You are on page 1of 197

Record 2016/07 | GeoCat 90096

Geodynamic Synthesis of the


Phanerozoic of eastern Australia
Second Edition

Revised and updated by Champion, D.C., from Champion, D. C., Kositcin, N., Huston, D. L., Mathews, E.
and Brown, C (2009).

APPLYING GEOSCIENCE TO AUSTRALIA’S MOST IMPORTANT CHALLENGES www.ga.gov.au


Geodynamic Synthesis of the Phanerozoic
of eastern Australia
Second Edition

GEOSCIENCE AUSTRALIA
RECORD 2016/07

Revised and updated by Champion, D.C., from Champion, D. C., Kositcin, N., Huston, D. L., Mathews,
E. and Brown, C (2009).
Department of Industry, Innovation and Science
Minister for Resources, Energy and Northern Australia: The Hon Josh Frydenberg MP
Assistant Minister for Science: The Hon Karen Andrews MP
Secretary: Ms Glenys Beauchamp PSM

Geoscience Australia
Chief Executive Officer: Dr Chris Pigram
This paper is published with the permission of the CEO, Geoscience Australia

© Commonwealth of Australia (Geoscience Australia) 2016

With the exception of the Commonwealth Coat of Arms and where otherwise noted, this product is
provided under a Creative Commons Attribution 4.0 International Licence.
(http://creativecommons.org/licenses/by/4.0/legalcode)

Geoscience Australia has tried to make the information in this product as accurate as possible.
However, it does not guarantee that the information is totally accurate or complete. Therefore, you
should not solely rely on this information when making a commercial decision.

Geoscience Australia is committed to providing web accessible content wherever possible. If you are
having difficulties with accessing this document please email clientservices@ga.gov.au.

ISSN 2201-702X (PDF)

ISBN 978-1-925297-01-0 (PDF)

GeoCat 90096

Bibliographic reference: Champion, D. C., 2016. Geodynamic Synthesis of the Phanerozoic of


eastern Australia. Second Edition. Record 2016/07. Geoscience Australia, Canberra.
http://dx.doi.org/10.11636/Record.2016.007
Contents

Executive Summary..................................................................................................................................1
1 Introduction ............................................................................................................................................3
1.1 Orogens versus orogenies ...............................................................................................................6
1.2 Eastern Australia Orogenic Zones and tectonic periods – definitions .............................................6
1.3 Time Scale definition and time and space terminology ...................................................................8
2 Geological and geodynamic syntheses for eastern Australia ...............................................................9
2.1 Lachlan Orogen and related margins ..............................................................................................9
2.1.1 Introduction .................................................................................................................................9
2.1.2 Late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma): Rodinian break-up .......12
2.1.3 Early to latest Cambrian (ca. 515 Ma to ca. 490 Ma): Delamerian Orogeny ...........................18
2.1.4 Latest Cambrian to early Silurian (ca. 490 Ma to ca. 430 Ma) .................................................23
2.1.5 Middle Silurian to Late Devonian (ca. 430 Ma to ca. 380 Ma) .................................................31
2.1.6 Late Devonian to early Carboniferous (ca. 380 to ca. 350 Ma) ...............................................36
2.1.7 Middle Carboniferous to earliest Triassic (ca. 350 Ma to 250 Ma)...........................................38
2.2 North Queensland region ...............................................................................................................39
2.2.1 Introduction ...............................................................................................................................39
2.2.2 Late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma): Rodinian break-up .......39
2.2.3 Early to latest Cambrian (ca. 515 to 490 Ma)...........................................................................42
2.2.4 Latest Cambrian to early Silurian (ca. 490 Ma to ca. 430 Ma) .................................................43
2.2.5 Middle Silurian to Late Devonian (ca. 430 Ma to ca. 380 Ma) .................................................50
2.2.6 Late Devonian to early Carboniferous (ca. 380 to ca. 350 Ma) ...............................................53
2.2.7 Middle Carboniferous to earliest Triassic (ca. 350 Ma to 250 Ma)...........................................55
2.3 New England Orogen.....................................................................................................................58
2.3.1 Introduction ...............................................................................................................................58
2.3.2 Late Neoproterozoic to latest Cambrian (ca. 600 to ca. 490 Ma) ............................................60
2.3.3 Latest Cambrian to early Silurian (ca. 490–430 Ma) ................................................................62
2.3.4 Middle Silurian to late Devonian (ca. 430–380 Ma): intra-oceanic island arc ..........................69
2.3.5 Late Devonian to Late Triassic overview (380–220 Ma) ..........................................................70
2.3.6 Late Devonian to late Carboniferous (ca. 380–305 Ma): continental arc .................................71
2.3.7 Late Carboniferous to late Permian (ca. 305 to 265 Ma): back-arc extension and
orocline formation ..............................................................................................................................76
2.3.8 Late Permian to mid Triassic (ca. 265 Ma to ca. 230 Ma): Hunter-Bowen Orogeny ...............80
2.4 Thomson Orogen, cover basins, and Koonenberry Belt................................................................82
2.4.1 Introduction ...............................................................................................................................82
2.4.2 Late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma): Rodinian breakup .........83
2.4.3 Early to latest Cambrian (ca. 515 Ma to 490 Ma).....................................................................89
2.4.4 Latest Cambrian to early Silurian (ca. 490 Ma to 430 Ma) .......................................................92
2.4.5 Middle Silurian to late Devonian (ca. 430 Ma to 380 Ma) ........................................................96
2.4.6 Late Devonian to early Carboniferous (ca. 380 Ma to ca. 350 Ma) .........................................99
2.4.7 Early Carboniferous to late Permian (ca. 350 Ma to ca. 265 Ma) ..........................................102
2.4.8 Late Permian to late Triassic (ca. 265 Ma to 230 Ma) ...........................................................104
3 Regional overview of the tectonic development of eastern Australia in the Phanerozoic .................107
3.1 Introduction ..................................................................................................................................107

Geodynamic Synthesis of the Phanerozoic of eastern Australia iii


3.2 Tectonic summary of eastern Australia by time period ................................................................114
3.2.1 Late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma): Rodinian breakup .......114
3.2.2 Early to latest Cambrian (ca. 515 Ma to 490 Ma)...................................................................121
3.2.3 Latest Cambrian to early Silurian (ca. 490 Ma to 430 Ma) .....................................................129
3.2.4 Middle Silurian to Late Devonian (ca. 430 Ma to ca. 380 Ma) ...............................................139
3.2.5 Late Devonian to early Carboniferous (ca. 380 to ca. 350 Ma) .............................................152
3.2.6 Early Carboniferous to late Permian (ca. 350 Ma to ca. 265 Ma) ..........................................157
3.2.7 Late Permian to late Triassic (ca. 265 Ma to 230 Ma) ...........................................................163
4 References ........................................................................................................................................169
Acknowledgements ..............................................................................................................................191

iv Geodynamic Synthesis of the Phanerozoic of eastern Australia


Executive Summary

This report presents the second edition of a ‘Geodynamic Synthesis of the Phanerozoic of Eastern
Australia’, the first edition being published in 2009. The original Record was undertaken with two aims:
(1) to better understand the tectonic and geodynamic setting of known mineral deposits within eastern
Australia, and (2) to provide a predictive capability for extending potential regions of known
mineralisation and delineating regions with potential for new mineral styles and commodities. This was
achieved by synthesising geological and metallogenic data on a regional, largely orogenic, basis. This
allowed the identification of geological and metallogenic events and geodynamic cycles, and the
production of regional geological syntheses and accompanying time-space-event plots. These
regional syntheses were used to produce an interpreted geological, metallogenic and geodynamic
synthesis of eastern Australia. This synthesis provided the geodynamic framework to constrain known
mineralisation and allow a predictive capability for potential new mineralisation.

Since the first edition, significant new geological information and tectonic models have appeared for
most parts of eastern Australia, along with the definition and/or redefinition of many geological
provinces, sub-provinces and igneous associations. This new information, tectonic models and names
have been incorporated into this second edition.

Like the first edition, this new edition has concentrated on the Tasmanides or Tasman Orogen or
Tasman Orogenic Belt of eastern Australia (Figure 1.1, Figure 1.2). This essentially corresponds to the
Palaeozoic and early Mesozoic rocks east of the old Gondwanan (Delamerian) continental margin. A
large focus of this second edition has been to update the geology of the Thomson Orogen and
adjacent regions. This includes an enhanced discussion of the geology of the Koonenberry region and
the Warburton Basin and their implications for the Thomson Orogen. Section 3, ‘Metallogenic Events
in Phanerozoic Eastern Australia’, from the first edition has not been updated, and has not been
reproduced in the new edition.

The geology and tectonic development of eastern Australia, particularly the Tasman Orogen has been
the focus of numerous studies, with a voluminous literature, including numerous orogen-based or
more regional reviews. This research has led to a plethora of tectonic models with perhaps the
majority of differences focussed on the Lachlan Orogen. Importantly, however, there is a general
consensus that since the late part of the Neoproterozoic, eastern Australia has, broadly, been in three
fundamental tectonic states:

• Rodinian-breakup (rifting) and ensuing passive margin (late Neoproterozoic)––in its simplest form,
the opening of the Pacific Ocean and development of the Australian component of the Gondwana
margin (e.g., Li and Powell, 2001; Direen and Crawford, 2003; Cawood, 2005).
• Alternating extension and compression in an overall convergent arc environment, commencing
during the early Cambrian and continuing through to the Mesozoic. This resulted in accretionary
growth, expressed as the Tasman Orogen (or Tasman Orogenic Belt) and its components (the
Lachlan, Thomson, Mossman and New England orogens). A number of orogenic events are
recorded, from the Delamerian Orogeny (Cambrian) through to the Hunter-Bowen Orogeny
(Permian-Triassic). The Tasman Orogen was effectively terminated at ca. 230 Ma with the
stepping out of the plate margin to the east following the Hunter-Bowen Orogeny, with ensuing
extension leading to subsequent breakup.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 1


• Rifting and passive margin (± hotspot activity), related to rifting of crustal fragments and opening of
ocean basins, and especially related to break-up in the late Mesozoic to early Cenozoic of the
Australian component of the Gondwana supercontinent.

Although broadly simple, in detail the tectonic evolution of eastern Australia is clearly complex. Not
only is there evidence for diachronous events, and possible arc switches, it is also clear that the
current configuration of eastern Australian geological provinces (especially Palaeozoic to early
Mesozoic components) may, at least in part, represent an amalgamation of allochthonous and
autochthonous terranes. Controversy and uncertainty involves: the original positions of potentially
allochthonous blocks, such as the Selwyn Block; the role of strike-slip or other lateral movement,
including orocline formation; the presence of domains, such as the Melbourne Zone, which are
missing evidence for major deformation events; as well as the nature of possible oceanic arc remnants
and whether they represent island arcs, continental arcs, or magmatic arcs at all. Another area of
controversy concerns the actual positions of, and number of, arcs, as best exemplified in the
Ordovician and Silurian rocks of the Lachlan Orogen, but also in the New England Orogen. Equally
important controversies concern the actual location of arcs and discriminating between arc-forearc and
backarc environments, e.g., the interpretation of the Hodgkinson Province in northern Queensland.

Much debate regarding tectonic reconstructions for eastern Australia centres around the Lachlan
Orogen, which has, numerous features including its width, the presence of interpreted oceanic arc
remnants (Macquarie Arc) interspersed between Ordovician turbidite packages, and variable
deformation, that are not easy to explain by simple models. Further uncertainty concerns the Thomson
Orogen, particularly the nature and age of the basement, and its relationship to the other orogens,
particularly with the Lachlan Orogen. The orogen is largely undercover, and as such poorly
understood.

