You are on page 1of 17

RESEARCH ARTICLE Proxies for Basement Structure and Its Implications

10.1029/2018JB016829
for Mesoproterozoic Metallogenic Provinces
Key Points:
• The satellite gravity data provide
in the Gawler Craton
clues to the regional framework of
the Gawler Craton, which aligns
J. G. Motta1 , P. G. Betts2 , C. R. de Souza Filho1, S. Thiel3 , S. Curtis3 , and R. J. Armit2
with different geophysical data sets 1
• The Gawler Craton is formed by a
Institute of Geosciences, University of Campinas, Campinas, Brazil, 2School of Earth, Atmosphere and Environment,
threefold Archean substratum that Monash University, Clayton, Victoria, Australia, 3Department for Energy and Mining, Geological Survey of South
has undergone changes in the Australia, Adelaide, South Australia, Australia
Proterozoic down to its deep crustal
levels
• Regional structures were reactivated Abstract The link between mineral resources and crustal‐rooted structures has been proposed for many
through time and act as pathways
for regional fluid flow throughout of the world's most significant mineral provinces. Here we utilize a new approach by interpreting potential
the craton field data, including satellite gravity data, and high‐resolution continental‐scale magnetotelluric data,
constrained with aeromagnetic, and seismic tomography and reflection data, to determine the distribution of
Supporting Information: crustal‐scale faults in the Archean to Proterozoic Gawler Craton (South Australia). The eastern flank of the
• Supporting Information S1
craton hosts the supergiant Olympic Dam iron oxide‐copper‐gold (IOCG) deposit within a larger Olympic
IOCG province. The central part of the craton contains gold‐only deposits, which define the Central Gawler
Correspondence to: Gold province. Both of these provinces are part of a Mesoproterozoic mineral system with an extensive
J. G. Motta, hydrothermal alteration footprint, which formed during complicated tectonic mode switches. We show that
jgmotta@gmail.com both types of mineralization are located in proximity to crustal‐scale structures that appear to connect
deep crustal fragments, which likely record the amalgamation of the Archean nucleus of the craton during
Citation: the Neoarchean with subsequent reworking during the Mesoproterozoic. Many of these structures do not
Motta, J. G., Betts, P. G., de Souza Filho, have a surface expression but coincide with gradients in magnetism, gravity, and electric resistivity
C. R., Thiel, S., Curtis, S., & Armit, R. J.
(2019). Proxies for basement structure anomalies, the latter data set suggesting they acted as fluid pathways extending to the lower crust. The
and its implications for results indicate that the first‐order controls on the distribution of IOCG and Central Gawler Gold
Mesoproterozoic metallogenic metallogenic provinces are inherited from earlier tectonic events, which formed major crustal boundaries
provinces in the Gawler Craton.
Journal of Geophysical Research: Solid and related structures that are prone to reworking during later tectonism.
Earth, 124, 3088–3104. https://doi.org/
10.1029/2018JB016829
1. Introduction
Received 4 OCT 2018
Accepted 8 FEB 2019 Large mineral systems punctuate Earth's evolution and are related to major transient tectonic disturbances
Accepted article online 14 FEB 2019 that affect the whole lithosphere (Hagemann et al., 2016; Richards & Mumin, 2013; Rosenbaum et al., 2005;
Published online 7 MAR 2019
Sillitoe, 2003; Tassara et al., 2017). Most large mineral systems require the existence of geological structures
(e.g., sets of faults, shear zones, and folds) that provide physical pathways for crucial large‐scale fluid
plumbing through the lithosphere to the crust (Huston et al., 2016). However, the identification of these
structures is difficult because rocks are often reworked, structures are reactivated, and deformation
superimposed over time or is buried by younger rocks (Heron et al., 2016). Furthermore, the task of
distinguishing between shallow and deep penetrating structures is challenging due to data resolution, the
ambiguity of geophysical techniques, and the physical properties of geological materials (Huston et al.,
2016; Korsch & Doublier, 2016).
Satellite‐derived measurements of the Earth's gravity field can be used to observe density variations within
the planet (Barzaghi et al., 2015; Ebbing et al., 2013). Applications of these data for evaluation of the
lithosphere have concentrated on the mantle (e.g., density and stress state and determination of the
Mohorovic discontinuity; Bouman et al., 2016; Panet et al., 2014; Uieda & Barbosa, 2017; Van der Meijde
©2019. The Authors. et al., 2013), with much less emphasis on the crustal structures (Bouman et al., 2015; Braitenberg, 2015;
This is an open access article under the Ebbing et al., 2018; Van der Meijde et al., 2015). Crustal studies have shown distinctive characteristics in
terms of the Creative Commons
Attribution‐NonCommercial‐NoDerivs the framework of the African and South American continents (Braitenberg, 2015) with data from the
License, which permits use and distri- Gravity Field and Steady‐State Ocean Circulation mission. However, the data have no sufficient spatial
bution in any medium, provided the resolution to evolve into the study of the small‐scale crustal structure. Some relationships between features
original work is properly cited, the use
is non‐commercial and no modifica- in the gravity data and known orogenic gold mineralization within greenstone belts in Zimbabwe were
tions or adaptations are made. identified (Braitenberg, 2015). Nevertheless, the interpretation of these data to understand the structure of

MOTTA ET AL. 3088


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Figure 1. (a) Simplified basement geology and (b) shear zones and existent sedimentary cover within the Gawler Craton, with outlines of the Olympic and Central
Gawler provinces after Reid and Fabris (2015). Acronyms: mineral deposits, PH, Prominent Hill; OD, Olympic Dam; CA, Carrapateena; TA, Tarcoola; shear
zones, KjMZ, Kalinjala Mylonite Zone; KSZ, Karari; ISZ, Ifould; TSZ, Tallacootra; CSZ, Coorabie; BSZ, Bulgunnia; YeSZ, Yerda; KoSZ, Kooniba; YaSZ, Yarlbrinda.
The interpreted extent of the Mesoarchean terrane in the southern Gawler Craton is as in Curtis and Thiel (2019).

the Earth's crust and geological provinces that host major mineral deposits remains underutilized. This
study presents a novel approach to address the relationship between crustal architecture and mineral
systems.
We have developed a study case on the application of satellite gravity data over the Gawler Craton in South
Australia (Figure 1), which hosts the Olympic iron oxide‐copper‐gold (IOCG; Bastrakov et al., 2007; Skirrow
et al., 2007) and the Central Gawler Gold (CGG; Ferris & Schwarz, 2003; Fraser et al., 2007) metallogenic
provinces. Both provinces formed during the earliest Mesoproterozoic (approximately 1,600 to 1,570 Ma;
Skirrow et al., 2007; Fraser et al., 2007) over an Archean‐Palaeoproterozoic cratonic shield (Hand et al.,
2007). Models for the genesis and geodynamic setting of these provinces are varied, including plume‐
modified orogenesis and magmatism (Betts et al., 2007, 2009), the development of extensional sedimentary
basins (Cherry et al., 2017; McPhie, Kamenetsky, Chambefort, et al., 2011; McPhie et al., 2016), a convergent
plate‐boundary settings (Direen & Lyons, 2007; Fraser et al., 2007), and postsubduction settings (Skirrow
et al., 2018). Despite significant differences between these models, most suggest that the craton‐scale mag-
matism related to the emplacement of the early Mesoproterozoic Hiltaba Suite plutons and coeval Gawler
Range Volcanics (Allen et al., 2008; Betts et al., 2009) supplied the thermal energy that drove crustal‐scale
fluids and associated mineralization (Bastrakov et al., 2007; Fraser et al., 2007; Kontonikas‐Charos et al.,
2017; Oreskes & Einaudi, 1992; Skirrow et al., 2007).
Magnetic, seismic reflection and magnetotelluric data have provided information about the near‐surface
controls on mineralization throughout the Gawler Craton (e.g., Direen & Lyons, 2007; Gow et al.,
1994). Several crustal‐scale fluid pathways responsible for fluid transport through the craton's basement
and cover sequences have been identified in regional seismic reflection data (Betts et al., 2016,
Drummond et al., 2006; Wise et al., 2015, see Figure S1 in the supporting information), deep imaging
magnetotelluric data (Curtis & Thiel, 2019; Heinson et al., 2006, 2018; Thiel et al., 2005; Thiel et al.,
2016; Thiel & Heinson, 2010; Thiel & Heinson, 2013), and potential field data (Baines et al., 2011; Betts
et al., 2003; Direen & Lyons, 2007; Stewart & Betts, 2010a, 2010b). The challenging question of how the
deep crustal penetrating structures connect with the upper crustal structures and their relationship with
large metallogenic provinces has been addressed with different traditional geophysical methods.
However, an integrative model of the regional architecture of the deep crust of the Gawler Craton
is lacking.

MOTTA ET AL. 3089


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

In this study, we present a novel application of satellite gravity data combined with magnetotelluric and
magnetic data to identify crustal‐scale structures spatially associated with the IOCG and the CGG metallo-
genic provinces on the Gawler Craton (Figure 1). The geometry of the concealed Archean basement of the
Gawler Craton is presented, where the presence of crustal breaks within different geological blocks controls
the location of the metallogenic provinces regionally. We discuss our geophysical interpretation in the
context of existing geochronology data.

