essling, A.M.,
Minerals, v. 6, article 85. 1971, Precision measurement of half-lives and specific activities of 235U and
Courtney-Davies, L., Tapster, S.R., Ciobanu, C.L., Cook, N.J., Verdugo-Ihl, 238U: Physical Review C, v. 4, p. 1889–1906.

M.R., Ehrig, K.J., Kennedy, A.K., Gilbert, S.E., Condon, D.J., and Wade, Jagodzinski, E.A., 2005, Compilation of SHRIMP U-Pb geochronological
B.P., 2019a, A multi-technique evaluation of hydrothermal hematite U-Pb data Olympic domain, Gawler craton, South Australia, 2001–2003: Geosci-
isotope systematics: Implications for ore deposit geochronology: Chemical ence Australia, Record 2005/020.
Geology, v. 513, p. 54–72. Johnson, J.P., 1993, The geochronology and radiogenic isotope systematics
Courtney-Davies, L., Ciobanu, C.L., Verdugo-Ihl, M.R., Dmitrijeva, M., of the Olympic Dam copper-uranium-gold-silver deposit, South Australia:
Cook, N.J., Ehrig, K., and Wade, B.P., 2019b, Hematite geochemistry and Unpublished Ph.D. thesis, Canberra, Australia, The Australian National
geochronology resolve genetic and temporal links among iron-oxide cop- University.
per gold systems, Olympic Dam district, South Australia: Precambrian Johnson, J.P., and Cross, K.C., 1995, U-Pb geochronological constraints on
Research, v. 335, article 105480. the genesis of the Olympic Dam Cu-U-Au-Ag deposit, South Australia:
Courtney-Davies, L., Ciobanu, C.L., Verdugo-Ihl, M.R., Slattery, A., Cook, Economic Geology, v. 90, p. 1046–1063.
N.J., Dmitrijeva, M., Keyser, W., Wade, B.P., Dominick, U.I., Ehrig, K., Keppler, H., and Wyllie, P.J., 1991, Partitioning of Cu, Sn, Mo, W, U, and Th
Xu, J., and Kontonikas-Charos, A., 2019c, Zircon at the nanoscale records between melt and aqueous fluid in the systems haplogranite-H2O-HCl and
metasomatic processes leading to large magmatic-hydrothermal ore depos- haplogranite-H2O-HF: Contributions to Mineralogy and Petrology, v. 109,
its: Minerals, v. 9, article 364. p. 139–150.
Courtney-Davies, L., Ciobanu, C.L., Verdugo-Ihl, M.R., Cook, N.J., Ehrig, Keyser, W., Ciobanu, C.L., Cook, N.J., Dmitrijeva, M., Courtney-Davies,
K., Wade, B.P., Zhu, Z.Y., and Kamenetsky, V., 2020, ∼1760 Ma magnetite- L., Feltus, H., Gilbert, S., Johnson, G., and Ehrig, K., 2019, Iron-oxides
bearing protoliths in the Olympic Dam deposit, South Australia: Implications constrain BIF evolution in terranes with protracted geological histories:

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas
16 SCIENTIFIC COMMUNICATIONS