It is also evident that there are significant differences between the southern and northern parts of the
Tasman Orogen, and so models for the Lachlan and Thomson orogens also have to be able to explain
the geology of the north Queensland region. The latter region has some unique advantages to help
unravel the tectonic development of the Tasman orogen. In particular the north Queensland region
has both Proterozoic basement rocks and rocks spanning the age range of the Tasman Orogen
exposed in a condensed section, i.e., successive continental margins largely developed on top of, or
close to, each other through the Paleozoic. This is best exemplified in the Townsville–Greenvale
region.

In this report the geology and interpreted geodynamic environments are discussed by time periods,
from the Late Neoproterozoic to the Middle Triassic, in two parts. Geological summaries and time-
space-event plots are presented for each orogen in Part 1. Interpreted geological and geodynamic
synthesis for the whole Tasman Orogen, are presented in Part 2, including discussion of the range of
geodynamic models that have been suggested in the literature, including major problems and
difficulties with all tectonic models and areas of uncertainty. We also present one possible tectonic
interpretation.

2 Geodynamic Synthesis of the Phanerozoic of eastern Australia


1 Introduction

DC Champion, updated from DC Champion and N Kositcin (2009)

This report presents the second edition of a ‘Geodynamic Synthesis of the Phanerozoic of Eastern
Australia’; the first edition being published in 2009 (Champion et al., 2009). The original Record was
undertaken with the aims: (1) to better understand the tectonic and geodynamic setting of existing
mineral deposits within eastern Australia, and (2) to provide a predictive capability, within the
synthesised geodynamic framework, for extending potential regions of known mineralisation and for
delineating regions with potential for new mineral styles and commodities. That Record combined the
Mineral Systems approach of Wyborn (1997) with the ‘Five Questions’ methodology adopted by the
Predictive Mineral Discovery Cooperative Research Centre (pmd*CRC;
http://www.pmdcrc.com.au/RESprograms.html; Barnicoat, 2007, 2008). It was targeted at the first of
the ‘Five Questions’, namely, constraining and understanding the regional and local geodynamic
environment as the first step in delineating mineral systems. This was achieved by synthesising
geological and metallogenic data on a regional, largely orogenic, basis. It allowed for the identification
of geological and metallogenic events and geodynamic cycles, and the production of regional
geological syntheses and accompanying time-space-event plots. These regional syntheses were used
to produce an interpreted geological, metallogenic and geodynamic synthesis of eastern Australia.
This synthesis provided the geodynamic framework to constrain known mineralisation and allow a
predictive capability for potential new mineralisation. Outputs were delivered in three parts: geological
summaries and time-space-event plots; interpreted geological and geodynamic synthesis; and known
and predicted metallogeny for eastern Australia.

Since the first edition significant new geological information has appeared in a number of publications,
including the Koonenberry region report and related papers (Greenfield et al., 2010, 2011), the
recently released Queensland Geology (Jell, 2013) with updated information for the New England,
Thomson and Mossman (called North Queensland in the first edition) orogens, new geological
information for the northern Thomson (Purdy et al., 2013), the southern Thomson and adjoining
Lachlan Orogen (e.g., Glen et al., 2013), new information and tectonic models for the Tasman Orogen,
especially the Lachlan Orogen (e.g., Gibson et al. 2011, 2015; Cayley, 2011; Cayley et al., 2011;
Fergusson, 2010; Fergusson et al., 2013; Moresi et al., 2014; Quinn et al., 2014), as well as a new
synthesis of the Tasman Orogen by Glen (2013). One result from the Queensland Geology volume
(Jell, 2013) has been the definition and/or redefinition of many geological provinces, sub-provinces
and igneous associations for that state. All these names have been adopted for this report.

Like the first edition, this new edition has concentrated on the Tasmanides or Tasman Orogen or
Tasman Orogenic Belt (Scheibner and Veevers, 2000; Veevers, 2000, 2004; Cawood, 2005; Glen,
2005) of eastern Australia (Figure 1.1, Figure 1.2). This essentially corresponds to the Palaeozoic and
early Mesozoic rocks east of the old Gondwanan (Delamerian) continental margin. A large focus of
this second edition has been to update the geology of the Thomson Orogen and local regions
surrounding that orogen. This includes greater discussion of the geology of the Koonenberry region
and the Warburton Basin and implications for the Thomson Orogen. Section 3, ‘Metallogenic Events in
Phanerozoic Eastern Australia’, from the first edition has not been updated, and has not been
reproduced in the new edition.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 3


Figure 1.1 Location map showing the distribution of eastern Australian orogens that form part of the Tasman
Orogen. Orogen names follow Glen (2005) and Jell (2013). Orogen boundaries from Glen (2005, 2013),
VandenBerg et al. (2000), Seymour and Calver (1995), Bain and Draper (1997), Jell (2013), and unpublished GA-
GSQ Nd isotope data for the eastern Thomson Orogen. The Bowen-Gunnedah-Sydney Basin outline is from
Geoscience Australia’s ‘Basins’ national data set.

4 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 1.2 Location map showing the distribution of eastern Australian orogens, basins and other regions. Orogen
names and boundaries follow Glen (2005), Bain and Draper (1997), and unpublished GA-GSQ Nd isotope data
for the eastern Thomson Orogen. The boundary of the Thomson Orogen follows Jell (2013) and extends north to
include the Greenvale, Cape River and Barnard provinces. The Koonenberry Belt is from Greenfield et al. (2011).
The Warburton, Georgina, Cooper, Darling, Tasmania, Cobar, Adavale, Galilee, Lakefield, Bowen-Gunnedah-
Sydney, Stansbury, and Eromanga and related basin outlines are from Geoscience Australia’s Provinces
database. The Belyando, Drummond, Timbury Hills basins, Warrabin Trough, and Anakie Inlier outlines are from
Jell (2013). The Selwyn Block delineates the area thought to have Proterozoic basement underlying the Lachlan
Orogen (e.g., Cayley et al., 2002, 2011).The Hay-Booligal Zone is another such possible block (Glen, 2013).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 5


1.1 Orogens versus orogenies
We use the term ‘orogen’ (e.g., Lachlan Orogen) simply to designate an orogenic province. It is used
in a similar sense to the historical use of fold belt (e.g., Lachlan Fold Belt). ‘Orogeny’ on the other
hand refers to an orogenic event, typically accompanied by deformation, metamorphism and
magmatism (e.g., Delamerian Orogeny). A number of major orogenies have occurred within the
eastern Australian orogens.

1.2 Eastern Australia Orogenic Zones and tectonic periods –


definitions
The geology and tectonic development of the eastern Australia, particularly the Phanerozoic
component––the Tasmanides––has been the focus of numerous past and continuing studies. This has
resulted in a voluminous literature including numerous orogen-based or more regional reviews (e.g.,
Murray, 1986, 1990; Murray et al., 1987; Coney, 1992; Seymour and Calver, 1995; Bain and Draper,
1997; Gray, 1997; Gray and Foster, 1997; Gray et al., 1997; Scheibner and Basden, 1998;
VandenBerg et al., 2000; Veevers, 2000; Cawood, 2005; Crawford et al., 1984, 2003a; Glen, 2005,
2013; Cayley et al., 2011; Glen et al., 2013). The focus of this research has had two important
outcomes. The first has been the recognition that eastern Australia, or more specifically the Tasman
Orogen (also Tasmanides and Tasman Orogenic Belt, TOB), can be broadly subdivided into a
relatively small number of, largely Paleozoic (and Mesozoic), complex and in part composite,
orogens––Lachlan, Thomson, New England and Mossman (called North Queensland in the first
edition of the report). These are located east of the more contiguous Proterozoic, and older, Australian
continent (e.g., Gray, 1997; Cawood, 2005). The latter is often interpreted to represent the post-
Rodinia breakup margin, corresponding to the old notion of the ‘Tasman Line’. The location and nature
of the boundary between rocks of the Tasman Orogen and the post-Rodinian margin is problematical,
however.

As summarised by Direen and Crawford (2003), apart from areas such as northern Queensland,
where the Tasman boundary is clear, for most of its extent the exact location of the Tasman Line is
enigmatic, and numerous variants of the line have been suggested. This is perhaps not surprising
given the potential complexity of old craton boundaries and subsequent reworking, and both Direen
and Crawford (2003) and Glen (2005) suggested the term was essentially meaningless, particularly in
southern Australia. Regardless of this, it is clear that a boundary, albeit transitional, exists. Where this
actually is depends in part on the definitions adopted (e.g., see discussion by Glen, 2005). For the
purpose of this report, we have taken the middle ground. We have largely concentrated on those rocks
east of the Delamerian Orogen, and the states these occur in, that is, Queensland, New South Wales,
Victoria and Tasmania. We have, however, also taken in consideration Rodinian break-up and
Delamerian Orogeny geology, but largely only where it occurs in the eastern States. We have not
considered in any detail the Tasmanides geology in other Australian states; though we do refer to it
where it is of significance, in particular the Delamerian Orogen in South Australia. Cawood (2005) has
grouped the Tasman Orogenic Zone and the Delamerian Orogen, and its continuation throughout
Gondwana, into the Terra Australis Orogen.

The definition of individual orogens and the subdivisions of these (zones, elements, regions, terranes,
superterranes, and provinces) are discussed in detail at the beginning of each section. We have, as
far as applicable, followed the boundaries used by the relevant state surveys (e.g., Seymour and
Calver, 1995; VandenBerg et al., 2000; Glen, 2005, 2013; Bain and Draper, 1997; and Jell, 2013).

6 Geodynamic Synthesis of the Phanerozoic of eastern Australia


The second important outcome of research, primarily since the 1980’s, is captured in the time-space-
event plots produced as part of this work and that is the recognition that there are broadly
contemporaneous orogenic events recorded in all the orogens. This has been recognised before, e.g.,
the ‘stages’ recognised by Scheibner and Basden (1996, 1998) and Korsch and Harrington (1987) in
NSW and the New England orogen, respectively. Glen (2005), amongst others, extended this concept
to define tectonic cycles for all the Tasmanides, with a cycle defined as the period immediately after
the previous orogeny, through to the next major orogeny. Cycles were named after the terminal
orogeny used, e.g., Benambran Cycle records the period from post-Delamerian Orogeny to the
Benambran Orogeny.

There are some potential difficulties with the cycle approach. These include the observation that
timings of orogenies are not contemporaneous across the Tasman Orogen––it is evident that the
timing(s) of the terminal orogeny in each cycle may be slightly different between orogens and even
within orogens. Although some of this in part reflects poor age constraints, it would appear that it is, in
part, real, and reflects the multiple stages of long-lived and possibly diachronous orogenies, such as
the Benambran Orogeny in Victoria (e.g., VandenBerg et al., 2000). Additional concerns include
orogeny names for multiple events, the best example being the Hunter-Bowen Orogeny which extends
some 35 million years (e.g., ~265 Ma to ~230 Ma; Holcombe et al., 1997b; Korsch et al., 2009b).
Whether this should be considered as single or multiple tectonic cycles is a debatable question.
Thirdly, within each cycle, there may be additional, usually minor orogenies, such as the Bindian
Orogeny in the Tabberabberan Cycle (e.g., Gray, 1997; Glen, 2005). There are also problems with
regional correlations. For example, the multiple orogenies of the Alice Springs Orogeny (450-300 Ma)
coincide reasonably well with the major orogenies of eastern Australia (e.g., Benambran,
Tabberabberan, Kanimblan). In areas potentially affected by both Alice Springs and eastern Australian
orogenies, e.g., the Koonenberry region (e.g., Figure 2.19), is it strictly valid to place rocks there into
cycles based on Tasman Orogen events? In north Queensland, for example, it appears that there may
be slight timing differences between the Kanimblan and Alice Springs orogenies, and both names
have been used (Section 2.2).