2. Geologic Setting
The Gawler Craton (Figure 1) records a complex geologic evolution that spans from the Archean eon to the
Mesoproterozoic era. The oldest rocks include the Mesoarchean Cooyerdoo Granite (>3.1 Ga; Fraser,
McAvaney, et al., 2010) and the Neoarchean to Early Palaeoproterozoic Mulgathing and Sleaford complexes
(Hand et al., 2007; Reid & Hand, 2012; Figure 1). Both complexes have mixed meta‐sedimentary (aluminous
and chemical sediments) and meta‐igneous rocks (volcanics ranging from ultramafic to felsic), and
abundant felsic magmatism with interpreted arc affinities (Swain, Hand, et al., 2005; Swain, Woodhouse,
et al., 2005; Hand et al., 2007). These complexes amalgamated during the Sleafordian Orogeny
(2,480–2,420 Ma; Hand et al., 2007; Swain, Hand, et al., 2005; Swain, Woodhouse, et al., 2005) and preserve
varying metamorphic conditions from low‐grade greenschist to granulite facies. The Cooyerdoo Granite
and associated rocks are predominantly ortho‐gneisses and granitoids enclosed in middle‐Proterozoic
meta‐volcano‐sedimentary sequences (Hand et al., 2007). Isotope geochemistry results (Nd and Hf systems)
by Curtis and Thiel (2019) on whole‐rock samples and magmatic and inherited zircons point to the existence
of an extended Mesoarchean terrane that extends further west than the exposed Cooyerdoo Granite and
underlies Neoarchean to Palaeoproterozoic crust (Figure 1a).
The Archean‐Palaeoproterozoic basement is overlain by several, variably metamorphosed volcano‐
sedimentary basins (Figure 1a), including the approximately 1,750 Ma Wallaroo Group in the eastern parts
of the craton and approximately 2,000 to 1,740 Ma sedimentary sequences in the south (Dutch et al., 2008;
Hand et al., 2007; Szpunar et al., 2011), and mostly covered by the approximately 1,790 to 1,740 Ma meta‐
sedimentary successions and igneous rocks in the north and west. Generally, the north and west provinces
are characterized by higher metamorphic grades compared to equivalent rocks in the south (Armit et al.,
2017; Betts et al., 2003; Direen et al., 2005; Dutch et al., 2008; Hand et al., 2007; Reid, Jagodzinski, Fraser,
et al., 2014; Reid, Jagodzinski, Armit, et al., 2014). The Fowler domain to the west comprises metapelitic
and meta‐igneous rocks with similar evolution to the sequences lying to the north (Hand et al., 2007;
Figure 1a). These basins were formed in a continental back‐arc setting following the amalgamation of the
Gawler Craton to the remainder of Australia during the Nuna supercontinent accretion (Armit et al.,
2017; Betts et al., 2016, 2008; Hand et al., 2007; Reid, Jagodzinski, Fraser, et al., 2014, Reid, Jagodzinski,
Armit, et al., 2014; Reid & Payne, 2017).
The Gawler Craton underwent several episodes of orogenic reworking during the Cornian (approximately
1,850 Ma), Kimban (approximately 1,730–1,690 Ma), Wartakan (approximately 1,630–1,610 Ma), and
Kararan (approximately 1,600–1,570 Ma) orogenies (Betts & Giles, 2006; Hand et al., 2007; Stewart &
Betts, 2010b), which led to development and reactivation of large crustal‐scale shear zones and related faults
(Figure 1b). During these events, the Archean and Proterozoic sequences were juxtaposed along arcuate
polyphase shear zones that trend NE‐SW dominantly in the western, E‐W in the central, NW‐SE to NE‐
SW in the northern, and N‐S to NW‐SE in the central to eastern parts of the craton (Hand et al., 2007;
Figure 1b). These shear zones are overprinted by a series of Mesoproterozoic E‐W to NE‐SW trending trans-
current faults, which were active during the Hiltaba Suite magmatism (Betts et al., 2009). These orogenic
events were succeeded by the onset of back‐arc granitic magmatism, which includes the approximately
1,850‐Ma Donington Suite and the approximately 1,690‐ to 1,670‐Ma Tunkillia Suite (Payne et al., 2010;
Reid et al., 2009). The approximately 1,650‐ to 1,610‐Ma St. Peter Suite preserves arc‐related granitoids
(Swain et al., 2008), which are subsequently succeeded by the approximately 1,600‐ to 1,575‐Ma within‐plate,
plume‐related felsic magmatism of the Hiltaba Suite granitoids and the predominantly felsic Gawler Range
Volcanics (GRV;approximately 1,600–1,575 Ma; Allen et al., 2008; Betts & Giles, 2006; Hand et al., 2007;
Pankhurst et al., 2011; Reid & Payne, 2017). During the Middle Mesoproterozoic sedimentary infilling in a
stable cratonic setting is characterized by the deposition of the Pandurra Formation (Figure 1a), which is

MOTTA ET AL. 3090


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

followed by the deposition of the sedimentary and volcanics rocks of the Adelaide Rift Complex during the
Neoproterozoic and further, widespread sedimentation during the early Paleozoic (Hand et al., 2007;
Figure 1b).
Seismic tomography data of the Gawler Craton shows a Moho discontinuity that varies in depth from 30 km
in the south‐east to 54 km in the north (Salmon et al., 2013; Sippl, 2016). Estimates of the Vp/Vs seismic wave
ratios suggest that the composition of the crust in the craton varies from felsic along its central to northern
parts to more mafic along the south (Sippl, 2016). The Moho surface is interpreted on deep seismic reflection
data (Kennett & Saygin, 2015) as a relatively flat surface varying from 40 to 45 km along the central and east-
ern parts of the craton (profiles 08GA‐G1 and 03GA‐OD1; Figures S1b and S1c; Fraser & Blewett, 2010; Wise
et al., 2015). However, the western part of the craton, imaged in the 13GA‐G1 seismic section (Doublier et al.,
2015, Figure S1a), shows the Moho changes from a flat and smooth morphology to an undulating surface
that reaches up to 50 km below the approximately 1,790‐ to 1,740‐Ma metasedimentary rocks of the northern
parts of the craton. Seismic reflection data support a layered crust in the Gawler Craton, which is given by
preponderant westward dipping structures to the west that change to eastward dipping in the east. These
structures have listric geometries (Figure S1) and are interpreted to be regional‐scale shear zones
(Doublier et al., 2015; Fraser & Blewett, 2010; Wise et al., 2015). These structures dissect the crust to deep
crustal levels, where they usually join zones of bland seismic character and merge into horizontal reflectors
parallel to the Moho (Doublier et al., 2015; Fraser & Blewett, 2010; Figure S1).
Electrical conductivity data suggests that the Gawler Craton is mostly composed of highly resistive
(>750 Ohm.m) anhydrous crust, interpreted to be Archean‐aged, which is truncated by several conductive
zones (<20 Ohm.m) that often coincide with regional shear zones extending into the upper mantle (Curtis
& Thiel, 2019; Heinson et al., 2006, 2018; Thiel et al., 2005; Thiel & Heinson, 2010; Thiel & Heinson,
2013). Beneath the Olympic Dam deposit (Figure 1), a prominent electrical discontinuity in the middle‐crust
to lower crust exists, with the latter becoming more conductive (<100 Ohm.m), which is possibly related to
interstitial graphite liberalization along structures fertilized by mantle fluids (Heinson et al., 2006) or sul-
fides (Heinson et al., 2018). Interpretations on the mantle architecture combined from seismology, electrical
resistivity, and geochemistry data (Skirrow et al., 2018) suggest that the IOCG and CGG provinces occur
along the boundary between the Archean/Proterozoic‐aged metasomatized subcontinental lithospheric
mantle and evolved, residual eclogitic crust derived from east directed crust consumption throughout the
approximately 1,730‐ to 1,690‐Ma Kimban Orogeny. Mantle metasomatism is interpreted to occur during
foundering of the lowermost lithospheric mantle at approximately 1,595 Ma, replacing the old, refractory
cratonic mantle root and facilitating magma and fluid ingress into the crust at the time of mineralization
(Skirrow et al., 2018). Continental‐scale results from Chopping and Kennett (2015) on the maximum depth
of the magnetization in the Gawler Craton vary from 40 to 55 km, with a homogenous distribution.
There is little consensus about the source of hydrothermal fluids associated with the IOCG Province, with
genetic models proposing a crustal source (Cherry et al., 2017; McPhie, Kamenetsky, Allen, et al., 2011;
McPhie, Kamenetsky, Chambefort, et al., 2011; McPhie et al., 2016), mantle/magmatic source (Creaser &
Cooper, 1993; Ciobanu et al., 2013; Heinson et al., 2018; Johnson & McCulloch, 1995; Kirchenbaur et al.,
2016; Kontonikas‐Charos et al., 2017; McPhie, Kamenetsky, Allen, et al., 2011; Schlegel et al., 2016), and
mixing of fluids from different sources (Bastrakov et al., 2007; Haynes et al., 1995; Huston et al., 2016;
Skirrow et al., 2007, 2018; Uvarova et al., 2017). The formation of the CGG deposits is not clear, as they
record characteristics of both orogenic gold and intrusion‐related systems (Fraser et al., 2007). Ore deposits
in both metallogenic provinces share relationships to major regional shear zones or subsidiary structures
(Bastrakov et al., 2007;Direen & Lyons, 2007; Skirrow et al., 2007) and are coeval with the emplacement
of the Hiltaba Suite granitoids. Reid and Fabris (2015) argue that differences in the metamorphic grade, com-
position, and porosity of the basement were influential on the hydrothermal system, leading to different
facies and deposition of metals. Structures interpreted from data obtained from potential‐field and electric
methods suggest that deep‐seated shear zone geometries control the formation of the IOCG and CGG pro-
vinces (Direen & Lyons, 2007; Drummond et al., 2006; Heinson et al., 2006, 2018). The collocation of the
Hiltaba Suite plutons, which are hosts to the deposits in the Olympic province (Direen & Lyons, 2007;
Skirrow et al., 2007), has been interpreted to be controlled by regional shear zones (McLean & Betts,
2003). In the vicinity of the Olympic Dam deposit, previous interpretations on deep seismic reflection data
point out to the presence of a representative step in the Moho in the vicinity of the Olympic Dam deposit

MOTTA ET AL. 3091


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Figure 2. Potential field data on the Gawler Craton. (a) Complete Bouguer anomaly. (b) Sediment‐free Complete Bouguer anomaly. (c) Reduced to magnetic pole
magnetic data upward continued to 25 km. (d) Second vertical derivative of the upward continued magnetic data. Moho depth and Vp/Vs ratios on b comes
from Sippl (2016). Mineral deposits: PH, Prominent Hill; OD, Olympic Dam; CA, Carrapateena; TA, Tarcoola.

(Drummond et al., 2006; Figure S1c) and regional joint low‐resistivity and ‐impedance structures (Heinson
et al., 2018; Wise et al., 2015).