The Iron Count prospect, Middleback Ranges, South Australia: Lithos, emplacement rates and porphyry copper deposits: Thermal modeling of the
v. 324–325, p. 20–38. Yerington batholith, Nevada. Economic Geology, v. 112, p. 1653–1672
Kirchenbaur, M., Maas, R., Ehrig, K., Kamenetsky, V.S., Strub, E., Ballhaus, Sillitoe, R.H., and Mortensen, J.K., 2010, Longevity of porphyry copper for-
C., and Münker, C., 2016, Uranium and Sm isotope studies of the super- mation at Quellaveco, Peru: Economic Geology, v. 105, p. 1157–1162.
giant Olympic Dam Cu-Au-U-Ag deposit, South Australia: Geochimica et Skirrow, R.G., Bastrakov, E.N., Barovich, K., Fraser, G.L., Creaser, R.A.,
Cosmochimica Acta, v. 180, p. 15–32. Fanning, C.M., Raymond, O.L., and Davidson, G.J., 2007, Timing of iron
Kontonikas-Charos, A., Ciobanu, C.L., Cook, N.J., Ehrig, K., Krneta, S., and oxide Cu-Au-(U) hydrothermal activity and Nd isotope constraints on metal
Kamenetsky, V.S., 2017, Feldspar evolution in the Roxby Downs Granite, sources in the Gawler craton, South Australia: Economic Geology, v. 102,
host to Fe-oxide Cu-Au-(U) mineralisation at Olympic Dam, South Austra- p. 1441–1470.
lia: Ore Geology Reviews, v. 80, p. 838–859. Skirrow, R.G., Van Der Wielen, S.E., Champion, D.C., Czarnota, K., and
Kozlov, V.D., 2011, Trace-element composition and origin of granitoids from Thiel, S., 2018, Lithospheric architecture and mantle metasomatism linked
the Shakhtama complex and Kukul’bei rare-metal complex (Aga zone, to iron oxide Cu‐Au ore formation: Multidisciplinary evidence from the
Transbaikalia): Russian Geology and Geophysics, v. 52, no. 5, p. 526–536. Olympic Dam region, South Australia: Geochemistry, Geophysics, Geosys-
Krneta, S., Ciobanu, C.L., Cook, N.J., Ehrig, K., and Kontonikas-Charos, A., tems, v. 19, p. 2673–2705.
2016, Apatite at Olympic Dam, South Australia: A petrogenetic tool: Lithos, Steiger, R.H., and Jäger, E., 1977, Subcommission on geochronology: Con-
v. 262, p. 470–485. vention on the use of decay constants in geo- and cosmochronology: Earth
Krneta, S., Ciobanu, C.L., Cook, N.J., Ehrig, K., and Kontonikas-Charos, A., and Planetary Science Letters, v. 36, p. 359–362.
2017, Rare earth element behaviour in apatite from the Olympic Dam Cu- Verdugo-Ihl, M.R., Ciobanu, C.L., Cook, N.J., Ehrig, K., Courtney-Davies,
U-Au-Ag deposit, South Australia: Minerals, v. 7, doi: 10.3390/min7080135. L., and Gilbert, S., 2017, Textures and U-W-Sn-Mo signatures in hematite
Krneta, S., Ciobanu, C., Cook, N., and Ehrig, K., 2018, Numerical model- from the Olympic Dam Cu-U-Au-Ag deposit, South Australia: Defining the
ing of REE fractionation patterns in fluorapatite from the Olympic Dam archetype for IOCG deposits: Ore Geology Reviews, v. 91, p. 173–195.
deposit (South Australia): Minerals, v. 8, article 342. Verdugo-Ihl, M.R., Ciobanu, C.L., Cook, N.J., Ehrig, K., and Courtney-
Li, Y., Selby, D., Condon, D., and Tapster, S., 2017, Cyclic magmatic-hydro- Davies, L., 2020, Defining early stages of IOCG systems: Evidence from
thermal evolution in porphyry systems: High-precision U-Pb and Re-Os iron-oxides in the outer shell of the Olympic Dam deposit, South Australia:
geochronology constraints on the Tibetan Qulong porphyry Cu-Mo deposit: Mineralium Deposita, v. 55, p. 429–452.
Economic Geology, v. 112, p. 1419–1440. Vigneresse, J.L., and Truche, L., 2020, Modeling ore generation in a mag-
Ludwig, K.R.A., 2012, User’s manual for Isoplot 3.75: A geochronological matic context: Ore Geology Reviews, v. 116, article 103223.
toolkit for Microsoft Excel: Berkeley Geochronology Centre Special Pub- von Quadt, A., Erni, M., Martinek, K., Moll, M., Peytcheva, and Heinrich,
lication, v. 5. C.A., 2011, Zircon crystallization and the lifetimes of ore-forming mag-
Maas, R., Kamenetsky, V.S., Ehrig, K., Meffre, S., McPhie, J., and Diemar, matic-hydrothermal systems: Geology, v. 39, p. 731−734.
G., 2011, Olympic Dam U-Cu-Au deposit, Australia: New age constraints: Wade, C.E., Payne, J.L., Barovich, K.M., and Reid, A.J., 2019, Heterogeneity
Mineralogical Magazine, Goldschmidt Conference abstract, v. 75, p. 1375. of the sub-continental lithospheric mantle and “non-juvenile” mantle addi-
Mattinson, J.M., 2005, Zircon U-Pb chemical abrasion (“CA-TIMS”) method: tions to a Proterozoic silicic large igneous: Lithos, v. 340–341, p. 87–107.