It is also evident that there is no reason why cycles or orogenies should actually be synchronous
within orogens let alone across all of eastern Australia (as shown elegantly in the Moresi et al. (2014)
model). Recognition of such cycles, however, has historically provided a methodology that has
allowed a more unified discussion of eastern Australian in a tectonic framework and recognition of
commonalities within and between orogens, and it was the approach we used in the first edition
(Champion et al., 2009). In the second edition, we have followed a modified approach, abandoning the
cycle nomenclature and using a more strictly time-based comparison, with the recognition that the
periods used reflect the timing of orogenies within the Lachlan Orogen. Hence, the geological
syntheses for regions and for eastern Australia presented herein have been documented on the basis
of time periods that correspond broadly to the Delamerian, Benambran, Tabberabberan, Kanimblan
and Hunter-Bowen Tectonic cycles of other workers. The modified approach allows us to not only
highlight commonalities across orogens but also highlight the differences within and between cratons
for the relevant regions and tectonic periods.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 7


1.3 Time Scale definition and time and space terminology
Time scale usage (Era, Period and Epoch nomenclature and boundary ages) (with the exceptions as
outlined below) follow the recommendations of the International Commission on Stratigraphy, as
published in August 2012 (http://www.stratigraphy.org/ICSchart/ChronostratChart2012.jpg). Apart from
minor changes in the first decimal place of some Permian Stages and the slight shifting of the top of
the Middle Triassic (from 237 Ma to ~235 Ma), ages are identical to the current 2015/01 version. The
only variations from the accepted nomenclature usage in this report are as follows:

• Early/Lower Permian instead of Cisuralian


• Late/Upper Permian instead of Guadalupian and Lopingian
• Early/Lower Carboniferous instead of Mississippian
• Late/Upper Carboniferous instead of Pennsylvanian
• Early/Lower Silurian instead of Llandovery and Wenlock
• Late/Upper Silurian instead of Ludlow and Pridoli
• Early/Lower Cambrian instead of Terreneuvian and Epoch 2
• Middle Cambrian instead of Epoch 3
• Late/Upper Cambrian instead of Furongian

Note that Lower, Middle and Upper are essentially equivalent to Early, Middle and Late, though
following the recommendations of Owen (2009), EML should be restricted to time, events etc, LMU to
locations in space. We have followed the latter recommendation and Early, Middle and Late are
shown on all time-space plots and used throughout the text in reference to time. When used as part of
time scale terminology, e.g., Early Devonian, the prefixes Early, Middle and Late are shown
capitalised regardless of whether they refer to the formal usage as listed by the International
Commission on Stratigraphy, or informally, as in the exceptions listed above. Where used more
generally (e.g., typically more non-specific time periods) the lower-case prefixes early, middle and late
have been used. More unambiguous text, e.g., ‘lower part of the Devonian’, has been used as much
as possible to avoid confusion.

8 Geodynamic Synthesis of the Phanerozoic of eastern Australia


2 Geological and geodynamic syntheses for
eastern Australia

by DC Champion, updating N Kositcin, DC Champion, E Mathews and C Brown (2009)

2.1 Lachlan Orogen and related margins


by DC Champion and N Kositcin

2.1.1 Introduction
The Lachlan Orogen, as used here, follows the general usage of numerous authors (e.g., Seymour
and Calver, 1995, 1998; Scheibner and Basden, 1996; Gray, 1997; Gray and Foster, 1997, 2004;
VandenBerg et al., 2000; Crawford et al., 2003a, b; Glen, 2005; Glen et al., 2007b), as shown in
Figure 1.1. The Orogen occurs within the central and eastern parts of New South Wales, Victoria and
northeastern Tasmania (Figure 1.1). It is bound to the west by the Delamerian Orogen, to the north by
the Thomson Orogen, and to the east by young oceanic crust of the New England Orogen. The
Lachlan Orogen is traditionally interpreted as those regions not having undergone the Delamerian
Orogeny (e.g., Glen, 2005). Included within the Orogen, however, are rocks which may be underlain
by Delamerian crust, e.g., the Selwyn Block of Cayley et al. (2002). The Delamerian-Lachlan boundary
is poorly defined in western New South Wales (e.g., compare Glen, 2005 and Glen, 2013), largely
because of younger cover rocks. Similarly, some conjecture concerns the western boundary of the
Orogen in Victoria. This is traditionally taken as the east-dipping Moyston Fault on the western side of
the Stawell Zone (e.g., VandenBerg et al., 2000; Crawford et al., 2003b), and results from the recent
seismic survey (Cayley et al., 2011) appear to confirm that the Moyston Fault does represent the
western boundary of the orogen. Work by Miller et al. (2005), however, has shown that rocks of the
western Stawell Zone have experienced Delamerian-aged deformation, indicating they are, at least
locally, part of the Delamerian Orogen. The western Stawell Zone, therefore, is probably best thought
of as a partly transitional boundary between the Lachlan and Delamerian Orogens, though is treated
here as being part of the former.

The Lachlan Orogen itself has been subdivided in a number of ways, usually largely based on state
lines. The geological surveys of both Victoria and Tasmania have each subdivided their state (and the
Lachlan Orogen component) into geological regions – called zones in Victoria (VandenBerg et al.,
2000; Figure 2.2) and elements in Tasmania (Seymour and Calver, 1995, 1998; Figure 2.3). The
Lachlan Orogen within each state has also been subdivided into terranes and superterranes. These
include the Whitelaw and Benambra Terranes in Victoria (e.g., VandenBerg et al., 2000; Figure 2.2;
though not used by Cayley et al., 2011), and the Bega, Narooma, Girilambone-Wagga and Macquarie
Arc terranes in NSW (e.g., Glen, 2005; Glen et al., 2007b; Figure 2.1). Glen and co-workers also
introduced the Adaminaby Superterrane (Figure 2.1), for those regions of the Lachlan with Ordovician
turbidites, consisting of the Bega, and Girilambone-Wagga terranes in NSW and the Bendigo,
Tabberabbera and Omeo zones in Victoria (parts of VandenBerg et al.’s (2000) Whitelaw and
Benambra Terranes). We refer to most of these zones and terranes (Figure 2.1, Figure 2.2,
Figure 2.3) specifically when discussing relevant state geology. We, however, prefer the simpler
terminology, also used by Gray (1997), VandenBerg et al. (2000), Gray et al. (2003), Glen (2005), and

Geodynamic Synthesis of the Phanerozoic of eastern Australia 9


more recently Cayley et al. (2011) and Glen (2013), of Western, Central and Eastern Lachlan
(Figure 2.1). The Western Lachlan – largely synonymous with the Whitelaw Terrane (VandenBerg et
al., 2000; Figure 2.1, Figure 2.2) is largely known from Victoria, but does extend into New South
Wales (Glen, 2005, 2013). The Central and Eastern Lachlan occur in both Victoria and New South
Wales (Figure 2.1, Figure 2.2), though the majority of the Eastern Lachlan lies in New South Wales.
Tasmania is more problematical. We follow VandenBerg et al. (2000) and Cayley et al. (2002) in
adopting the Selwyn Block model (see below) which effectively links western Tasmania with the
Melbourne Zone (Figure 2.2). In this scenario northeastern Tasmania is also included with the
Melbourne Zone (Western Lachlan), though it shares similarities with both the Melbourne Zone and
the Central Lachlan (e.g., Reed, 2001).

Figure 2.1 Zones, terranes and Lachlan subprovinces of New South Wales and Victoria. Figure modified after
Fergusson (2003), Gray and Foster (2004), Glen (2005, 2013), Crawford et al. (2007a) and Glen et al. (2007b).
Note in the definition of Glen et al. (2007b) the Bendigo Zone is also included in the Adaminaby Superterrane.
Victoria zones follow nomenclature of VandenBerg et al. (2000). The Western Lachlan Orogen as used here
comprises the Stawell Zone and its interpreted northern extension into New South Wales, and the Bendigo,
Melbourne and Hay-Booligal zones (e.g., Cayley et al., 2011; Glen, 2013).

10 Geodynamic Synthesis of the Phanerozoic of eastern Australia


As outlined by Cayley et al. (2002), rocks of the Melbourne Zone probably formed upon Delamerian
Orogen crust. Although this is not universally accepted (e.g., Gray and Foster, 2004; Spaggiari et al.,
2004), it indicates that the Melbourne Zone, like the Stawell Zone, can be thought of as transitional
Delamerian-Lachlan. In the following discussion, although focused on the Lachlan Orogen, parts of the
Delamerian Orogen are also included, particularly western Victoria, and Western Tasmania (including
King Island). The Delamerian in South Australia, especially in the Stansbury Basin, is also referred to
where necessary (see Figure 2.7). The Koonenberry region and Warburton Basin are included with
the Thomson Orogen discussion. Numerous significant reviews have been presented on the Lachlan
Orogen, both on a state (e.g., Scheibner and Basden, 1996; Gray, 1997; VandenBerg et al., 2000;
Birch, 2003; Seymour and Calver, 1995) and orogen basis (e.g., Gray and Foster, 2004; Glen, 2005,
2013; Cayley et al., 2011). As a consequence although the general distribution of rock types is
described, most attention is focused on identifying crustal blocks and interpreted geodynamic
environments. Similarly, numerous time-space plots for the majority of the Lachlan, on a state or
orogen-wide basis, have been presented previously (e.g., VandenBerg et al., 2000; Birch, 2003; Glen,
2005, 2013; Pogson and Glen, 2006; Seymour and Calver, 1998). Summaries of these are also
presented here (Figure 2.4, Figure 2.5, Figure 2.6), largely based on those previously published, with
minor modifications where required.

Figure 2.2 Geological orogens, terranes and zones of Victoria and parts of New South Wales (not coloured).
Figure modified from VandenBerg et al. (2000). The Selwyn Block delineates the area thought to have
Proterozoic basement underlying the Melbourne Zone (e.g., Cayley et al., 2002). The Whitelaw Terrane is largely
synonymous with the Western Lachlan Subprovince (Figure 2.1), though the location of the northern boundary is
uncertain.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 11


2.1.2 Late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma):
Rodinian break-up

2.1.2.1 Delamerian Orogen

2.1.2.1.1 Geological and tectonic summary

The late Neoproterozoic (ca. 600 Ma) to mid Cambrian geological history of southeastern Australia
records a cycle of continental rifting and ocean opening related to the breakup of Rodinia, formation of
passive margins and initiation of the Pacific Ocean (e.g., Cawood, 2005). This continued in
southeastern Australia until it was effectively ended by subduction (occurring by at least ca. 515 Ma;
Foden et al., 2006) and accretionary tectonics related to the Delamerian Orogeny (e.g., Berry and
Crawford, 1988; Crawford and Berry, 1992; Cayley, 2011). Glen (2005) called this period the
Delamerian Cycle, and suggested that it lasted more than 300 Ma, back to ca. 830–780 Ma, and
possibly earlier. Most of this time period falls outside the range or geographic coverage of this report
and is not covered here in any detail (see Drexel and Preiss, 1995; Calver and Walter, 2000; Foden et
al., 2002b; Crawford et al., 2003a; Glen, 2005, 2013; and references therein for more information).