3. Data and Methods


Satellite‐borne gravity, airborne magnetics, electrical resistivity, seismic tomography, and reflection data
were interpreted together. Satellite gravity information is derived from the satellite‐only Gravity
Observation Combination 05 model (GOCO05s; Mayer‐Guerr, 2015). The GOCO05s data were processed
from the gravity over the ellipsoid (4,000 m above sea level) truncated at 250/order into the simple
Bouguer anomaly assuming a 2,670‐kg/m3 reference density. The complete Bouguer anomaly (Figure 2a)

MOTTA ET AL. 3092


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Figure 3. (a–d) Composite image from the satellite‐derived, sediment‐free, complete Bouguer anomaly with shading by the second vertical derivative of the total
magnetic intensity data at 25‐km height with the regional observed structural framework (a) and interpreted framework over the data in a (b). Interpreted
structural breaks from the composite geophysical image (c). Ground gravity strings over the Gawler Craton (d). Key for the shear zones features: KjMZ, Kalinjala
Mylonite Zone; KSZ, Karari; ISZ, Ifould; TSZ, Tallacootra; CSZ, Coorabie; BSZ, Bulgunnia; YeSZ, Yerda; KoSZ, Kooniba; YaSZ, Yarlbrinda shear zones.
Mineral deposits: PH, Prominent Hill; OD, Olympic Dam; CA, Carrapateena; TA, Tarcoola.

was derived by reducing the simple Bouguer anomaly from the terrain correction, as computed from forward
modeling of the gravity effect of the rocks contained in the topographic surface by using tesseroids (spherical
prisms). The input for the topography surface height comes from the ETOPO1 model (Amante & Eakins,
2009). The sediment‐free complete Bouguer anomaly (SFCBA; Figure 2b) is obtained by removing the
gravity effect of the existent sedimentary cover as forward modeled with sediment depth data from
CRUST1 model (Laske et al., 2012) from the complete Bouguer anomaly. Forward modeling of 2.5‐D
models used the SFCBA along three profiles (see the location in Figure 3 and data profiles in Figure S2).
The purpose of the forward model was to constrain the basement framework of the craton. For the
construction of the forward models we have used density values from reference work (Telford et al., 1990;
Dentith & Mudge, 2014) that approximates to the composition of the major geologic groups as described

MOTTA ET AL. 3093


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

in the literature and reference values (e.g., lower crust). The densities of the modeled bodies are taken as
constant given the lack of constraining information on the density variations with depth. Furthermore,
given that the models are constructed using long‐wavelength information and relatively coarse spatial reso-
lution, we refrain from modeling short‐wavelength features. To constrain the regional dip of the interpreted
structures in the profiles, we have integrated interpretations of seismic reflection data (Figure S1), and the
presence of ground gravity gradient strings (see below). Information on the sedimentary cover depth in
the area for the forward models comes from depth to basement map produced by the Geological Survey of
South Australia (GSSA; Cowley, 2015).
A total magnetic intensity aeromagnetic map leveled at 80‐m height and gridded to 100‐m cell size was used
as compiled by the GSSA (Dhu, 2011). Aeromagnetic data are corrected with a variable reduction to the mag-
netic pole (RTP) of the International Geomagnetic field in 1990. The RTP data were upward continued
(Henderson & Zietzt, 1949) to 25 km above the Earth's surface to enhance long‐wavelength features
(Figure 2c). A second‐order vertical derivative was applied to the upward continued RTP data to enhance
the edges of the long‐wavelength anomalies (Figure 2d). The gravity and magnetic data are inspected in con-
junction to interpret geophysical regions according to their geophysical response, and changes in the trends
observed in the second vertical derivative of the magnetic data. Visual inspection of the composite image
(Figures 3a and 3b) was used for tracing observable textural breaks, which led to a map of interpreted breaks
in the framework of the craton (Figure 3c). To cross‐check our interpretations, we compare our interpreted
structural breaks to gravity gradient strings (worms) extracted from ground gravity data by the GSSA in 2011
(see Figure 3d). The multiscale edges technique was proposed by Archibald et al. (1999) and Hornby et al.
(1999) and consists in the application of the upward continuation filter (Henderson & Zietzt, 1949) for sev-
eral heights greater than the survey elevation and subsequent calculation of its horizontal gradient and gath-
ering of regions with maximum gradient as lines (strings). The regions of maximum gradient represent
regions of expressive change in the gravity field, which are likely to represent edges of geological bodies,
in this case, in different depth levels. The strings come from the results based on the Australian National
Gravity database processed into the Bouguer anomaly with a 0.5‐min resolution from 2008 as provided by
Geoscience Australia. The gradient strings were produced for upward continued surfaces of the gravity field
ranging from 100 m to 25 km, with the use of the multiscale edge detection Wizard by the INTREPID
Software. In this contribution, we opted to use only worms derived from heights greater than 15 km to
restrict the information to the deep part of the crust. Data and documentation for the gradient strings can
be found on the South Australian Resources Information Geoserver (SARIG) platform.
The 3‐D inversion results for electrical resistivity are derived from the Gawler Craton AusLAMP (Australian
Lithospheric Architecture Magnetotelluric Project) data (Robertson et al., 2016). The long‐period (10–
10,000 s) MT data were collected with a nominal spacing of half degree latitude and longitude (about 55‐
km spacing) across the Gawler Craton and resolve the midcrustal to upper mantle resistivity structure
(Robertson et al., 2016; Thiel et al., 2016).
The geometry of the Moho is constrained from the Australian Seismological Reference Model (Salmon et al.,
2013), determinations of the Moho depth by Sippl (2016) (H‐K stacking, receiver function inversion), and
three deep seismic reflection profiles: 13GA‐EG1 (Doublier et al., 2015), 08GA‐G1 (Fraser & Blewett,
2010), and 03GA‐OD1 (Drummond et al., 2006; Lyons & Goleby, 2005). Seismic reflection profiles also help
constrain interpretations on the regional‐scale structures (Figure S1). That was carried out through inspec-
tion of previous published interpretations and profile images to interpret the major structural trends, such as
the dip of regional reflectors, presence, and shape of form lines, features associated with shear zones and
extents of geological domains. To constrain the bulk composition of the crust in the Gawler Craton, we used
information on the Vp/Vs ratio based on Sippl (2016). The velocity ratio (Poisson ratio) constrains the bulk
composition of the crust with the empirical observation that rocks closer to mafic compositions have higher
ratios (Christensen, 1996).
Isotopic geochemistry data, including zircon magmatic and metamorphic crystallization ages, and whole
rock εNd determinations from Champion (2013) and a compilation from literature (Armit et al., 2017;
Belousova et al., 2009; Budd, 2006; Budd & Skirrow, 2007; Creaser, 1989; Creaser & Cooper, 1993; Fraser,
McAvaney, et al., 2010; Fraser & Neumann, 2010; Fraser & Neumann, 2010; Fraser, McAvaney, et al.,
2010; Fraser & Neumann, 2010; Howard, Hand, Barovich & Belousova, et al., 2011; Howard, Hand,

MOTTA ET AL. 3094


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Barovich, Payne, et al., 2011; Jagodzinski, 2005; Johnson, 1993; Johnson & Cross, 1995; Johnson &
McCulloch, 1995; Payne et al., 2006, 2010; Reid & Hand, 2012; Reid et al., 2009; Schaefer, 1998; Swain,
Hand, et al., 2005; Swain, Woodhouse, et al., 2005; Swain et al., 2008; Skirrow et al., 2007; Szpunar et al.,
2011; Turner et al., 1993, and references therein) served as constraints for the crustal evolution of the craton.

4. Crustal Framework Interpreted From Potential Field Data


We interpreted three distinct geophysical domains (domains A to C; Figure 3b) across the craton from the
potential field data (SFCBA and the magnetic maps; Figure 2). These domains are defined by their values
of the gravity anomaly, and texture and trend directions in the magnetic data (Figures 3a and 3c).
Domain A is a subcircular element with subdued gravity response down to −26 mGal. This domain hosts
discontinuous E‐W to NNE‐SSW directed trends in the magnetic texture (Figures 3b and 3c). The lowermost
values in the gravity anomaly within the domain (marked as 1 in Figure 3a), in the central‐western region,
are coincident with rocks of St. Peter Suite magmatic arc and the Archean Sleaford Complex (Figure 1).
Domain A is bounded to the north by Domain B, which is defined by sinuous gravity anomalies with values
that vary between 20 to 100 mGal (e.g., marked as 2: Figure 3a). Domain B is also characterized by breaks in
the magnetic data that trend NE‐SW to E‐W (Figures 3b and 3c) in the north and N‐S in the south‐eastern
part of the craton (Figure 3c). Domain C, located in the southeastern part of the craton, is characterized
by gravity anomaly values that reach 36 mGal and decrease toward the north (marked as 3 in Figure 3a).
The magnetic texture of Domain C is discontinuous with N‐S to NW‐SE directed anomalies (Figures 3b
and 3c). Domain C extends from the Sleaford Complex in the west to the Cooyerdoo Granite in the east
(Figure 3b), beneath the Gawler Range Volcanics (Figure 1). The high‐density anomalies (up to
102 mGal) in the far south (marked as 4 in Figure 3a) relate to the current continental margin. The structural
trends of the domains observed in the composite geophysical image (Figure 3c) are broadly consistent with
gravity gradient strings for the area (Figure 3d). These similarities are particularly evident within the central
to northern part of the craton, where strings for heights greater than 15 km correlate well with our inter-
preted framework map. Gravity gradient strings are more continuous and present smoother contours when
compared to the structural breaks interpreted from the composite image. This difference may reflect the
greater variation of magnetic susceptibility in rocks in contrast to density. That can be of significance when
it comes to the domain boundaries, as some of them are not necessarily highlighted in the strings but on the
interpreted breaks (e.g., western flank of the Domain C, Figures 3c and 3d). The similarity between the two
maps highlight that the interpreted breaks are potential discontinuities in the structural framework of
the craton.
The observed intricate array of breaks with varying NE‐SW, NW‐SE, and E‐W trends interpreted from the
composite geophysical image (Figure 3c) has a significant regional expression, extending throughout the
whole craton. These structures potentially penetrate the boundaries of the geophysical domains and con-
tinue below younger rocks (e.g., GRV, St Peter Suite; Figure 4, see below). These interpreted structural
breaks are in agreement with major geological trends and crosscut relationships of the craton observed at
the surface (e.g., Bulgunnia, Coorabie, Yerda, and Karari shear zones; Figures 1b and 3a). Structures with
N‐S to NW‐SE and E‐W to NE‐SW trends are interpreted as formed during Kimban and Kararan orogenies,
respectively, and are prominent in the central and northern parts of the craton. Thus, our interpreted frame-
work shows that structures observed from the surface potentially extend into the deeper levels of the
cratonic crust.
The subsurface architecture is modeled following our interpretation of the domains from the SFCBA data
(Figures 2b and 3b) along profiles (Figure 3b) as a set of regional scale wedge‐like blocks, which are bound
by shear systems (Figure 4). From west to east, domain B is juxtaposed against domain C with an undulated
ramp‐like geometry (Figures 4a and 4c). Domain A lies in the hanging wall of domains B and C. These
boundaries are moderate to steeply dipping (Figures 4b and 4c). The northern part of domain B is modeled
as a series of north dipping wedges reaching thicknesses up to 10 km (Figure 4c) overlying domain B and
representing a high‐metamorphic grade belt aging from 1,790–1,740 Ma. The Mesoproterozoic to
Neoproterozoic sedimentary packages (e.g., Pandurra Formation and Adelaide syncline) overlie domain B
and the GRV (PC in Figure 4a and far east in Figure 4b). The GRV occurs as a layer up to 5‐km‐thick covering
most of the central‐eastern Gawler Craton, following the 3‐km thickness estimated by Allen et al. (2008), and

MOTTA ET AL. 3095


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Figure 4. Two and a half dimensional forward models of the Gawler Craton from sediment‐free complete Bouguer anomaly data, according to the section traces on
Figure 3b. Backdrop image is the resistivity profile from 15 km down to the Moho depth. (a) Model 1. (b) Model 3. (c) Model 2. The thick continuous lines
represent modeled block limits; the thin discontinuous lines represents features correlating to significant shear zones, and interpreted breaks, and gravity gradient
3
strings. Key for shear zones as for Figure 3. Densities for the units: Mesoproterozoic‐Neoproterozoic sedimentary cover (PC)—2,550 kg/m , St Peter Suite—
3 3 3 3 3
2,600 kg/m , Gawler Range Volcanics—2,650 kg/m , Domains A and B—2,800 kg/m , domain C—2,750 kg/m , Paleoproterozoic high‐grade belts—3,000 kg/m ,
3
and mantle as 3,300 kg/m . Key for the shear zones features: KjMZ, Kalinjala Mylonite Zone; KSZ, Karari; ISZ, Ifould; TSZ, Tallacootra; CSZ, Coorabie; BSZ,
Bulgunnia; YeSZ, Yerda; KoSZ, Kooniba; YaSZ, Yarlbrinda shear zones. See Figure S3 for data profiles.

tentatively adjusted to fit the observed gravity (Figure 4). The crustal thicknesses are similar among the
blocks and range from 35 to 40 km, with over‐thickened areas in the central part of the craton covered by
domains A and C (Figure 4c), following geophysical constraints presented in Drummond et al. (2006),
Huston et al. (2016), and Sippl (2016; see Figure 2b).