Combined annealing and multi-step partial dissolution analysis for improved Walter, B.F., Gerdes, A., Kleinhanns, I.C., Dunkl, I., Von, H., Kreissl, S., and
precision and accuracy of zircon ages: Chemical Geology, v. 220, p. 47–66. Markl, G., 2018, The connection between hydrothermal fluids, mineraliza-
McPhie, J., Ehrig, K., Kamenetsky, M.B., Crowley, J.L., and Kamenetsky, tion, tectonics and magmatism in a continental rift setting: Fluorite Sm-Nd
V.S., 2020, Geology of the Acropolis prospect, South Australia, constrained and hematite and carbonates U-Pb geochronology from the Rhine graben
by high-precision CA-TIMS ages: Australian Journal of Earth Sciences, v. in SW Germany: Geochimica et Cosmochimica Acta, v. 240, p. 11–42.
67, p. 699–716. Weatherley, D.K., and Henley, R.W., 2013, Flash vaporization during earth-
Meffre, S., Ehrig, K., Kamenetsky, V.S., Chambefort, I., Maas, R., and quakes evidenced by gold deposits: Nature Geoscience, v. 6, p. 294–298.
McPhie, J., 2010, Pb isotopes at Olympic Dam: Constraining sulphide Zhou, H., Sun, X., Wu, Z., Liao, J., Fu, Y., Li, D., Hollings, P., Liu, Y., Lin, H.,
growth: Giant ore deposits down under: Quadrennial International Associa- and Lin, Z., 2017, Hematite U-Pb geochronometer: Insights from monazite
tion on the Genesis of Ore Deposits (IAGOD) Symposium, 13th, Adelaide, and hematite integrated chronology of the Yaoan gold deposit, Southwest
South Australia, 2010, Proceedings, p. 78–79. China: Economic Geology, v. 112, p. 2023–2039.
Oreskes, N., and Einaudi, M.T., 1990, Origin of rare earth element-enriched Zimmerman, A., Stein, H.J., Morgan, J.W., Markey, R.J., and Watanabe, Y.,
hematite breccias at the Olympic Dam Cu-U-Au-Ag deposit, Roxby Downs, 2014, Re-Os geochronology of the El Salvador porphyry Cu-Mo deposit,
South Australia: Economic Geology, v. 85, p. 1–28. Chile: Tracking analytical improvements in accuracy and precision over the
——1992, Origin of hydrothermal fluids at Olympic Dam; preliminary results past decade: Geochimica et Cosmochimica Acta, v. 131, p. 13–32.
from fluid inclusions and stable isotopes: Economic Geology, v. 87, p. 64–90.
Paton, C., Hellstrom, J., Paul, P., Woodhead, J., and Hergt, J., 2011, Iolite:
Freeware for the visualisation and processing of mass spectrometric data:
Journal of Analytical Atomic Spectrometry, v. 26, p. 2508–2518.
Reeve, J.S., Cross, K.C., Smith, R.N., and Oreskes, N., 1990, Olympic Dam
copper-uranium-gold-silver deposit Australasian Institute of Mining and
Metallurgy (AusIMM), Monograph 14, p. 1009–1035.
Reid, A., 2019, The Olympic Cu-Au province, Gawler craton: A review of
the lithospheric architecture, geodynamic setting, alteration systems, cover
successions and prospectivity: Minerals, v. 9, article 371.
Richards, J.P., 2018, A shake-up in the porphyry world?: Economic Geology, Liam Courtney-Davies received an M.Sci.
v. 113, no. 6, p. 1225–1233. degree (2015) from Royal Holloway, University
Schlegel, T.U., Wagner, T., Boyce, A., and Heinrich, C.A., 2017, A magmatic of London (UK) and a Ph.D. degree (2020) from
source of hydrothermal sulfur for the Prominent Hill deposit and associated
the University of Adelaide (Australia). His doctoral
prospects in the Olympic iron oxide copper‐gold (IOCG) province of South
Australia: Ore Geology Reviews, v. 89, p. 1058–1090. research within the FOX project, led by Dr. Cris-
Schmitz, M.D., and Schoene, B., 2007, Derivation of isotope ratios, errors, and tiana Ciobanu, focused on radioisotopic dating of
error correlations for U-Pb geochronology using 205Pb-235U-(233U)-spiked IOCG systems, development and application of
isotope dilution thermal ionization mass spectrometric data: Geochemistry, iron-oxide U-Pb geochronology, and synthesis and characterization of U-Pb
Geophysics, Geosystems, v. 8, Q08006, doi: 10.1029/2006GC001492. reference materials. His current postdoctoral research addresses nano- to
Schöpa, A., Annen, C., Dilles, J.H., Sparks, S., and Blundy, J.D., 2017, Magma atom-scale modification of zircon associated with ore deposits.

Downloaded from https://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4772/5140315/4772_courtney-davies+et+al.pdf


by Juan Rojas

You might also like