Figure 2.3 Tectonic elements of Tasmania. Figure modified from Seymour and Calver (1995, 1998). Eastern and
Western Tasmania terminology adopted from Spaggiari et al. (2003). The correlation of Tasmania with mainland
Australia is contentious. In the Selwyn Block model of Cayley et al. (2002), Tasmania is related to the Melbourne
Zone (Western Lachlan; Figure 2.1, Figure 2.2). East Tasmania is thought to correlate with either the Melbourne
Zone or the Tabberabbera Zone (e.g., Reed, 2001; see Figure 2.2).

12 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Neoproterozoic and Early Cambrian rocks occur in western Tasmania and King Island (Figure 2.6,
Figure 2.9), in the Glenelg and Grampians-Stavely Zones in Western Victoria (Figure 2.2, Figure 2.5),
and possible extensions into the Koonenberry region of northwestern New South Wales (Crawford et
al., 1997; Greenfield et al., 2010, 2011), and are extensively developed in the Delamerian Orogen in
South Australia (Figure 2.7). They include glacial-derived sediments (ca. 700 and 635 Ma; Calver and
Walter, 2000), ca. 780–725 Ma and ca. 600–580 Ma rift-related mafic magmatism (Crawford and
Berry, 1992, Crawford et al., 1997; Holm et al., 2003; Foden et al., 2002b), and the extensive, largely
marine, sedimentation of the Stansbury Basin (Drexel and Preiss, 1995; VandenBerg et al., 2000;
Crawford et al., 2003b; Figure 2.7), which also contains ca. 525 Ma rift-related magmatism. Rocks of
this age also include ultramafics and mafic rocks, interpreted as either oceanic floor and/or supra-
subduction zone remnants (Crawford et al., 2003a).

Widespread evidence for continental breakup in southeastern Australia, related to Rodinian rifting
(e.g., Cawood, 2005; Crawford et al., 2003a), is well recorded in rocks ca. 600 Ma in age and younger,
in western Tasmania and King Island (e.g., Calver and Walter, 2000; Calver et al., 2004; Meffre et al.,
2004), in South Australia (e.g., Drexel and Preiss, 1995; Foden et al., 2001), in western Victoria
(VandenBerg et al., 2000; Crawford et al., 2003b), and in the Koonenberry region, western New South
Wales (e.g., Crawford et al., 1997; Gilmore et al., 2007; Greenfield et al., 2010, 2011). As summarised
by Crawford et al. (1997; 2003a, b) many rocks of this age contain alkaline and/or tholeiitic
assemblages consistent with rift tectonics and a passive margin and mantle-plume magmatism.
Crawford et al. (2003a) suggested the extension was oriented largely northwest-southeast to explain
the distribution of rift volcanism at this time.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 13


Figure 2.4 Time-space plot for the New South Wales part of the Lachlan Orogen. Figure modified from, and based largely on, the time-space plots in the East Lachlan Orogen GIS (Pogson and Glen, 2006), with updates from Percival et al. (2011), Prendergast and
Offler (2012) and Quinn et al. (2014). Note: not all geological groups and units are shown (due to north-south geological variability).

14 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 2.5 Time-space plot for Victoria. Modified from, and based largely on, the time-space plot of VandenBerg et al. (2000). Refer to Figure 2.2 for location of zones. A simplified (and updated) version of this is presented in Figure 2.7.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 15


Figure 2.6 Time-space plot for Tasmania. Modified from, and based largely on, the time-space plot of Seymour and Calver (1998), with modifications by G. Green, J. Everard, C. Calver and M. McClenaghan (Mineral Resources Tasmania, pers. comm., 2008).
Refer to Figure 2.3 for location of elements.

16 Geodynamic Synthesis of the Phanerozoic of eastern Australia


2.1.2.1.2 Geological history and Time-Space plot explanation (Figure 2.1 to Figure 2.9)

• Neoproterozoic mafic magmatism in western Tasmania thought to be rift-related (Holm et al., 2003;
Foden et al., 2002b). Ages poorly constrained, probably ca. 725 Ma, but possibly ca. 780 Ma and
older.
• Deformation in western Tasmania––the Wickham deformation––constrained by granites (King
Island and northwest Tasmania) to have occurred at ca. 777 Ma and 760 Ma respectively (Turner
et al., 1998).
• Glacially-derived ca. 700 Ma, ca. 640 Ma and ca. 582–575 Ma sediments in northwest Tasmania
(Calver and Walters, 2000; Calver et al., 2004; Kendall et al., 2007).
• ca. 600–570 Ma shelf and rift successions (including glacial-derived sediments) in most elements
of Western Tasmania (Seymour and Calver, 1995; 1998; Meffre et al., 2004), including the Togari,
Success Creek and Weld River Groups. Many of these, such as the Crimson Creek Formation,
contain mafic tholeiitic magmatic rocks including picrites, consistent with rifting and a mantle plume
(Crawford and Berry, 1992, Crawford et al., 2003a).
• Deposition of Cambrian sediments of the Moralana Supergroup in the Glenelg Zone, Victoria
(VandenBerg et al., 2000; Crawford et al., 2003b), and more extensively in South Australia (Drexel
and Preiss, 1995; Figure 2.7), as part of the Stansbury Basin. In the Glenelg Zone, sediments are
dominantly deep-water turbidites, but also include mafic magmatic rocks (VandenBerg et al., 2000;
Crawford et al., 2003b). Rift-related, within-plate and more oceanic MORB-like magmatism occurs
within the Moralana Supergroup, in both South Australia and western Victoria—the Truro Volcanics
and correlatives (Drexel and Preiss, 1995, Belperio et al., 1998; VandenBerg et al., 2000; Foden et
al., 2002b; Crawford et al., 2003b; Figure 2.7). Most of these appear to have ages of ca. 525 Ma
though, Gibson et al. (2015) suggested some of these mafic rocks and the enclosing sediments
may be as old as Neoproterozoic sequences such as in the Koonenberry region to the north.
Deposition of the Stansbury Basin appears to have ended by ca. 514 Ma, based on granite
emplacement ages (Foden et al., 2006).
• Fault-bounded, mafic-ultramafic complexes in western Victoria (VandenBerg et al., 2000; Crawford
et al., 2003b) and in western Tasmania, interpreted as oceanic floor and/or supra-subduction zone
remnants (e.g., Crawford and Keays, 1987; Crawford et al., 1984, 2003b). These mafic-ultramafic
complexes either pre-date (VandenBerg et al., 2000), or are contemporaneous with, early
Delamerian deformation (e.g., ca. 510 Ma age for the Heazlewood Ultramafic Complex, Tasmania;
Turner et al., 1998). Gibson et al. (2015) have suggested some of the ultramafic bodies may
represent underlying mantle emplaced during hyper-extended margin processes.

2.1.2.2 Lachlan Orogen

Late Neoproterozoic and Early (to Late) Cambrian rocks also occur within the Lachlan Orogen
(Figure 2.8). These fall into two main categories:

• Mafic and ultramafic Cambrian (and older?) igneous rocks, preserved along major faults,
especially those that delineate zone boundaries in the Victorian part of the Lachlan (e.g.,
VandenBerg et al., 2000, Spaggiari et al., 2003, 2004). These are similar to those recorded in the
Delamerian Orogen; all have been subdivided into a number of chemical associations, including an
ultramafic and a tholeiitic-boninitic association (e.g., Crawford and Keays, 1987; Crawford et al.,
1984, 2003b; VandenBerg et al., 2000)., and are typically interpreted as having formed in a
suprasubduction zone environment (e.g., Crawford and Keays, 1987; Crawford et al., 1984, 2003b;
Spaggiari et al., 2003, 2004). In the Stawell Zone, the mafic volcanics (Magdala Volcanics;
Figure 2.8, Figure 2.9) have a back-arc signature and are underlain by continental-derived
turbiditic sediments (Crawford et al., 2003b; Squire et al., 2006). These authors interpreted the

Geodynamic Synthesis of the Phanerozoic of eastern Australia 17


succession to represent a distal back-arc environment, related to a west-dipping subduction zone
to the east. This is also consistent with the observation that in the Bendigo and Tabberabbera
zones, the mafic-ultramafic successions appear to form the basement to those zones (e.g.,
Spaggiari et al., 2003, 2004; Korsch et al., 2008; Figure 2.5), and represent oceanic crust as
suggested by Gray, Foster and co-workers (e.g., Gray and Foster, 2004).
• These units are overlain, conformably in places, by Cambrian, deep marine, often pelagic
sedimentation (VandenBerg et al., 2000; Spaggiari et al., 2003). Examples, such as those in the
Bendigo Zone and further east, were not affected by the Delamerian Orogeny (Spaggiari et al.,
2003, 2004). Western examples, including those in the western Stawell Zone, were deformed
during the Delamerian Orogeny (e.g., Miller et al., 2005).
• Interpreted Delamerian-age and older rocks of the Selwyn Block (VandenBerg et al., 2000, Cayley
et al., 2002, Cayley, 2011; Figure 2.8). The Selwyn Block hypothesis suggests the presence of
older continental basement beneath the Melbourne Zone, which links through to Tasmania and the
offshore South Tasman Rise (Cayley et al., 2002; Cayley, 2011). Cayley (2011) coined the term
‘VanDieland’ for this microcontinent. Although some authors have suggested oceanic crust, only,
as basement for the Lachlan Orogen (e.g., Gray, 1997; Gray and Foster, 1997), the presence of
the Selwyn Block appears to be confirmed by the recent central Victorian seismic survey (e.g.,
Korsch et al., 2008; Cayley et al., 2011). The seismic results show the Melbourne Zone to be
underlain by basement distinct from zones to the west.

These associations provide important tectonic constraints. The geological and seismic evidence
suggests that parts of the Lachlan Orogen have either oceanic crust basement e.g., Bendigo Zone, or
older Neoproterozoic–Cambrian continental basement, e.g., Melbourne Zone. This supports the idea
that parts of the now contiguous Lachlan Orogen were separated from one another in the Early
Palaeozoic, as suggested by many authors (e.g., VandenBerg et al., 2000; Cayley et al., 2002; Gray
and Foster, 2004; Cayley, 2011). Crawford et al. (2003a) have suggested that the Selwyn Block
represented a part of Australia that was rifted off at 600 Ma; though see Cayley (2011) for an alternate
viewpoint.

2.1.3 Early to latest Cambrian (ca. 515 Ma to ca. 490 Ma): Delamerian Orogeny

2.1.3.1 Geological and tectonic summary

The break-up of Rodinia and associated extension of southeastern Australia between the late
Neoproterozoic (ca. 600 Ma) and the early part of the Cambrian was halted with the development of
subduction and accompanying contractional orogenesis—the Delamerian Orogeny. In South Australia
and western Victoria, the Delamerian Orogeny commenced at ca. 515 Ma (e.g., Foden et al., 2002a,
2006; Figure 2.7). A similar time is evident for the Delamerian Orogeny (sometimes called Tyennan
Orogeny) in Tasmania (Seymour and Calver, 1995; Figure 2.6). In South Australia, the Delamerian
Orogeny was long-lived, from at least ca. 515 to 490 Ma (e.g., Drexel and Preiss, 1995; Foden et al.,
2006), and perhaps earlier (e.g., Turner et al., 2009; Gibson et al., 2011). Turner et al. (2009) have
suggested that the Delamerian Orogeny may have commenced significantly earlier in southeastern
Australia, based on new ages (ca. 554–525 Ma) for deformation in rocks in the Adelaide Fold Belt. As
noted by those authors these new dates suggest that contraction in southeastern Australia occurred
more or less contemporaneously with that recorded in the Ross Orogen in Antarctica, and not
diachronously as commonly suggested. In the scenario of Turner et al. (2009) Cambrian rocks of the
Kanmantoo and Normanville groups (Figure 2.7) may well have been deposited in a syn-orogenic
setting, analogous to similar-aged sediments in Antarctica.