5. Evolution Constraints
The crustal evolution of the geophysical domains was assessed using pub-
lished isotope geochronology data (Figure 5; see Figure S3 for sample loca-
tion). The crustal evolution diagram (Figure 5) shows that the three
domains present common trends in their evolution, with shared paths of
evolution from juvenile to evolved crust, with little juvenile material addi-
tion, apart from the St Peter Suite arc (Figure 5; Hand et al., 2007; Reid,
Jagodzinski, Armit, et al., 2014). From 2,500 Ma onward, the three
domains share several reworking events leading to the formation of rocks
with mildly (near zero) to strongly negative εNd values (< −5.0; Figure 5).
Juvenile rocks younger than 2,000 Ma are restricted to domains A and B
(Figure 5), including those related to rocks of the Hiltaba suite magma-
tism and the Wallaroo and 2,000‐ to 1,740‐Ma meta‐sedimentary and
igneous rocks to the south (Figure 1a). In contrast, domain C is the only
domain that contains crystallization ages earlier than the Sleafordian oro-
geny (e.g., >2.8 Ga, up to 3,100 Ma), which are related to slightly negative
to positive εNd values in the Cooyerdoo Granite. Thus, the age constraints
Figure 5. Geochronology data for the three major geophysical domains in a
suggest that the interpreted geophysical domains represent crustal
crustal evolution diagram (n = 224 samples). Curves for the depleted mantle
model and chondritic unfractionated reservoir (CHUR) follow Michard et al. domains with slightly different evolution trajectories before the
(1985) and Goldstein et al. (1984), respectively. Data come from literature Sleafordian Orogeny (Figure 5). The contemporary reworking at approxi-
compilation, see section 3 for references. mately 2,500 Ma indicates that they were amalgamated by that time and

MOTTA ET AL. 3096


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

have a shared evolution into the Paleoproterozoic to Mesoproterozoic within the Mulgathing and Sleaford
complexes and Paleoproterozoic belts. These trends align with crustal growth patterns recorded in in situ zir-
con Lu‐Hf isotope data sets (Reid & Payne, 2017) reflecting reworking within the Nuna Supercontinent
(Betts et al., 2016).

6. Crustal Breaks as Proxies for Regional Fluid Flow


Application of electrical resistivity data has provided constraints on fluid transfer systems throughout the
crust in the Gawler Craton (Curtis & Thiel, 2019; Heinson et al., 2006, 2018; Skirrow et al., 2018; Thiel
et al., 2005, 2016; Thiel & Heinson, 2013). We investigate the electrical resistivity data in the context of
our interpreted structures to ascertain whether the framework relates to regional fluid transfer. The
interpreted geophysical domains have distinct patterns along resistivity data profiles (Figure 4). They are
marked by the presence of resistive crust (>100 Ohm.m) in the interior of the domains A, B, and C reaching
values as high as >1,000 Ohm.m (e.g., the central part of Figure 4c). The contacts between the domains vary
in resistivity. High resistive zones mark the boundary from domains A to B (Figure 4b) and the boundary
from A to C and from A to B (Figure 4c). In contrast, the boundary from domains B to C is more complex in
Model 1 (Figure 4a), which has smaller resistivity values (as low as 100 Ohm.m). Some regions in the west
and central parts of the craton, like the area between the Karari, Kooniba, and Yarlbrinda shear zones
(from 200 to 400 km in Model 2; Figures 4c and 6), which has a higher resistivity (>1,000 Ohm.m). The
deposits of the CGG province are confined to such areas or occur close to steep electrical resistivity gradi-
ents within these domains on magnetotelluric data (Figure 4). The IOCG province lies close to areas with
steep gradients in the resistivity varying from >1000 Ohm.m (to the west) to a linear belt of decreased resis-
tivity (≪1000 Ohm.m) to the east (Figure 6). In contrast to other deposits in the IOCG province, those in
the central‐northern Gawler Craton (e.g., Prominent Hill) lie over gradient zones in the resistivity maps
that are perpendicular to oblique to regional scale shear zones (e.g., Bulgunnia shear zone; Figure 6).
Features that relate to decreased resistivity in the northern part of the craton and within domain C
(Figure 6) represent favorable areas for preexisting fluid flow (e.g., Curtis & Thiel, 2019; Heinson et al.,
2006, 2018).
Our data show that some of the regional structures identified in potential field data and surface mapping are
not exactly coincident with the boundaries of resistivity zones in the deep crust (see resistivity depth slice at
38‐km depth; Figure 6). The resistivity anomalies in the plan section are usually larger than or misfit the
structures. That raises three possibilities: poor horizontal resolution of the data, that the changes in the resis-
tivity occur over broader areas of influence rather than the shear‐zone itself, and that changes in the
resistivity have happened after the establishment of the cratonic framework. In the latter case, the resistivity
grain of the craton would not necessarily follow the previous structures and could, then, have a verticalized
distribution, which is discordant to the listric geometry of the shear zones.
Deposits within the IOCG province occur at distances ranging from 50 to up to 200 km from the boundary of
the geophysical domains B to C (Figures 4a and 4b), where geophysical breaks (shear zones) trending E‐W to
NE‐SW intersect N‐S to NW‐SE striking low‐resistivity anomalies (Figure 6). These E‐W to NE‐SW shear
zones crosscut domain C and can be traced into domain A (Figure 6), connecting to the Yerda Shear Zone
(Figure 1a) and connecting to the CGG deposits in central Gawler Craton. The region around the
Olympic Dam deposit lies over an undulation (greater than 3 km) in the Moho topography (Figure 4a), in
the vicinity of a structural break imaged in the magnetic data (Figures 2c and 2d), and close to a significant
low‐resistivity zone (Figure 6). The CGG deposits also occur close to the boundary of the geophysical
domains in the center of the study area (Figures 4 and 6), and regional‐scale shear zones (e.g., Yerda,
Yarlbrinda, and Kooniba shear zones), which connect to undulations in Moho (Figure 4c) with steep gradi-
ents in the resistivity data.
From the observed patterns in the electrical resistivity and the presence of interpreted major structural
breaks, it seems that the structural framework of the craton may have acted as a control on regional perme-
ability for fluids, especially along structures that intersect the major boundaries (also see Heinson et al.,
2018). However, this framework could have been reactivated several times during the protracted tectono‐
thermal evolution of the craton (e.g., Hall et al., 2018).

MOTTA ET AL. 3097


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Figure 6. Structural interpretation map for the Gawler Craton draped over the horizontal resistivity depth slice at 38‐km
depth from the Australian Lithospheric Architecture Magnetotelluric Project (AUSLAMP) magnetotelluric array
data set. Structural breaks interpreted from potential field data that relate to changes in the electrical structure are
depicted. Key for the shear zones features: KjMZ, Kalinjala Mylonite Zone; KSZ, Karari; ISZ, Ifould; TSZ, Tallacootra;
CSZ, Coorabie; BSZ, Bulgunnia; YeSZ, Yerda; KoSZ, Kooniba; YaSZ, Yarlbrinda shear zones. Key for mineral deposits:
Mineral deposits: PH, Prominent Hill; OD; Olympic Dam; CA, Carrapateena; TA, Tarcoola.

7. Discussion
Our observations on the crustal structure and isotope geochronology suggest that the domains A and B
represent the substratum of the Sleaford and Mulgathing complexes, respectively. Extending the
Mulgathing Complex further north below the Paleoproterozoic metamorphosed sedimentary basins is con-
sistent with previous interpretations constrained from geochronology data inferring the presence of under-
lying Archean crust (Armit et al., 2017; Reid, Jagodzinski, Armit, et al., 2014). Domain C is interpreted as an
extension of the poorly exposed Cooyerdoo Granite (Fraser, McAvaney, et al., 2010), which represents an
extended Mesoarchean basement unit during the Sleafordian Orogeny. An extended Mesoarchean basement
is also interpreted by Curtis and Thiel (2019) from εNd and εHf data on magmatic and inherited zircons from
igneous and meta‐sedimentary rocks and magnetotelluric data along the southern part of the craton, which
is limited on the eastern flank of the craton with the operation long‐lived structures (e.g., Roopena Fault,
Kalinjala Shear Zone). Mesoproterozoic magmatism of the St. Peter Suite is coincident with the domain
A; however, it is characterized by juvenile isotopic responses and thus is unlikely to have formed above

MOTTA ET AL. 3098


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Figure 7. The interpreted crustal framework of the Gawler Craton in contrast to the mantle lithospheric structure by
Skirrow et al. (2018). (1) Interpretations and data provided by Skirrow et al. (2018) and (2) for data from Thiel and
Heinson (2013).