18 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 2.7 Time-space plots for various regions of the Delamerian Orogen in South Australia, Northern Territory, New South Wales, Victoria and Queensland, and surrounding regions of the Thomson and Lachlan orogens. Modified from Drexel and Preiss (1995),
Jago et al. (2002), Radke (2007), Glen et al. (2013), VandenBerg et al. (2000), Cayley et al. (2011), Greenfield et al. (2010, 2011), Jell (in Withnall and Hutton et al., 2013.)

Geodynamic Synthesis of the Phanerozoic of eastern Australia 19


Regardless, most studies suggest that the orogeny had commenced by 515 Ma, with a number of
deformation events documented. VandenBerg et al. (2000), amongst others, have suggested that
there may have been two deformation stages in western Victoria, at ca. 515 and ca. 490 Ma. They
indicated that the first phase is evident in the Glenelg Zone, the second only in the Grampians-Stavely
Zone (Figure 2.5). Miller et al. (2005) showed that the Delamerian Orogeny also affects the western
part of the Stawell Zone, based on metamorphic ages of ca. 490–500 Ma. Western Tasmania also
records at least two discrete deformation events, separated by a significant extensional event. The
Delamerian (~Tyennan) Orogeny in western Tasmania was initiated, by collision, around 510 Ma
(Berry and Crawford 1988; Crawford and Berry, 1992; Crawford et al., 2003a), with subsequent post-
collisional (back-arc) extension, and renewed contractional deformation around 495 Ma (Crawford and
Berry, 1992; Turner et al., 1998; Figure 2.6).

VandenBerg et al. (2000) showed that, in many respects, western Victoria and western Tasmania
share similar geological histories. The Delamerian Orogeny in western Tasmania, and perhaps also
western Victoria, has been suggested to have been triggered by arc-continent collision, related to
east-dipping subduction east of Tasmania, around 515–510 Ma (Berry and Crawford, 1988;
VandenBerg et al., 2000; Crawford et al., 2003a; Gibson et al., 2011). In both areas collision has been
suggested to have been accompanied by the accretion of Cambrian forearc boninitic crust—the
Tasmanian mafic-ultramafic complex in Tasmania (Crawford and Berry, 1992; Crawford et al., 2003a),
and the Dimboola Igneous Complex in western Victoria (VandenBerg et al., 2000; Crawford et al.,
2003b; Gibson et al., 2011; Figure 2.5, Figure 2.6). Other authors, e.g., Miller et al. (2005) and Cayley
(2011) have questioned the need for an east-dipping subduction zone in western Victoria. They
present a slightly different model, suggesting that emplacement of the Dimboola Igneous Complex
occurred at a later time to that in Tasmania, and was related to backthrusting and west-dipping
subduction. Regardless, nearly all workers suggest west-dipping subduction was operative by the end
of the Delamerian Orogeny (e.g., Gibson et al., 2011). As noted by Squire et al. (2006), this is
consistent with the observed succession (forearc to backarc) evident in many of the mafic-ultramafic
complexes (seafloor remnants) preserved in both the Delamerian and Lachlan orogens (Crawford et
al., 2003b). The presence of these rocks in the Lachlan Orogen strongly indicates an oceanic setting
for much of the Lachlan at this time (e.g., Crawford et al., 1984; Gray, 1997; Fergusson, 2003; Gray
and Foster 2004; Spaggiari et al., 2003).

High temperature, low pressure, metamorphism and metamorphic complexes were developed in both
western Victoria and western Tasmania, with syn-tectonic I- and S-type granites emplaced in the
Glenelg River Metamorphic Complex in Western Victoria (Turner et al., 1993; VandenBerg et al.,
2000; Gray et al., 2002; Crawford et al., 2003b; Gibson et al., 2011). Both regions underwent post-
collisional extension (possibly in a backarc environment), leading to emplacement of calcalkaline
volcanics; the Mount Read Volcanics, correlatives, and intrusives, in Tasmania (Crawford and Berry,
1992; Crawford et al., 2003a), and the Mount Stavely Volcanic Complex in Victoria (Crawford et al.,
1996, 2003b; VandenBerg et al., 2000). The Selwyn Block also contains Cambrian calcalkaline
volcanics—the Jamison-Licola volcanics (e.g., VandenBerg et al., 2000, Spaggiari et al., 2003;
Figure 2.8, Figure 2.9)—which have many similarities to, and have been correlated with, the Mount
Read Volcanics (Crawford et al., 2003a, b). Crawford et al. (2003b) suggested that these volcanics
may represent ‘along strike continuations’ of the Mount Read Volcanics. Cayley (2011), in his model,
also correlated these volcanics with the Mount Stavely Volcanics in western Victoria, suggesting they
all formed along strike of each other. Interpreted environments for these rocks are similar, that is, a
post-collisional, continental-rift setting (Crawford et al., 1996; VandenBerg et al., 2000; Cayley et al.,
2002). Gray, Foster and co-workers (e.g., Spaggiari et al., 2003, 2004; Gray and Foster, 2004) have
invoked an alternate model for the calcalkaline volcanics of the Melbourne Zone. In their ocean floor

20 Geodynamic Synthesis of the Phanerozoic of eastern Australia


basalt model, these rocks are interpreted to be remnants of a Cambrian arc system––the Jamieson
Island Arc––which were subsequently structurally incorporated into the Melbourne Zone.

Calcalkaline volcanism was followed by deformation at ca. 500 Ma to 490 Ma, perhaps related to the
arrival of the VanDieland microcontinent to the Gondwana margin, either further south, e.g., Cayley
(2011), or close to its present position, e.g., Gibson et al. (2011). Subsequent extension resulted in rift
basins, and deposition of molasse-type sediments, e.g., Owen Conglomerate (Seymour and Calver,
1995) in Tasmania. Emplacement of ca. 495–485 Ma post-tectonic granites in western Victoria and
South Australia (e.g., VandenBerg et al., 2000; Foden et al., 2002a, 2006; Figure 2.7) constrain the
end of the Delamerian cycle. At this time, subduction is best recorded well to the east, in the
Ordovician Macquarie Arc (Crawford et al., 2007a). Cayley (2011) has suggested this jump in
subduction to the east may also relate to the arrival of the VanDieland microcontinent.

Figure 2.8 Distribution of various lithological and chemical associations of the Neoproterozoic to Cambrian
volcanic rocks of Victoria. Figure modified from VandenBerg et al. (2000). Rocks at Ceres, Phillip Island, Glen
Creek and Jamieson-Licola are interpreted to form part of the Selwyn Block (VandenBerg et al., 2000; Cayley et
al., 2002). Others (e.g., Gray and Foster, 1997, 2004; Spaggiari et al., 2004) interpret the Jamieson-Licola rocks
as remnants of an island arc.

2.1.3.2 Geological history and Time-Space plot explanation (Figure 2.4 to Figure 2.9)

• Stage 1 Delamerian deformation in Western Victoria (Grampians Zone) and Tyennan deformation
in Western Tasmania, in the Early to Middle Cambrian (from ca. 520–515 Ma), e.g., Berry and
Crawford (1988), Crawford and Berry (1992), Turner et al. (1998), VandenBerg et al. (2000),
Foden et al. (2002a, 2006), Miller et al. (2005), Cayley (2011).
• This deformation was accompanied by the emplacement of numerous allochthonous blocks,
including mafic (tholeiites, boninites, eclogites), ultramafic and sedimentary rocks (Berry and
Crawford, 1988; Crawford and Berry, 1992; Turner et al., 1998; VandenBerg et al., 2000). Dating

Geodynamic Synthesis of the Phanerozoic of eastern Australia 21


of these rocks indicate Middle Cambrian ages in Tasmania (515–510 Ma; Turner et al., 1998). This
is consistent with ages for the obduction event there, which is bracketed between ca. 515 and ca.
500 Ma (from a tonalite in the ophiolite; Black et al., 1997; and the overlying Mount Read
Volcanics; Berry and Crawford, 1988; Crawford et al., 2003a). Similar ages are recorded in
western Victoria (see VandenBerg et al., 2000; Crawford et al., 2003b), though Cayley (2011),
suggests this event may have been slightly younger there.
• Significant deformation and metamorphism is recorded in Tasmania, e.g., Ulverstone
Metamorphics, Forth Metamorphic Complex, Glenelg River Metamorphic Complex (VandenBerg et
al., 2000, Crawford et al., 2003b; Gray et al., 2002, 2003; Berry et al., 2007). Berry et al. (2007)
recorded ages of ca. 510 Ma for metamorphism. In Victoria syntectonic magmatism is recorded in
the Glenelg Zone of Victoria, representing the eastern extent of this magmatism which is best
developed in South Australia (e.g., Turner et al., 1993; Drexel and Preiss, 1995; Foden et al.,
2002a, 2006). Maher et al. (1997) report an age of ca 504 Ma for a tonalite from drill core and
suggested deformation was probably of the same age.
• Post-collision Middle Cambrian east-west extension (e.g., Crawford and Berry, 1992) led to
formation of the Dundas Trough and the mixed-derivation sedimentary succession of the Lower
Dundas Group and correlatives in Tasmania (Seymour and Calver, 1995). Marine sedimentation of
the Nargoon Group in the Grampians-Stavely Zone of Victoria (VandenBerg et al., 2000; Crawford
et al., 2003b), and similar aged deep-marine sedimentation occurred within the Stawell Zone at
this time (St Arnaud Group; e.g., VandenBerg et al., 2000).
• The mainly felsic, calcalkaline Mount Read Volcanics, and associated volcano-sedimentary
successions (Crawford and Berry, 1992; Munker and Crawford, 2000) formed around this time, as
did similar rocks (e.g., Mount Stavely Volcanic Complex) in Victoria (VandenBerg et al., 2000;
Crawford et al., 1996; 2003b; Squire et al., 2006). In Tasmania, the volcanics were intruded by late
dacites and granites. Similar-aged calcalkaline volcanic rocks (e.g., Licola and Jamieson
Volcanics) occur as windows in the Melbourne Zone (VandenBerg et al., 2000; Spaggiari et al.,
2003). These have compositions similar to the Mount Read Volcanics of Tasmania (Crawford et
al., 1996, 2003a, b).
• A second phase of deformation occurred between 505 Ma and 495 Ma in western Victoria,
Tasmania and the Selwyn Block (VandenBerg et al., 2000; Seymour and Calver, 1995). In
Tasmania, this deformation has been subdivided into an early phase of north-south contraction
and later east-west contraction (e.g., Seymour and Calver, 1995). Berry et al. (2007) recorded no
metamorphism of this age in western Tasmania, although they did in offshore samples which they
related to the Ross Orogeny in Antarctica.
• Although now reported from the western part of the Stawell Zone (Miller et al., 2005), the Delamerian
Orogeny did not affect rocks in the Lachlan Orogen. The deformation is, however, recorded in the
Selwyn Block underlying the Melbourne Zone (VandenBerg et al., 2000; Cayley et al., 2002; although
see Spaggiari et al., 2003). The presence of deformed calcalkaline volcanic rocks which have been
equated with the Mount Read Volcanics in Tasmania (Crawford et al., 2003a), strongly suggests that
deformation was synchronous with the second phase of the deformation as recorded elsewhere. On
the basis of the correlation between the Selwyn Block and Tasmania, this deformation has been
referred to as the Tyennan Orogeny (VandenBerg et al., 2000; Cayley et al., 2002).
• Further extension, possibly commencing between the two Middle-Late Cambrian deformations,
resulted in deposition of coarse sediment, e.g., Owen Conglomerate and correlatives, and upper
Dundas Group, in Tasmania (Munker and Crawford, 2000; Crawford et al., 2003a; Seymour and
Calver, 1995) and post-tectonic magmatism, including A-types, in western Victoria and South
Australia, ca. 495–485 Ma (Foden et al., 2002a, 2006).

22 Geodynamic Synthesis of the Phanerozoic of eastern Australia


• Continuing marine, mostly deep-water turbiditic sedimentation in the Stawell Zone (e.g., St Arnaud
Group), and largely pelagic sedimentation in the Bendigo and Tabberabbera zones (e.g., Goldie
Chert) of Victoria (VandenBerg et al., 2000; Crawford et al., 2003b; Gray et al., 2003). The
Bendigo and Tabberabbera zones contain evidence of deposition in an oceanic environment
distant from continental Australia. As summarised by VandenBerg et al. (2000), these include the
pelagic sedimentation, the apparently conformable relationship with ocean floor rocks, and the lack
of evidence for the Delamerian Orogeny in these zones.

2.1.4 Latest Cambrian to early Silurian (ca. 490 Ma to ca. 430 Ma)

2.1.4.1 Geological and tectonic summary

The Lachlan Orogen is dominated by two contrasting rock packages through this time interval:

• Voluminous deep water quartz-rich turbidites of cratonic provenance and pelagic sediments. These
occur throughout the Lachlan Orogen, with little evidence for significant volcanic input. They are
conformable on oceanic crust in a number of regions.
• Calcalkaline magmatism and volcaniclastics, and marine sediments with common carbonates—
largely related to the Macquarie Arc (e.g., Glen, 2004; Crawford et al., 2007a, Glen et al., 2011);
called the Macquarie Volcanic Belt by Quinn et al. (2014).

These rock packages are thought to have been juxtaposed as part of the Benambran Orogeny (e.g.,
VandenBerg et al., 2000; Glen et al., 2007b; Moresi et al., 2014; although cf. Quinn et al. 2014), which
appears to have occurred in two main pulses, ca. 440 and 430 Ma (e.g., Glen et al., 2007b). Syn-
tectonic, largely S-type magmatism accompanied this deformation.

During the late Cambrian to the end of the Ordovician and early part of Silurian, most of the Lachlan
Orogen, in NSW, Victoria and Tasmania, was the site of deep marine sedimentation (Figure 2.4,
Figure 2.5, Figure 2.6, Figure 2.9; e.g., Vandenberg et al., 2000; Percival et al., 2011). This comprises
quartz-rich turbiditic successions (Western, Central and Eastern Lachlan) and late Middle to Late
Ordovician black shale-dominated sediments (Central and Eastern Lachlan). In a number of regions,
e.g., Bendigo and Tabberabbera zones (e.g., VandenBerg et al., 2000; Fergusson and VandenBerg,
2003; Spaggiari et al., 2003), this sedimentation appears to be conformable upon mafic and ultramafic
rocks interpreted as oceanic crust (see previous section). In most areas of the Lachlan Orogen, this
deep water sedimentation ended with the Benambran Orogeny. Sedimentation, however, continued in
both the Melbourne Zone (Victoria) and northeastern Tasmania, and both regions apparently show no
evidence for the Benambran Orogeny (e.g., Fergusson and VandenBerg, 2003; Seymour and Calver,
1995; Figure 2.5, Figure 2.6, Figure 2.9).

Most workers imply the presence of one or more submarine fans for the Ordovician deep marine
sedimentation in the Lachlan Orogen (e.g., Fergusson and VandenBerg, 2003; Gray and Foster, 2004;
Glen, 2005; Glen et al., 2007b). This may have reflected uplift of the Delamerian Orogen in the early
Ordovician (e.g., VandenBerg et al., 2000; Fergusson and VandenBerg, 2003), although there is
uncertainty regarding the relative positions of the respective parts of the Lachlan Orogen at this time.
The major change in sedimentation recorded by the switch to black shale-dominated pelagic
sediments during the late Middle Ordovician and their localisation largely to the Central and Eastern
Lachlan is, in part, contemporaneous with an early phase of the Benambran deformation (ca. 455 Ma)
recorded in the Western Lachlan (VandenBerg et al., 2000; Gray et al., 2003; Miller et al., 2005).

Various workers have attempted to subdivide the Ordovician sediments into geological domains and
terranes; most recently by Glen et al. (2009) who recognised 4 tectonostratigraphic terranes on the

Geodynamic Synthesis of the Phanerozoic of eastern Australia 23


basis of stratigraphic and biostratigraphic differences. Most controversy surrounds the geographic
location and relative position of the Ordovician sediments with respect to the contemporaneous Early
Ordovician to earliest Silurian Macquarie Arc. The Ordovician sediments contain very little evidence of
volcanic detritus, and it would appear, as suggested by numerous workers (e.g., Gray and Foster,
2004; Meffre et al., 2007), that the contemporaneous Macquarie Arc was disconnected from the deep
marine sedimentation, perhaps by hundreds of kilometres or more (Meffre et al., 2007). This is
consistent with the interpreted oceanic environment for the Macquarie Arc (e.g., Percival and Glen,
2007). More recent work, however, suggests that the Macquarie Arc and Ordovician sediments may
be more closely associated, i.e., it is likely that parts of the arc were in closer proximity with the deep
marine sedimentation. Evidence for this comes from a number of sources. Firstly, as pointed out by
Glen et al. (2011), ages and Hf isotopic signatures of detrital zircons in sediments associated with
phase 1 of the arc (ca. 480 Ma) suggest a provenance similar to that of the Ordovician sediments.
Secondly, there is evidence for some Early Ordovician detrital zircons (ca. 476 Ma) within the
Ordovician sediments (Gilmore et al., 2012). Importantly, these sediments also contain VHMS
mineralisation which has Pb isotopic model ages of ca. 490–470 Ma consistent with a back-arc
environment (Huston et al., 2016), i.e., not too distant from an arc. Thirdly, recent detailed mapping
accompanied by detailed fossil dating by Quinn et al. (2014) has shown that not only do the two
sequences interfinger in the Kiandra area, but that episodes of volcanism in the arc can be matched to
switches in sedimentation in the turbiditic sequences on both sides of the arc, from turbiditic
sandstones to cherts and black mudstones. The converse also appears to be the case, i.e., hiatuses
in volcanism are matched by a return to turbiditic sandstones in the surrounding sediments.

The Macquarie Arc consists of calcalkaline and shoshonitic volcanics, intrusions and volcaniclastic
and carbonate-rich successions, typically suggested to have formed in an intra-oceanic arc setting
(e.g., Crawford et al., 2007a), although Wyborn (1992), and more recently Quinn et al. (2014), have
suggested that the largely shoshonitic magmatism reflected reworking of older (Cambrian?)
subduction-modified lithosphere. The arc is preserved as elongate remnants, mostly in New South
Wales (Figure 2.9), but also includes the Kiandra Group in northern Victoria (Fergusson and
VandenBerg, 2003). Recent work (e.g., Percival et al., 2011; Quinn et al. 2014) has suggested that the
easternmost belt––the Rockley Volcanic Belt––is possibly younger and therefore not part of the
Macquarie Arc. Crawford et al. (2007a, b and companion papers) suggests that the Macquarie Arc
was built up in four successive phases of growth, with apparently two distinct (east and west)
provinces whose relative positions are uncertain (Percival and Glen, 2007). The arc is thought to have
accreted to eastern Australia in the Early Silurian as part of the Benambran Orogeny (Glen et al.,
2007b; Meffre et al., 2007).

Subduction zone polarity for the Macquarie Arc is uncertain and may have been west-dipping, east-
dipping or a mixture of east- and west-dipping subduction (e.g., see models in Meffre et al., 2007). The
original orientation of the arc is also uncertain–more recent models have suggested a more east-west
strike to the arc, e.g., Fergusson et al. (2013), Cayley (2011), and Gibson et al. (2011). Many models
also include significant strike-slip motion south of the arc.

Finally, it should be noted that although general consensus favours an arc interpretation for these
rocks, other interpretations have been invoked. Wyborn (1992), for example, suggested the arc-like
geochemical signature reflected a previous (Cambrian), not current, subduction environment. Wyborn
(1992) speculated that a plume-type environment could explain the Ordovician shoshonitic, and
subsequent Silurian-Devonian felsic, magmatism. These models are consistent with the new
stratigraphic mapping of Quinn et al. (2014) who suggested a rift environment, located behind an arc
further to the east.

24 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Late Cambrian to Ordovician deep marine sediments, with minor Middle to Late Cambrian limestone
and associated clasts, breccia units and small outcrops of mafic volcanics, comprise the Wagonga
Group. These were deposited in the Batemans Bay and Narooma region (Lewis et al., 1994; Glen et
al., 2004; Prendergast and Offler, 2012; Stokes et al., 2015), in what has also been termed the
Narooma Terrane (Glen et al., 2004; Figure 2.1). The upper part of this succession has an increasing
continental component, interpreted to represent the increasing influence of mainland Australia (Glen et
al., 2004; Glen, 2005).

Glen et al. (2004) interpreted the Narooma Terrane as being oceanic in origin; Miller and Gray (1997), in
contrast, presented an accretionary wedge model to explain this terrane. The latter viewpoint is supported
by the more recent work by Prendergast et al. (2012), Prendergast and Offler (2012) and Stokes et al.
(2015) who document alkaline and subalkaline mafic compositions which they attributed to backarc and
ocean island seamount settings, respectively. Stokes et al. (2015) have also suggested that the Wagonga
Group is interdigitated with the Ordovician Adaminaby Group turbidites, and Prendergast et al. (2012)
showed both units have similar metamorphic histories, supporting an accretionary complex origin.

Ordovician sedimentation is not restricted to the Lachlan Orogen. Ordovician shallow marine and
terrestrial sedimentation was deposited on top of rocks of the Delamerian Orogen, e.g., in western
Tasmania (Seymour and Calver, 1995; Figure 2.6) and in the Koonenberry region (Gilmore et al.,
2007). Post-tectonic Ordovician A- and I-type magmatism occurred in the Delamerian Orogen, e.g., in
the Glenelg Zone, Victoria, and South Australia (Turner et al., 1993; Foden et al., 2002a, 2006; Gray
et al., 2002). No sedimentation is known from western Victoria at this time.