the evolved Archean basement and is therefore interpreted to represent a sequence that was accreted via
overthrusting to the Gawler Craton basement at a convergent margin setting (Betts et al., 2011; Swain
et al., 2008). Neoarchaean basement rocks with similar in age and composition to the Sleaford Complex
(Hand et al., 2007) were buried at this time. Domain B (Figure 3b) coincides with concealed,
Palaeoproterozoic amphibolite to granulite facies metamorphosed approximately 1,790‐ to 1,740‐Ma
Paleoproterozoic sedimentary basins and parts of the Mulgathing Complex (Armit et al., 2017; Hand
et al., 2007; Reid, Jagodzinski, Armit, et al., 2014).
Many of the interpreted shear zones in the area (e.g., this work; Baines et al., 2011; Direen et al., 2005;
Stewart & Betts, 2010a; Thomas et al., 2008) have density, magnetic, and electric resistivity expressions in
the deep crust (Heinson et al., 2006, 2018; Thiel et al., 2005; Thiel & Heinson, 2013; Figures 2, 3, and 6).
These shear zones have disrupted the geometry of the basement of the craton. However, their deep crustal
traces differ spatially from their surface expression. Furthermore, the data lack spectral and spatial resolu-
tion to provide details of the finer‐scale structure, due to the long‐wavelength nature of the satellite gravity
data and representation of the field in a relatively small degree and order (250, nearly 80 km in plane coor-
dinates; Mayer‐Guerr, 2015).
The electrical conductors observed across the craton (e.g., Figures 3, 5, and 6) are interpreted as zones of
fluid flow along shear zones probably during the Paleoproterozoic to Mesoproterozoic, which decreased
the resistivity through aqua‐carbonic fluid entrapment (Heinson et al., 2006; Thiel & Heinson, 2013) or sul-
fide precipitation (Heinson et al., 2018). The significant domain boundaries within Gawler Craton
(Figures 3b and 6) appear to have acted as agents for energy and fluid flux between different geochemical
reservoirs across the craton basement along secondary structures (Bastrakov et al., 2007; Direen & Lyons,
2007; Fraser et al., 2007; Haynes et al., 1995; Kontonikas‐Charos et al., 2014, 2017; Uvarova et al., 2017)
including shallow levels in the overlying sedimentary basins and basement (Cherry et al., 2017; Reid &
Fabris, 2015). This is evident from the regional alteration and chemical footprint observed in both northern
and southern sectors of the IOCG province (Kontonikas‐Charos et al., 2014, 2017) and similarities in timing
and composition of the hydrothermal alterations between the IOCG and CGG provinces (Bastrakov et al.,
2007; Fraser et al., 2007). This fundamental role of the major structures as pathways persists through time
due to reactivation associated with tectonic switches, as previously envisaged around the Olympic Dam
deposit (Direen & Lyons, 2007; Heinson et al., 2006, 2018; Korsch & Doublier, 2016; Stewart & Betts,
2010a, 2010b). However, the resistivity structure could also have been affected throughout the late
Mesoproterozoic or by younger events (Foden et al., 2002; Hall et al., 2018).
Our interpreted crustal architecture aligns with the mantle architecture interpreted by Skirrow et al. (2018)
(Figure 7), where both IOCG and CGG deposits lie along major structural breaks that connect different

MOTTA ET AL. 3099


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

compartments of the mantle to the crust along long‐lasting lithospheric scars (in the sense of Heron et al.,
2016). The oldest crustal fragment represented by Domain C is mostly rooted by the metasomatized litho-
spheric mantle. Domains A and B reside over a residual crust underplate and depleted mantle lithosphere,
respectively. Tectonic switches and crustal foundering of the ancient Archean cratonic lithosphere in a post-
subduction scenario could be responsible for reactivation of the inherited trans‐crustal structures. These set-
tings provided the environment for the transfer of deep‐seated, mantle‐derived fluids through the whole
lithosphere and ingress into the crust. For instance, the CGG deposits lie in a lithospheric region that has
undergone crustal reworking during the early Mesoproterozoic Wartaken event and Kararan Orogeny and
the mixture of evolved and primitive magmas and associated fluids. Areas that are proximal to crustal
boundaries with a prolonged crustal reworking history are recognized as important sites for orogenic gold
deposits to form (Huston et al., 2016). In contrast, the IOCG deposits lie in an inward position to the crustal
domain boundaries during Mesoproterozoic tectonism, likely a back‐arc setting above a highly heteroge-
neous, evolved crust, forming in a postsubduction setting (Skirrow et al., 2018). Following this rationale,
the Olympic Province is located on the upper plate of a suture zone where the lithosphere is characterized
by the hydrated mantle (Skirrow et al., 2018), which was refertilized during plume impingement (Betts
et al., 2009). Furthermore, our results suggest that the formation of the Mesoproterozoic IOCG deposits in
the Gawler Craton could have formed similarly to mineral deposits formed in the Phanerozoic Andean chain
(Rosenbaum et al., 2005; Sillitoe, 2003), a notion previously introduced by Richards and Mumin (2013) on
different grounds (e.g., overall hydrothermal alteration and isotope geochemistry). The envisioned frame-
work also locates the CGG province above a region of refertilized and metasomatized lithospheric mantle,
similar to the Olympic province. However, the gold deposits are located closer to the margin of the over‐
riding plate at the time of the mineralization. Similarities to modern systems include the necessity of pro-
found refertilization of the lithospheric mantle for gold deposits to form in the Jurassic (Tassara et al.,
2017), with the impingement of a mantle plume for the fertilization of the lithosphere required for the depos-
its to form. This is a younger analog for the Precambrian plume‐modified orogenesis proposed by Betts et al.
(2009) for the Gawler Craton. Thus, the interpreted crustal framework in this work correlates with observa-
tions in the lithospheric mantle of the Gawler Craton, with both playing a major role during the mineral
deposit formation and can be used as a proxy for its exploration (Hagemann et al., 2016; Huston et al.,
2016). It is important to notice that the spatial resolution of the data sets interpreted here and those of pre-
vious studies are not the same. Therefore, geological and tectonic features identified at different crustal levels
and in the lithosphere will not exactly correlate. Nevertheless, there is a broad relationship between the dif-
ferent examined lithospheric levels that constrain the architecture of Gawler Craton.

8. Conclusion
The use of satellite gravity data proved to be a powerful tool to provide insights into the deep crustal archi-
tecture of the Gawler Craton and should be considered in evaluations of first‐order crustal structure else-
where. The satellite gravity data provide more salient geological information when analyzed in
conjunction with ancillary data (e.g., aeromagnetic, electrical resistivity, and seismic tomography and deep
seismic reflection data). This study provides a new perspective on the value of satellite‐gravity data sets to
resolve major, regional‐scale tectonic features that relate to the formation of mineral deposits. In particular,
the data allowed comparison from the surface geological constraints gleaned from geology and geochemical
data sets to the deep crustal framework of the craton. The three domains of Archean crust in the craton had
an expanded crustal extension. The major crustal structure was possibly formed during the Sleafordian
Orogeny due to the amalgamation of the extended Mesoarchean Cooyerdoo Granite to the Neoarchean
Mulgathing and Sleaford complexes and the basement that underlies these complexes. Structures in the
Archean‐aged framework were formed and subsequently reactivated during several tectonic events, includ-
ing the approximately 1,590‐Ma tectonic‐magmatic event, which included emplacement of the Hiltaba Suite
and GRV and the associated IOCG and gold‐only mineralization.
The interpreted crustal structures and elements correlate with prominent changes in the electric resistivity
of the craton throughout the crust. The areas associated with decreased electric resistivity are interpreted as
the product of fluid transport through crust with structurally enhanced permeability. The association of the
observed crustal structures and domains to the IOCG and gold‐only systems suggests that these structural
zones are first‐order controls on the distribution of the fluids during the Mesoproterozoic mineralizing

MOTTA ET AL. 3100


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

event. The interpreted crustal framework hosts inherited, spatially significant, trans‐crustal boundaries
within the Gawler Craton that influenced the distribution of the metallogenic provinces. This association
suggests the coupling of the crust and the lithospheric mantle structure, which controlled the geometry of
the crust developing above. This coupling gave rise to the mineralizing systems, with an extensive minera-
lization footprint within both IOCG and CGG provinces in the craton.