Ordovician mafic volcanics, chert and serpentinite and other (Cambrian–Ordovician) ultramafic rocks,
interpreted as oceanic crust, occur within the Eastern Lachlan in New South Wales, e.g., Jindalee
Group (Warren et al., 1995; Lyons et al., 2000; Figure 2.4). They also occur between the remnants of
the Macquarie Arc (e.g., Glen et al., 2011). As indicated by Glen (2005) and Meffre et al. (2007), these
probably represent remnants of the basement to the Ordovician sediments, brought up by subsequent
deformation, analogous to the older examples in Victoria. Possible Late Ordovician ultramafic rocks
also occur in the Western and Central Lachlan, interpreted as intrusives into the Ordovician sediments
(e.g., Lyons et al., 2000; Fraser et al., 2014). Burton et al. (2012, 2013) report basaltic rocks of MORB
and OIB composition of probable Early Ordovician age within the Girilambone Group in the Cobar
region, e.g., Narrama Formation.

Deformation associated with the Benambran Orogeny commenced in the Western Lachlan, e.g., Stawell
and Bendigo zones, at ca. 455–440 Ma (VandenBerg et al., 2000; Gray et al., 2003; Gray and Foster,
2004) and ca. 440 Ma (and slightly older) in other parts of the Lachlan (e.g., Collins and Hobbs, 2001; Gray
et al., 2003; Glen et al., 2007b). The resulting Benambran Orogeny affected most of the Lachlan Orogen.
Exceptions to this are the Melbourne Zone and eastern and western Tasmania, where deep-marine
sedimentation continued throughout the Benambran Orogeny (VandenBerg et al., 2000; Seymour and
Calver, 1995, 1998). Late Ordovician to early Silurian turbiditic sedimentation is also recorded in the
Central and Eastern Lachlan in Victoria, e.g., the syn-tectonic Yalmy Group (VandenBerg et al., 2000;
Willman et al., 2002; VandenBerg, 2003; Glen et al., 2007b; Figure 2.5). These rocks have significantly
more quartz than underlying rocks and reflect a change in provenance, which VandenBerg et al. (2000)
suggested was possibly related to uplift associated with the first phase of the Benambran Orogeny.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 25


26 Geodynamic Synthesis of the Phanerozoic of eastern Australia
Figure 2.9 Generalised distribution of rocks in the Lachlan Orogen, by time period. A: late Neoproterozoic to early
Cambrian (ca. 600 Ma to ca. 515 Ma), B: early to latest Cambrian (ca. 515 Ma to ca. 490 Ma), C: latest Cambrian
to early Silurian (ca. 490 to ca. 430 Ma), D: Middle Silurian to Late Devonian (ca. 430 Ma to ca. 380 Ma), E: Late
Devonian to early Carboniferous (ca. 380 Ma to ca. 350 Ma), F: Middle Carboniferous to earliest Triassic (ca. 350
Ma to ca. 250 Ma).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 27


Benambran deformation was accompanied by crustal thickening and uplift, regional, locally significant,
metamorphism (in rocks of the Adaminaby Superterrane, such as in the Wagga-Omeo Metamorphic
Belt) and additionally also marked the end of recorded arc volcanism in the Macquarie Arc (Crawford et
al., 2007a). Significant higher grade metamorphism (e.g., Gray, 1997; Gray et al., 2003), as well as syn-
tectonic S-type magmatism (e.g., Collins and Hobbs, 2001), accompanied the orogeny. Both are largely
concentrated in two, non-parallel, belts, in the Central Lachlan, and the Eastern Lachlan, respectively
(Figure 2.9). Collins and Hobbs (2001) suggested that at least some of the thermal input was provided
by mantle-derived magmatism. The orogeny probably involved at least two discrete late deformation
events at ca. 440 Ma and 430–425 Ma (Glen et al., 2007b) and resulted in a complex arrangement of
terranes, particularly in eastern NSW. Phase 2 of the Benambran Orogeny (ca. 430–425 Ma), best
developed in the Eastern Lachlan, is thought to be related to renewed east-west to more oblique
contraction to strike-slip deformation (with a north-south component) (e.g., VandenBerg et al., 2000;
Collins and Vernon, 1994; VandenBerg, 2003; Pogson and Glen, 2006; Glen et al., 2007b).

The end result of the Benambran Orogeny was effectively cratonisation of much of the Western
Lachlan (VandenBerg, 2003), and complex accretion of a number of terranes in the central and
eastern part of the orogen (VandenBerg et al., 2000; Willman et al., 2002; Glen, 2005; Glen et al.,
2007b, 2009; Meffre et al., 2007). These include accretion of the Macquarie Arc and what Glen et al.
(2009) call the Albury-Bega Terrane (e.g., Glen, 2005; Glen et al., 2007b, 2009) and perhaps the
Narooma Terrane (Glen, 2005; although see Prendergast et al., 2012 for an alternative view). Willman
et al. (2002) suggest that this event was responsible for amalgamation of parts of their Benambra
Terrane (Central and Eastern Lachlan), although VandenBerg et al. (2000) and Moresi et al. (2014),
for example, indicate younger strike-slip movement (Bindian Orogeny) and other tectonic adjustments
may be responsible for bringing the central and eastern zones into their final positions.

Overall, the Lachlan Orogen appears to record a relatively simple ‘passive-margin’ to deep marine
environment and an interpreted oceanic arc environment. These are often depicted as an oceanic
backarc basin (marginal sea) behind an oceanic arc and west-dipping slab (e.g., Coney, 1992; Glen et
al., 1998; Cayley et al., 2002; Fergusson, 2003; Gray et al., 2003; Gray and Foster, 2004; Glen, 2005).
In detail, however, the situation is more complex and controversy exists with respect to the position of
terranes, the number and location of subduction zones and mechanisms of terrane accretion (e.g.,
Coney, 1992; Gray, 1997; Soesoo et al., 1997 and discussion papers; VandenBerg et al., 2000;
Collins and Hobbs, 2001; Cayley et al., 2002; Willman et al., 2002; Cas et al., 2003; Fergusson, 2003;
Gray et al., 2003; Spaggiari et al., 2003, 2004; Glen, 2005; Glen et al., 2007b, 2009; Cayley, 2011;
Moresi et al., 2014), and even whether subduction was present at all, e.g., Wyborn (1992), Quinn et al
(2014). Many models invoke a number of subduction zones (e.g., Gray, 1997; Soesoo et al., 1997,
Collins and Hobbs, 2001; Fergusson, 2003; Spaggiari et al., 2003, 2004) to explain the across-orogen
variations in magmatism, metamorphism, deformation, especially vergence changes, and the
presence of blueschists. The actual mechanisms and detail of these reconstructions have important
implications for mineralisation, given that the Benambran Orogeny coincides with significant lode Au
(e.g., Bendigo) and arc-related Cu-Au (e.g., Cadia) mineralisation (e.g., VandenBerg et al., 2000;
Crawford et al., 2007a).

Neither phase of the Benambran deformation appear to have significantly affected the Melbourne
Zone in Victoria (VandenBerg et al., 2000), the Mathinna Supergroup in Tasmania (Seymour and
Calver, 1995), or Western Tasmania, as marine sedimentation continued in these areas throughout
this period. Sedimentation in Western Tasmania commenced with post-Delamerian extension and
formation of rift or sag basins, with deep water sediments ca. 490 Ma to 470 Ma, followed by extensive
Ordovician carbonate-dominated sedimentation and deeper water clastic-dominated sedimentation
around the end of the Ordovician (Seymour and Calver, 1995, 1998; Figure 2.6, Figure 2.9).
Deposition in this region continued intermittently until the Middle Devonian, and the Benambran and

28 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Bindian orogenies appear to be absent. Hiatuses recorded in several elements (e.g., Sheffield
Element), however, may correspond to the Benambran Orogeny (Seymour and Calver, 1995, 1998),
as may the switch from carbonate- to clastic-dominated sedimentation. It is possible that the apparent
absence of these orogenies may simply reflect the presence of the Selwyn Block, such that
deformation was largely partitioned around it (e.g., VandenBerg et al., 2000). Reed (2001) has inferred
an earlier Benambran deformation in the older western part of the Mathinna Supergroup. Detrital
zircon data has been interpreted to suggest western and eastern Tasmania were separate during this
period (e.g., Black et al., 2004). In this regard, Cayley (2011) has invoked a model where by the
Selwyn block––his VanDieland microcontinent––moved northward over 300 km, by dextral strike-slip,
into close to its present position during the Benambran cycle. In this model the northward movement
of the Selwyn Block plays an important role in the deformation of the Stawell and Bendigo zones
(floored by ocean crust), caught between the Selwyn Block and mainland Australia.

2.1.4.2 Geological history and Time-Space plot explanation (Figure 2.4 to Figure 2.9)

Ordovician deep marine turbiditic sediments occur in the Western, Central and Eastern Lachlan: in
Victoria and NSW, e.g., St Arnaud?, Castlemaine, Adaminaby, Sunbury, Wagga and Girilambone
groups (e.g., VandenBerg et al., 2000; Crawford et al., 2003b; Glen, 2005; Watkins and Meakin, 1996;
Lyons et al., 2000; Colquhoun et al., 2005; Percival et al., 2011)––called the Adaminaby Superterrane
by Glen et al. (2007b); and in Tasmania—the Mathinna Supergroup (Seymour and Calver, 1995).
These are well developed sequences, the major exception being the Melbourne Zone where rocks of
this age are poorly developed (Fergusson and VandenBerg, 2003). Burton et al. (2012, 2013) report
basaltic rocks of MORB and OIB composition of probable Ordovician age within the Girilambone
Group in the Cobar region, e.g., in the Kaiwilta and Mount Dijou Volcanic members of the Early
Ordovician Narrama Formation. Percival et al. (2011) have shown this volcanism (their Figure 5) as
being contemporaneous with chert deposition in the Adaminaby and Wagga groups.

Post-tectonic A- and I-type magmatism, ca. 495–485 Ma, with no accompanying sedimentation, occur
within the mainland Delamerian Orogen at this time e.g., in the Glenelg Zone, Victoria, and South
Australia (Foden et al., 2006). Western Tasmania records deep water sedimentation followed by
extensive Ordovician carbonate-dominated sedimentation (e.g., Gordon Group) (see below; Seymour
and Calver, 1995).

Late Ordovician, mainly pelagic, black-shale dominated rocks, e.g., Bendoc Group (Lewis et al., 1994;
VandenBerg et al., 2000; Seymour and Calver, 1995, 1998; Colquhoun et al., 2005; Percival et al., 2011),
appear to be confined to the Central and Eastern Lachlan in NSW and Victoria, and eastern Tasmania.
They are largely absent from most of the Melbourne Zone, only apparently occurring around the eastern
margins (Mount Easton Shale; VandenBerg et al., 2000; Fergusson and VandenBerg, 2003). Deep water
clastic sedimentation appears at the top of these successions, at least locally (e.g., VandenBerg et al.,
2000; Colquhoun et al., 2005). Turbiditic sandstones, shale and mudstones of the Sunbury group, of similar
age to the Bendoc group and Mount Easton Shale, occur in the margin of the Bendigo and Melbourne
zones (VandenBerg et al., 2000). The relationship with the Bendigo Zone is uncertain.

Late Cambrian to Ordovician deep marine sediments, which overlie Cambrian mafic volcanics with
OIB and MORB affinities (Lewis et al., 1994; Glen et al., 2004; Prendergast and Offler, 2012; Stokes
et al., 2015), of the Wagonga Group, were also deposited during this time in southern New South
Wales (the Narooma Terrane of Glen et al., 2004). The upper part of this succession has an
increasing continental component (Glen et al., 2004).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 29


Glen et al. (2009) split the turbidite and black shale packages into 4 terranes: the Bendigo Terrane
(Castlemaine Group), the Hermidale Terrane (Girilambone Group), the Melbourne Terrane (Sunbury
Group) and the Albury-Bega terranes (Wagga and Bendoc Groups).