Acknowledgments References
Geoscience Australia and GSSA are
thanked for providing the geological, Allen, S. R., McPhie, J., Ferris, G., & Simpson, C. (2008). Evolution and architecture of a large felsic Igneous Province in western Laurentia:
geochemistry, and geophysical data. We The 1.6 Ga Gawler Range Volcanics, South Australia. Journal of Volcanology and Geothermal Research, 172(1–2), 132–147. https://doi.
must also acknowledge that the mag- org/10.1016/j.jvolgeores.2005.09.027
netotelluric data were collected in col- Amante, C., & Eakins, B. W. (2009). ETOPO1 1 arc‐minute global relief model: Procedures, data sources and analysis, NOAA Technical
laboration between the GSSA and the Memorandum NESDIS NGDC‐24. https://doi.org/10.7289/V5C8276M
University of Adelaide with PACE Archibald, N., Gow, P., & Boschetti, F. (1999). Multiscale edge analysis of potential field data. Exploration Geophysics, 30(1‐2), 38–44.
initiative funding by GSSA. The MT https://doi.org/10.1071/EG999038
instruments are part of the national Armit, R., Betts, P. G., Schaefer, B. F., Yi, K., Kim, Y., Dutch, R. A., et al. (2017). Late Palaeoproterozoic evolution of the buried northern
instrument pool under the auspices of Gawler Craton. Precambrian Research, 291, 178–201. https://doi.org/10.1016/j.precamres.2017.01.023
Australian National Seismic Imaging Baines, G., Giles, D., Betts, P. G., & Backé, G. (2011). Locating a major Proterozoic crustal boundary beneath the Eastern Officer Basin,
Resource (ANSIR) and funded by Australia. Precambrian Research, 191(3–4), 120–140. https://doi.org/10.1016/j.precamres.2011.09.001
NCRIS (National Collaborative Barzaghi, R., Migliaccio, F., Reguzzoni, M., & Albertella, A. (2015). The Earth gravity field in the time of satellites. Rendiconti Lincei, 26(S1),
Research Infrastructure Scheme). The 13–23. https://doi.org/10.1007/s12210‐015‐0382‐9
Australian Academy of Sciences, Bastrakov, E. N., Skirrow, R. G., & Davidson, G. J. (2007). Fluid evolution and origins of iron oxide Cu‐Au prospects in the Olympic Dam
Department of Education and Training, district, Gawler Craton, South Australia. Economic Geology, 102(8), 1415–1440. https://doi.org/10.2113/gsecongeo.102.8.1415
and the Australian Embassy in Brazil Belousova, E. A., Reid, A. J., Griffin, W. L., & O'Reilly, S. Y. (2009). Rejuvenation vs. recycling of Archean crust in the Gawler Craton, South
are thanked for funding J. G. M. to visit Australia: Evidence from U–Pb and Hf isotopes in detrital zircon. Lithos, 113(3–4), 570–582. https://doi.org/10.1016/j.lithos.2009.06.028
Monash University through the 2017 Betts, P. G., Armit, R. J., Stewart, J., Aitken, A. R. A., Ailleres, L., Donchak, P., et al. (2016). Australia and Nuna. Geological Society, London,
Australia‐Brazil Postgraduate Research Special Publications, 424(1), 47–81. https://doi.org/10.1144/SP424.2
Internship program. The Brazilian Betts, P. G., & Giles, D. (2006). The 1800–1100 Ma tectonic evolution of Australia. Precambrian Research, 144(1–2), 92–125. https://doi.org/
National Council for Scientific and 10.1016/j.precamres.2005.11.006
Technological Development (CNPQ) is Betts, P. G., Giles, D., & Aitken, A. (2011). Palaeoproterozoic accretion processes of Australia and comparisons with Laurentia.
acknowledged for the research project International Geology Review, 53(11–12), 1357–1376. https://doi.org/10.1080/00206814.2010.527646
(Proc. Nr. 401316/2014‐9) and research Betts, P. G., Giles, D., Foden, J., Schaefer, B. F., Mark, G., Pankhurst, M. J., et al. (2009). Mesoproterozoic plume‐modified orogenesis in
grants to J.G.M. and C.R.S.F. (Proc. Nr. eastern Precambrian Australia. Tectonics, 28, TC3006. https://doi.org/10.1029/2008TC002325
309712/2017‐30). ARAMZ Geo is Betts, P. G., Giles, D., & Schaefer, B. F. (2008). Comparing 1800–1600 Ma accretionary and basin processes in Australia and Laurentia:
thanked for providing access to the Possible geographic connections in Columbia. Precambrian Research, 166(1–4), 81–92. https://doi.org/10.1016/j.precamres.2007.03.007
Leapfrog Geo 3D modeling software. Betts, P. G., Giles, D., Schaefer, B. F., & Mark, G. (2007). 1600–1500 Ma hotspot track in eastern Australia: Implications for Mesoproterozoic
J. G. M. is thankful to João Paulo Araújo continental reconstructions. Terra Nova, 19(6), 496–501. https://doi.org/10.1111/j.1365‐3121.2007.00778.x
Pitombeira for help with geochemical Betts, P. G., Valenta, R. K., & Finlay, J. (2003). Evolution of the Mount Woods Inlier, northern Gawler Craton, Southern Australia: An
data. S. C. and S. T. publish with the integrated structural and aeromagnetic analysis. Tectonophysics, 366(1–2), 83–111. https://doi.org/10.1016/S0040‐1951(03)00062‐3
permission of the Director of the Bouman, J., Ebbing, J., Fuchs, M., Sebera, J., Lieb, V., Szwillus, W., et al. (2016). Satellite gravity gradient grids for geophysics. Scientific
Geological Survey of South Australia. Reports, 6(August 2015)(1), 21050. https://doi.org/10.1038/srep21050
Editor Paul Tregoning and the reviewer Bouman, J., Ebbing, J., Meekes, S., Fattah, R. A., Fuchs, M., Gradmann, S., et al. (2015). GOCE gravity gradient data for lithospheric
Michael Doublier, and the anonymous modeling. International Journal of Applied Earth Observation and Geoinformation, 35(PA), 16–30. https://doi.org/10.1016/j.
reviewer are thanked for taking their jag.2013.11.001
time to provide very constructive feed- Braitenberg, C. (2015). Exploration of tectonic structures with GOCE in Africa and across‐continents. International Journal of Applied
back on the manuscript. The geophysi- Earth Observation and Geoinformation, 35(PA), 88–95. https://doi.org/10.1016/j.jag.2014.01.013
cal data, which are original to this Budd, A. (2006). The Tarcoola Goldfeld of the Central Gawler Gold province, and the Hiltaba Association Granites, Gawler craton, South
contribution (satellite gravity data and Australia, (PhD thesis). Australian National University, Canberra, Australia.
resistivity data derived from magneto- Budd, A. R., & Skirrow, R. G. (2007). The nature and origin of gold deposits of the Tarcoola Goldfield and implications for the central
telluric surveys), and isotope geochem- prospect: The nature and origin of gold deposits of the Tarcoola Goldfield and. Society of Economic Geologists, 102(8), 1541–1563. https://
istry and geochronology data are made doi.org/10.2113/gsecongeo.102.8.1541
available to interested readers on the Champion, D. C. (2013). Neodymium depleted mantle model age map of Australia: explanatory notes and user guide, Geoscience Australia
following link https://monash.figshare. Record 2013/44 (Vol. 44). Canberra: Geoscience Australia. https://doi.org/10.11636/Record.2013.044
com/s/8c1a225e9465598bec49 under Cherry, A. R., McPhie, J., Kamenetsky, V. S., Ehrig, K., Keeling, J. L., Kamenetsky, M. B., et al. (2017). Linking Olympic Dam and the
the DOI https://doi.org/10.26180/ Cariewerloo Basin: Was a sedimentary basin involved in formation of the world's largest uranium deposit? Precambrian Research,
5c1c4dff46114. Data sets derived from 300(December 2016, 168–180. https://doi.org/10.1016/j.precamres.2017.08.002
third party (e.g., airborne magnetics Chopping, R., & Kennett, B. L. N. (2015). Maximum depth of magnetisation of Australia, its uncertainty, and implications for Curie depth.
data, gravity gradient strings, and pas- GeoResJ, 7, 70–77. https://doi.org/10.1016/j.grj.2015.06.003
sive seismology stations) should be Christensen, N. I. (1996). Poisson's ratio and crustal seismology. Journal of Geophysical Research, 101(B2), 3139–3156. https://doi.org/
taken in the original repositories from 10.1029/95JB03446
the respective authors as cited. Ciobanu, C. L., Wade, B. P., Cook, N. J., Schmidt Mumm, A., & Giles, D. (2013). Uranium‐bearing hematite from the Olympic Dam Cu–U–
Au deposit, South Australia: A geochemical tracer and reconnaissance Pb–Pb geochronometer. Precambrian Research, 238, 129–147.
https://doi.org/10.1016/j.precamres.2013.10.007
Cowley, W. (2015). Depth to basement map 2015. Department of State Development, Resources and Energy, South Australia.
Creaser, R. A. (1989). The geology and petrology of Middle Proterozoic felsic magmatism of the Stuart Shelf, South Australia, (PhD thesis).
La Trobe University.
Creaser, R. A., & Cooper, J. A. (1993). U–Pb geochronology of middle Proterozoic felsic magmatism surrounding the Olympic Dam Cu‐U‐
Au‐Ag and Moonta Cu‐Au‐Ag deposits, South Australia. Economic Geology, 88(1), 186–197. https://doi.org/10.2113/gsecongeo.88.1.186
Curtis, S., & Thiel, S. (2019). Identifying lithospheric boundaries using magnetotellurics and Nd isotope geochemistry: An example from
the Gawler Craton, Australia. Precambrian Research, 320(November 2018, 403–423. https://doi.org/10.1016/j.precamres.2018.11.013

MOTTA ET AL. 3101


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Dentith, M. C., & Mudge, S. T. (2014). Geophysics for the mineral exploration geoscientist. Cambridge: Cambridge University Press. Retrieved
from http://www.cambridge.org/us/academic/subjects/earth‐and‐environmental‐science/solid‐earth‐geophysics/geophysics‐mineral‐
exploration‐geoscientist
Dhu, T. (2011). Reduced to pole total magnetic intensity grid of South Australia. Adelaide, SA.: Department of Primary Industries and
Resources, South Australia.
Direen, N. G., Cadd, A. G., Lyons, P., & Teasdale, J. P. (2005). Architecture of Proterozoic shear zones in the Christie Domain, western
Gawler Craton, Australia: Geophysical appraisal of a poorly exposed orogenic terrane. Precambrian Research, 142(1–2), 28–44. https://
doi.org/10.1016/j.precamres.2005.09.007
Direen, N. G., & Lyons, P. (2007). Regional crustal setting of iron oxide Cu‐Au mineral systems of the olympic dam region, South Australia:
Insights from potential‐field modeling. Economic Geology, 102(8), 1397–1414. https://doi.org/10.2113/gsecongeo.102.8.1397
Doublier, M. P., Dutch, R. A., Clark, D., Pawley, M. J., & Fraser, G. L. (2015). Interpretation of the western Gawler Craton section of seismic
line 13GA‐EG1.
Drummond, B., Lyons, P., Goleby, B., & Jones, L. (2006). Constraining models of the tectonic setting of the giant Olympic Dam iron oxide‐
copper‐gold deposit, South Australia, using deep seismic reflection data. Tectonophysics, 420(1–2), 91–103. https://doi.org/10.1016/j.
tecto.2006.01.010
Dutch, R., Hand, M., & Kinny, P. D. (2008). High‐grade Paleoproterozoic reworking in the southeastern Gawler Craton, South Australia.
Australian Journal of Earth Sciences, 55(8), 1063–1081. https://doi.org/10.1080/08120090802266550
Ebbing, J., Bouman, J., Haagmans, R., Meekes, J. A. C., & Fattah, R. A. (2013). Advancements in satellite gravity gradient data for crustal
studies. The Leading Edge, 32(August(8), 900–906. https://doi.org/10.1190/tle32080900.1
Ebbing, J., Haas, P., Ferraccioli, F., Pappa, F., Szwillus, W., & Bouman, J. (2018). Earth tectonics as seen by GOCE—Enhanced satellite
gravity gradient imaging. Scientific Reports, 8(1), 16356. https://doi.org/10.1038/s41598‐018‐34733‐9
Ferris, G., & Schwarz, M. (2003). Proterozoic gold province of the central Gawler Craton. MESA Journal, 31(2‐3), 207–246. https://doi.org/
10.1016/j.stueduc.2005.05.011
Foden, J., Song, S. H., Turner, S., Elburg, M., Smith, P. B., Van Der Steldt, B., & Van Penglis, D. (2002). Geochemical evolution of litho-
spheric mantle beneath S.E. South Australia. Chemical Geology, 182(2–4), 663–695. https://doi.org/10.1016/S0009‐2541(01)00347‐3
Fraser, G., & Blewett, R. (2010)). Geological interpretation of deep seismic reflection and magnetotelluric line 08GA‐G1: Eyre Peninsula,
Gawler Craton, South Australia. In R. Korsch & N. Kositcin (Eds.), GOMA (Gawler Craton‐Officer Basin‐Musgrave Province‐Amadeus
Basin), South Australian Seimsic and MT workshop (Vol. 2010, pp. 81–95). Perth, Australia.
Fraser, G., McAvaney, S., Neumann, N., Szpunar, M., & Reid, A. J. (2010). Discovery of early Mesoarchean crust in the eastern Gawler
Craton, South Australia. Precambrian Research, 179(1–4), 1–21. https://doi.org/10.1016/j.precamres.2010.02.008
Fraser, G. L., & Neumann, N. (2010). New SHRIMP U–Pb zircon ages from the Gawler Craton and Curnamona Province, South Australia,
2008–2010. Record 2010/16.
Fraser, G. L., Skirrow, R. G., Schmidt‐Mumm, A., & Holm, O. (2007). Mesoproterozoic gold in the central Gawler craton, South Australia:
Geology, alteration, fluids, and timing. Economic Geology, 102(8), 1511–1539. https://doi.org/10.2113/gsecongeo.102.8.1511
Goldstein, S. L., Onions, R. K., Hamilton, P. J., O'Nions, R. K., & Hamilton, P. J. (1984). A S m‐N d isotopic study of atmospheric dusts and
particulates from major river systems. Earth and Planetary Science Letters, 70, 221–236. https://doi.org/10.1016/0012‐821X(84)90007‐4
Gow, P. A., Wall, V. J., Oliver, N. H. S., & Valenta, R. K. (1994). Proterozoic iron oxide (Cu‐U‐Au‐REE) deposits: Further evidence of
hydrothermal origins. Geology, 22(7), 633–636. https://doi.org/10.1130/0091‐7613(1994)022<0633:PIOCUA>2.3.CO;2
Hagemann, S. G., Lisitsin, V. A., & Huston, D. L. (2016). Mineral system analysis: Quo vadis. Ore Geology Reviews, 76, 504–522. https://doi.
org/10.1016/j.oregeorev.2015.12.012
Hall, J. W., Glorie, S., Reid, A. J., Collins, A. S., Jourdan, F., Danišík, M., & Evans, N. (2018). Thermal history of the northern Olympic
Domain, Gawler Craton; correlations between thermochronometric data and mineralising systems. Gondwana Research, 56, 90–104.
https://doi.org/10.1016/j.gr.2018.01.001
Hand, M., Reid, A., & Jagodzinski, L. (2007). Tectonic framework and evolution of the Gawler craton, Southern Australia. Economic
Geology, 102(8), 1377–1395. https://doi.org/10.2113/gsecongeo.102.8.1377
Haynes, D. W., Cross, K. C., Bills, R. T., & Reed, M. H. (1995). Olympic dam ore genesis: a fluid‐mixing model. Economic Geology, 90(2),
281–307. https://doi.org/10.2113/gsecongeo.90.2.281
Heinson, G., Didana, Y., Soeffky, P., Thiel, S., & Wise, T. (2018). The crustal geophysical signature of a world‐class magmatic mineral
system. Scientific Reports, 8(1), 10608. https://doi.org/10.1038/s41598‐018‐29016‐2
Heinson, G. S., Direen, N. G., & Gill, R. M. (2006). Magnetotelluric evidence for a deep‐crustal mineralizing system beneath the Olympic
Dam iron oxide copper‐gold deposit, southern Australia. Geology, 34(7), 573–576. https://doi.org/10.1130/G22222.1
Henderson, R. G., & Zietzt, I. (1949). The upward continuation of anomalies in total magnetic intensity fields. Geophysics, 14(4), 517–534.
https://doi.org/10.1190/1.1437560
Heron, P. J., Pysklywec, R. N., & Stephenson, R. (2016). Lasting mantle scars lead to perennial plate tectonics. Nature Communications,
7(1), 1–7. https://doi.org/10.1038/ncomms11834
Hornby, P., Boschetti, F., & Horowitz, F. G. (1999). Analysis of potential field data in the wavelet domain. Geophysical Journal
International, 137(1), 175–196. https://doi.org/10.1046/j.1365‐246x.1999.00788.x
Howard, K., Hand, M., Barovich, K., & Belousova, E. (2011). Provenance of late Palaeoproterozoic cover sequences in the central Gawler
Craton: Exploring stratigraphic correlations in eastern Proterozoic Australia using detrital zircon ages, Hf and Nd isotopic data.
Australian Journal of Earth Sciences, 58(5), 475–500. https://doi.org/10.1080/08120099.2011.577753
Howard, K. E., Hand, M., Barovich, K. M., Payne, J. L., Cutts, K. A., & Belousova, E. A. (2011). U–Pb zircon, zircon Hf and whole‐rock Sm‐
Nd isotopic constraints on the evolution of Paleoproterozoic rocks in the northern Gawler Craton. Australian Journal of Earth Sciences,
58(6), 615–638. https://doi.org/10.1080/08120099.2011.594905
Huston, D. L., Mernagh, T. P., Hagemann, S. G., Doublier, M. P., Fiorentini, M., Champion, D. C., et al. (2016). Tectono‐metallogenic
systems—The place of mineral systems within tectonic evolution, with an emphasis on Australian examples. Ore Geology Reviews,
76(September 2015), 168–210. https://doi.org/https://doi.org/10.1016/j.oregeorev.2015.09.005
Jagodzinski, E. A. (2005). Compilation of SHRIMP U–Pb geochronological data, Olympic domain, Gawler Craton, South Australia, 2001–
2003. Record 2005/20.
Johnson, J. P. (1993). The geochronology and radiogenic isotope systematics of the Olympic Dam copper‐uranium‐gold‐silver deposit,
South Australia (PhD thesis). Australian National University, Canberra, Australia.
Johnson, J. P., & Cross, K. C. (1995). U–Pb geochronological constraints on the genesis of the Olympic Dam Cu‐U‐Au‐Ag deposit, South
Australia. Economic Geology, 90(5), 1046–1063. https://doi.org/10.2113/gsecongeo.90.5.1046

MOTTA ET AL. 3102


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Johnson, J. P., & McCulloch, M. T. (1995). Sources of mineralising fluids for the Olympic Dam deposit (South Australia): SmNd isotopic
constraints. Chemical Geology, 121(1–4), 177–199. https://doi.org/10.1016/0009‐2541(94)00125‐R
Kennett, B. L. N., & Saygin, E. (2015). The nature of the Moho in Australia from reflection profiling: A review. GeoResJ, 5, 74–91. https://
doi.org/10.1016/j.grj.2015.02.001
Kirchenbaur, M., Maas, R., Ehrig, K., Kamenetsky, V. S., Strub, E., Ballhaus, C., & Münker, C. (2016). Uranium and Sm isotope studies of
the supergiant Olympic Dam Cu‐Au‐U‐Ag deposit, South Australia. Geochimica et Cosmochimica Acta, 180, 15–32. https://doi.org/
10.1016/j.gca.2016.01.035
Kontonikas‐Charos, A., Ciobanu, C. L., & Cook, N. J. (2014). Albitization and redistribution of REE and Y in IOCG systems: Insights from
Moonta‐Wallaroo, Yorke Peninsula, South Australia. Lithos, 208, 178–201. https://doi.org/10.1016/j.lithos.2014.09.001
Kontonikas‐Charos, A., Ciobanu, C. L., Cook, N. J., Ehrig, K., Krneta, S., & Kamenetsky, V. S. (2017). Feldspar evolution in the Roxby
Downs Granite, host to Fe‐oxide Cu‐Au‐(U) mineralization at Olympic Dam, South Australia. Ore Geology Reviews, 80, 838–859. https://
doi.org/10.1016/j.oregeorev.2016.08.019
Korsch, R. J., & Doublier, M. P. (2016). Major crustal boundaries of Australia, and their significance in mineral systems targeting. Ore
Geology Reviews, 76, 211–228. https://doi.org/10.1016/j.oregeorev.2015.05.010
Laske, G., Masters, G., Ma, Z., & Pasyanos, M. E. (2012). CRUST1.0: An updated global model of the Earth's crust, EGU General Assembly
(Vol. 14, p. 3743). Vienna, Austria.
Lyons, P., & Goleby, B. R. (2005). The 2003 Gawler Seismic Survey: Notes of the seismic workshop held at Gawler Craton State of Play 2004.
Geoscience Australia Record 2015/19.
Mayer‐Guerr, T. (2015). The combined satellite gravity field model GOCO05s. EGU General Assembly Conference Abstracts, 17, 12364.
McLean, M. A., & Betts, P. G. (2003). Geophysical constraints of shear zones and geometry of the Hiltaba Suite granites in the western
Gawler Craton, Australia. Australian Journal of Earth Sciences, 50(4), 525–541. https://doi.org/10.1046/j.1440‐0952.2003.01010.x
McPhie, J., Kamenetsky, V., Allen, S., Ehrig, K., Agangi, A., & Bath, A. (2011). The fluorine link between a supergiant ore deposit and a
silicic large igneous province. Geology, 39(11), 1003–1006. https://doi.org/10.1130/G32205.1
McPhie, J., Kamenetsky, V. S., Chambefort, I., Ehrig, K., & Green, N. (2011). Origin of the supergiant Olympic Dam Cu‐U‐Au‐Ag deposit,
South Australia: Was a sedimentary basin involved. Geology, 39(8), 795–798. https://doi.org/10.1130/G31952.1
McPhie, J., Orth, K., Kamenetsky, V., Kamenetsky, M., & Ehrig, K. (2016). Characteristics, origin and significance of Mesoproterozoic
bedded clastic facies at the Olympic Dam Cu‐U‐Au‐Ag deposit, South Australia. Precambrian Research, 276, 85–100. https://doi.org/
10.1016/j.precamres.2016.01.029
Michard, A., Gurriet, P., Soudant, M., & Albarede, F. (1985). Nd isotopes in French Phanerozoic shales: external vs. internal aspects of
crustal evolution. Geochimica et Cosmochimica Acta, 49(2), 601–610. https://doi.org/10.1016/0016‐7037(85)90051‐1
Oreskes, N., & Einaudi, M. T. (1992). Origin of hydrothermal fluids at Olympic Dam: Preliminary results from fluid inclusions and stable
isotopes. Economic Geology, 87(1), 64–90. https://doi.org/10.2113/gsecongeo.87.1.64
Panet, I., Pajot‐Métivier, G., Greff‐Lefftz, M., Métivier, L., Diament, M., & Mandea, M. (2014). Mapping the mass distribution of Earth's
mantle using satellite‐derived gravity gradients. Nature Geoscience, 7(2), 131–135. https://doi.org/10.1038/ngeo2063
Pankhurst, M. J., Schaefer, B. F., Betts, P. G., Phillips, N., & Hand, M. (2011). A Mesoproterozoic continental flood rhyolite province, the
Gawler Ranges, Australia: The end member example of the Large Igneous Province clan. Solid Earth, 2(1), 25–33. https://doi.org/
10.5194/se‐2‐25‐2011
Payne, J. L., Barovich, K. M., & Hand, M. (2006). Provenance of metasedimentary rocks in the northern Gawler Craton, Australia:
Implications for Palaeoproterozoic reconstructions. Precambrian Research, 148(3‐4), 275–291. https://doi.org/10.1016/j.
precamres.2006.05.002
Payne, J. L., Ferris, G., Barovich, K., & Hand, M. (2010). Pitfalls of classifying ancient magmatic suites with tectonic discrimination dia-
grams: An example from the Paleoproterozoic Tunkillia Suite, southern Australia. Precambrian Research, 177(3‐4), 227–240. https://doi.
org/10.1016/j.precamres.2009.12.005
Reid, A. J., & Fabris, A. (2015). Influence of preexisting low metamorphic grade sedimentary successions on the distribution of iron oxide
copper‐gold mineralization in the Olympic Cu‐Au Province, Gawler Craton. Economic Geology, 110(8), 2147–2157. https://doi.org/
10.2113/econgeo.110.8.2147
Reid, A. J., Flint, R., Maas, R., Howard, K., & Belousova, E. A. (2009). Geochronological and isotopic constraints on Palaeoproterozoic skarn
base metal mineralization in the central Gawler Craton. Ore Geology Reviews, 36(4), 350–362. https://doi.org/10.1016/j.
oregeorev.2009.09.001
Reid, A. J., & Hand, M. (2012). Mesoarchean to Mesoproterozoic evolution of the southern Gawler Craton, South Australia. Episodes,
35(March, 216–225.
Reid, A. J., Jagodzinski, E. A., Armit, R. J., Dutch, R. A., Kirkland, C. L., Betts, P. G., & Schaefer, B. F. (2014). U–Pb and Hf isotopic evidence
for Neoarchean and Paleoproterozoic basement in the buried northern Gawler Craton, South Australia. Precambrian Research, 250,
127–142. https://doi.org/10.1016/j.precamres.2014.05.019
Reid, A. J., Jagodzinski, E. A., Fraser, G. L., & Pawley, M. J. (2014). SHRIMP U–Pb zircon age constraints on the tectonics of the Neoarchean
to early Paleoproterozoic transition within the Mulgathing Complex, Gawler Craton, South Australia. Precambrian Research, 250, 27–49.
https://doi.org/10.1016/j.precamres.2014.05.013
Reid, A. J., & Payne, J. L. (2017). Magmatic zircon Lu‐Hf isotopic record of juvenile addition and crustal reworking in the Gawler Craton,
Australia. Lithos, 292‐293, 294–306. https://doi.org/10.1016/j.lithos.2017.08.010
Richards, J. P., & Mumin, A. H. (2013). Magmatic‐hydrothermal processes within an evolving Earth: Iron oxide‐copper‐gold and porphyry
Cu ± Mo ± Au deposits. Geology, 41(7), 767–770. https://doi.org/10.1130/G34275.1
Robertson, K., Heinson, G., & Thiel, S. (2016). Lithospheric reworking at the Proterozoic???Phanerozoic transition of Australia imaged
using AusLAMP Magnetotelluric data. Earth and Planetary Science Letters, 452, 27–35. https://doi.org/10.1016/j.epsl.2016.07.036
Rosenbaum, G., Giles, D., Saxon, M., Betts, P. G., Weinberg, R. F., & Duboz, C. (2005). Subduction of the Nazca Ridge and the Inca Plateau:
Insights into the formation of ore deposits in Peru. Earth and Planetary Science Letters, 239(1–2), 18–32. https://doi.org/10.1016/j.
epsl.2005.08.003
Salmon, M., Kennett, B. L. N., & Saygin, E. (2013). Australian Seismological Reference Model (AuSREM): Crustal component. Geophysical
Journal International, 192(1), 190–206. https://doi.org/10.1093/gji/ggs004
Schaefer, B. F. (1998). Insights into Proterozoic tectonics evolution from the Southern Eyre Peninsula. University of Adelaide.
Schlegel, T. U., Wagner, T., Boyce, A., & Heinrich, C. A. (2016). A magmatic source of hydrothermal sulfur for the Prominent Hill deposit
and associated prospects in the Olympic iron oxide copper‐gold (IOCG) province of South Australia. Ore Geology Reviews, 89, 1058–1090.
https://doi.org/10.1016/j.oregeorev.2016.09.002

MOTTA ET AL. 3103


Journal of Geophysical Research: Solid Earth 10.1029/2018JB016829

Sillitoe, R. H. (2003). Iron oxide‐copper‐gold deposits: An Andean view. Mineralium Deposita, 38(7), 787–812. https://doi.org/10.1007/
s00126‐003‐0379‐7
Sippl, C. (2016). Moho geometry along a north–south passive seismic transect through Central Australia. Tectonophysics, 676, 56–69.
https://doi.org/10.1016/j.tecto.2016.03.031
Skirrow, R., Bastrakov, E. N., & Barovich, K. (2007). Timing of iron oxide copper‐gold hydrothermal activity and isotope constraints on
metal sources in the Gawler Craton, South Australia Prospect: Timing of iron oxide Cu‐Au‐(U) hydrothermal activity and Nd isotope
constraints. Economic Geology, 102(8), 1441–1470. https://doi.org/10.2113/gsecongeo.102.8.1441
Skirrow, R. G., van der Wielen, S. E., Champion, D. C., Czarnota, K., & Thiel, S. (2018). Lithospheric architecture and mantle metaso-
matism linked to iron oxide Cu‐Au ore formation: Multidisciplinary evidence from the Olympic Dam Region, South Australia.
Geochemistry, Geophysics, Geosystems, 19(8), 2673–2705. https://doi.org/10.1029/2018GC007561
Stewart, J. R., & Betts, P. G. (2010a). Implications for Proterozoic plate margin evolution from geophysical analysis and crustal‐scale
modeling within the western Gawler Craton, Australia. Tectonophysics, 483(1–2), 151–177. https://doi.org/10.1016/j.tecto.2009.11.016
Stewart, J. R., & Betts, P. G. (2010b). Late Paleo‐Mesoproterozoic plate margin deformation in the southern Gawler Craton: Insights from
structural and aeromagnetic analysis. Precambrian Research, 177(1–2), 55–72. https://doi.org/10.1016/j.precamres.2009.11.004
Swain, G., Woodhouse, A., Hand, M., Barovich, K., Schwarz, M., & Fanning, C. M. (2005). Provenance and tectonic development of the late
Archaean Gawler Craton, Australia; U–Pb zircon, geochemical and Sm‐Nd isotopic implications. Precambrian Research, 141(3–4),
106–136. https://doi.org/10.1016/j.precamres.2005.08.004
Swain, G., Woodhouse, A., Hand, M., Barovich, K., Schwarz, M., & Fanning, C. M. (2008). Petrogenesis of the St Peter Suite, southern
Australia: arc magmatism and Proterozoic crustal growth of the South Australian Craton. Precambrian Research, 166(1‐4), 283–296.
https://doi.org/10.1016/j.precamres.2007.07.028
Swain, G. M., Hand, M., Teasdale, J., Rutherford, L., & Clark, C. (2005). Age constraints on terrane‐scale shear zones in the Gawler Craton,
southern Australia. Precambrian Research, 139(3–4), 164–180. https://doi.org/10.1016/j.precamres.2005.06.007
Szpunar, M., Hand, M., Barovich, K., Jagodzinski, E., & Belousova, E. (2011). Isotopic and geochemical constraints on the Paleoproterozoic
Hutchison Group, southern Australia: Implications for Paleoproterozoic continental reconstructions. Precambrian Research, 187(1–2),
99–126. https://doi.org/10.1016/j.precamres.2011.02.006
Tassara, S., González‐Jiménez, J. M., Reich, M., Schilling, M. E., Morata, D., Begg, G., et al. (2017). Plume‐subduction interaction forms
large auriferous provinces. Nature Communications, 8(1), 843–847. https://doi.org/10.1038/s41467‐017‐00821‐z
Telford, W. M., Geldart, L. P., & Sheriff, R. E. (1990). Applied geophysics (2nd ed.). Cambridge: Cambridge University Press. Retrieved from
http://www.cambridge.org/us/academic/subjects/earth‐and‐environmental‐science/solid‐earth‐geophysics/applied‐geophysics‐2nd‐
edition
Thiel, S., & Heinson, G. (2010). Crustal imaging of a mobile belt using magnetotellurics: An example of the Fowler Domain in South
Australia. Journal of Geophysical Research, 115, B06102. https://doi.org/10.1029/2009JB006698
Thiel, S., & Heinson, G. (2013). Electrical conductors in Archean mantle‐Result of plume interaction? Geophysical Research Letters, 40,
2947–2952. https://doi.org/10.1002/grl.50486
Thiel, S., Heinson, G., & White, a. (2005). Tectonic evolution of the southern Gawler Craton, South Australia, from electromagnetic
sounding. Australian Journal of Earth Sciences, 52(6), 887–896. https://doi.org/10.1080/08120090500304281
Thiel, S., Soeffky, P., Krieger, L., Regenauer‐Lieb, K., Peacock, J., & Heinson, G. (2016). Conductivity response to intraplate deformation:
Evidence for metamorphic devolatilization and crustal‐scale fluid focusing. Geophysical Research Letters, 43, 11,236–11,244. https://doi.
org/10.1002/2016GL071351
Thomas, J. L., Direen, N. G., & Hand, M. (2008). Blind orogen: Integrated appraisal of multiple episodes of Mesoproterozoic deformation
and reworking in the Fowler Domain, western Gawler Craton, Australia. Precambrian Research, 166(1–4), 263–282. https://doi.org/
10.1016/j.precamres.2007.05.006
Turner, S., Foden, J., Sandiford, M., & Bruce, D. (1993). Sm‐Nd isotopic evidence for the provenance of sediments from the Adelaide fold
belt and southeastern Australia with implications for episodic crustal addition. Geochimica et Cosmochimica Acta, 57(8), 1837–1856.
https://doi.org/10.1016/0016‐7037(93)90116‐E
Uieda, L., & Barbosa, V. C. F. (2017). Fast nonlinear gravity inversion in spherical coordinates with application to the South American
Moho. Geophysical Journal International, 208(1), 162–176. https://doi.org/10.1093/gji/ggw390
Uvarova, Y. A., Pearce, M. A., Liu, W., Cleverley, J. S., & Hough, R. M. (2017). Geochemical signatures of copper redistribution in IOCG‐
type mineralization, Gawler Craton, South Australia. Mineralium Deposita, 53(4), 477–492. https://doi.org/10.1007/s00126‐017‐0749‐1
Van der Meijde, M., Julià, J., & Assumpção, M. (2013). Gravity derived Moho for South America. Tectonophysics, 609, 456–467. https://doi.
org/10.1016/j.tecto.2013.03.023
Van der Meijde, M., Pail, R., & Bingham, R. (2015). Introduction to the special issue on GOCE Earth science applications and models.
International Journal of Applied Earth Observation and Geoinformation, 35(PA), 1–3. https://doi.org/10.1016/j.jag.2014.09.008
Wise, T., Reid, A. J., Jakica, S., Fabris, A. J., Wielen, S. E., Van Der Ziramov, S., & Pridmore, D. (2015). Olympic Dam seismic revisited:
reprocessing of deep crustal seismic using partially preserved amplitude processing. MESA, 78(3), 17–28.

MOTTA ET AL. 3104

You might also like