Ordovician mafic volcanics, chert and serpentinite and other (Cambrian–Ordovician) ultramafic rocks
occur within the Central and Eastern Lachlan in New South Wales, e.g., Jindalee Group (Warren et
al., 1995; Lyons et al., 2000). As outlined by Glen et al. (2009) the mafic rocks often lie along or close
to major faults. They have MORB-like chemistry leading Glen et al. (2009) to suggest they represent
oceanic crust. As discussed by Percival et al. (2011) many of these ultramafic (± basalt, gabbro, chert
and jasper) units, interpreted as seafloor and mantle, occur as blocks within younger units. Intrusive
ultramafic units are also recorded. These include the Alaskan-type ultramafic complexes (ultramafic
and feldspathic rocks) of the Fifield suite (e.g., Barron et al., 2004) which mostly occur in a >300km
belt north and south of Nyngan, within the Girilambone Group. Fraser et al. (2014) dated one of these
units–the Honeybugle Diorite–at 442± 2.1 Ma, an age similar to the last phase of (shoshonitic)
magmatism in the adjacent Macquarie Arc (e.g., Crawford et al., 2007b). Notably, Barron et al. (2004)
indicated a shoshonitic affinity for the feldspathic components of the Fifield Complexes.

Shoshonitic and less potassic calcalkaline volcanics, intrusions and volcaniclastic and carbonate-rich
successions were formed in the Ordovician to earliest Silurian. These rocks are preserved in four
zones: the Junee-Narromine Volcanic Belt, the Kiandra Volcanic Belt, the Molong Volcanic Belt, and
the Rockley-Gulgong Volcanic Belt (Figure 2.1). It should be noted however, that recent work (Percival
et al., 2011; Quinn et al., 2014) suggests the Rockley-Gulgong Volcanic Belt may, in part, be younger.
The volcanic belts have collectively been called the Macquarie Arc (e.g., Glen, 2004, 2005; Crawford
et al., 2007a), and more recently, the Macquarie Volcanic Belt, by Quinn et al. (2014), who questioned
the suggested arc origin. Detailed work was undertaken on the Macquarie Arc in a joint project by the
Geological Survey of New South Wales and University of Tasmania in the late 1990’s and early
2000’s. As outlined in a series of papers from that work (e.g., Crawford et al, 2007a,b; Percival and
Glen, 2007; Glen et al., 2007a), the Macquarie Arc was suggested to have been built up in four
successive phases from the Early Ordovician to the Late Ordovician to earliest Silurian. Percival and
Glen (2007) documented the presence of two distinct (east and west) provinces within the Macquarie
Arc which may not have been together until accretion.

Early Benambran Orogeny east-west shortening deformation and east vergence is recorded in the
Western Lachlan, e.g., Stawell and Bendigo zones, ca. 455–440 Ma (VandenBerg et al., 2000; Gray et
al., 2003; Gray and Foster, 2004). This early deformation may be diachronous, commencing in the
west and younging to the east (e.g., Gray and Foster, 1997). The Benambran Orogeny is also
recorded in other parts of the Lachlan Orogen (e.g., Collins and Hobbs, 2001; Gray et al., 2003; Glen
et al., 2007b). Deformation was accompanied by crustal thickening, uplift, regional, locally significant,
metamorphism (in rocks of the Adaminaby Superterrane, such as in the Wagga-Omeo metamorphic
belt; Gray, 1997; VandenBerg et al., 2000; Willman et al., 2002; Gray et al., 2003; Gray and Foster,
1997, 2004; Glen et al., 2007b) and also marked the end of recorded arc volcanism in the Macquarie
Arc (Crawford et al., 2007a).

Late Ordovician to early Silurian turbiditic sedimentation is recorded in the Central and Eastern
Lachlan in Victoria, e.g., the syn-tectonic Yalmy Group (VandenBerg et al., 2000; Willman et al., 2002;
VandenBerg, 2003; Glen et al., 2007b). This sedimentation was contemporaneous with deep-marine
sedimentation in the Melbourne Zone in Victoria, and western and eastern Tasmania (Seymour and
Calver, 1995). Possible correlatives of the Yalmy Group may also be present in New South Wales
(Percival et al., 2011).

30 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Phase 2 of the Benambran Orogeny at ca. 430–425 Ma, is best developed in the Eastern Lachlan as
east-west to more oblique contraction to strike-slip deformation (with a north-south component) (e.g.,
VandenBerg et al., 2000; Collins and Hobbs, 2001; VandenBerg, 2003; Pogson and Glen, 2006; Glen
et al., 2007b). This deformation is not apparent everywhere, e.g., Colquhoun et al. (2005) indicate
deformation in the Cargelligo region of New South Wales had ended prior to ca. 432 Ma intrusion of
post-tectonic S-type granites.

Continuing regional metamorphism, such as in the Wagga-Omeo Metamorphic Belt and Cooma
region, accompanied the second phase of the Benambran deformation (Foster and Gray, 1997; Gray
et al., 2003), and was associated with extensive S-type syn- to post-tectonic magmatism (Collins and
Hobbs, 2001; Ickert and Williams, 2011). The S-type magmatism is concentrated within two large non-
parallel belts (e.g., Chappell et al., 1991; Collins and Hobbs, 2001), one in the Central Lachlan and the
other in the Eastern Lachlan. Available recent geochronology questions this proposed zonation and
rather indicates that ca. 435–425 Ma magmatism is widespread across the Central and Eastern
Lachlan Orogen (Ickert and Williams, 2011; Bodorkos et al., 2013, 2015; Fraser et al., 2014; Chisholm
et al., 2014b,c)

No phase of Benambran deformation appears to have significantly affected the Melbourne Zone in
Victoria (VandenBerg et al., 2000), or the Mathinna Supergroup in Tasmania (Seymour and Calver,
1995), as these regions were the sites of continued marine sedimentation throughout this period.
Similarly, sedimentation in Western Tasmania continued throughout this time. Sedimentation there
followed post-Delamerian extension which resulted in the formation of rift or sag basins, that
subsequently filled with deep water sediments including conglomerates, such as the ca. 490 to 470
Ma Owen Conglomerate (e.g., Seymour and Calver, 1995). This was followed by extensive Ordovician
carbonate-dominated sedimentation (e.g., Gordon Group) and a subsequent switch to deeper water
clastic-dominated sedimentation (e.g., Eldon Group) around the end of the Ordovician (Seymour and
Calver, 1995, 1998). Deposition continued intermittently until the Middle Devonian, and although
hiatuses are recorded in several elements (e.g., Sheffield element), evidence for the Benambran and
Bindian orogenies appear to be absent.

2.1.5 Middle Silurian to Late Devonian (ca. 430 Ma to ca. 380 Ma)

2.1.5.1 Geological and tectonic summary

This time period in the Lachlan Orogen (Figure 2.1 to Figure 2.9) is marked by widespread extensional
episodes with accompanying basin formation and ubiquitous extrusive and intrusive magmatism.
Differing geological histories are observed within the Orogen, in particular between the Western
Lachlan (Whitelaw Terrane) and Central and Eastern Lachlan (Benambra Terrane), suggesting these
regions may have been separate for most of this cycle (e.g., Gray and Foster, 1997, 2004;
VandenBerg et al., 2000; Willman et al., 2002; Spaggiari et al., 2004). In addition, the regions
associated with the Selwyn Block (Cayley et al., 2002), also record a unique event history.

Within the Central and Eastern Lachlan, extension developed within and across the juxtaposed
Ordovician turbidite successions and the Macquarie Arc remnants in New South Wales (Meakin and
Morgan, 1999; Lyons et al., 2000; Glen et al., 2007b; Figure 2.4, Figure 2.9), and within the Ordovician
turbidite successions in Victoria (e.g., VandenBerg et al., 2000; Figure 2.5). This resulted in
widespread largely deep to shallow marine sedimentation (including carbonates) in troughs,
platforms/shelves and other rift zones (Pogson and Watkins, 1998; Meakin and Morgan, 1999; Lyons
et al., 2000; VandenBerg et al., 2000; VandenBerg, 2003; Colquhoun et al., 2005; Glen, 2005). This is
interpreted to reflect post-Benambran east-west extension or transtension, and the formation of rift

Geodynamic Synthesis of the Phanerozoic of eastern Australia 31


basins (e.g., VandenBerg et al., 2000; Willman et al., 2002; Colquhoun et al., 2005; Glen, 2005;
Cooney and Mantaring, 2007), related to an easterly roll-back migration of the arc following the
Benambran Orogeny (e.g., Gray and Foster, 2004; Glen et al., 2004; Spaggiari et al., 2004).
Accompanying magmatism includes S- and I-type granites that are often bimodal, with intermediate
magmatism uncommon (Chappell and White; 1992; Lyons et al., 2000; Collins and Hobbs, 2001; Gray
et al., 2003; Rossiter, 2003; Black et al., 2005). Glen (2005) suggested that volcanism is absent from
basins north of the Lachlan Transverse Zone, which he suggested reflected poorly understood
differences in the thermal structure of the region.

These basins were inverted during the poorly-defined, latest Silurian (–Earliest Devonian) Bindian
Orogeny (ca. 420–410 Ma; Figure 2.4, Figure 2.5). This orogeny, most evident in Victoria and New
South Wales, has been suggested to be transpressive and to have resulted in significant strike-slip
movement between the western and central Lachlan (VandenBerg et al., 2000; Willman et al., 2002;
Glen, 2005). Willman et al. (2002), for example, suggested up to 600 km of dextral movement
between the Whitelaw and Benambra terranes, largely accommodated along what they called the
Baragwanath Transform. Most of this strike-slip movement is thought to have occurred during the
Bindian deformation (Willman et al., 2002). This movement may also relate to the effects of rotation
and orocline formation, for example, as suggested by Cayley (2011, and co-workers, e.g., Moresi et
al., 2014). Other workers (e.g., Spaggiari et al., 2003), find no evidence for significant strike-slip
movement, at least along the Whitelaw-Benambra terrane boundary. It is possible that deformation of
this age may relate more to continuing subduction-accretion effects if the multiple subduction zone
models of Gray (1997), Gray and Foster (1997), Soesoo et al. (1997), and Spaggiari et al. (2003,
2004), are correct. This, however, does not necessarily negate some strike-slip effects in the Lachlan
Orogen (e.g., Glen et al., 2009). In contrast, Bindian deformation appears to relate more to east-west
contraction in far eastern Victoria (e.g., Willman et al., 2002).

Renewed extension and development of rift basins continued in the Early Devonian in the Central and
Eastern Lachlan following the Bindian Orogeny (e.g., Willman et al., 2002). This resulted in deep to
shallow marine sedimentation (including carbonates) and widespread bimodal and felsic volcanism in
new and existing basins and rifts in the Central and Eastern Lachlan in both Victoria and New South
Wales (Meakin and Morgan, 1999; Lyons et al., 2000; VandenBerg et al., 2000, Willman et al., 2002;
Colquhoun et al., 2005; Glen, 2005). This Early Devonian sedimentation and volcanism is thought to
reflect post-Bindian orogenic extension, and appears to be reflected even in basins that do not record
the Bindian Orogeny (e.g., Pogson and Watkins, 1998; Meakin and Morgan, 1999; Glen, 2004).
Watkins (in Meakin and Morgan, 1999) suggested that the volcanism reflected a backarc environment.

The Bindian Orogeny appears to be absent from northeastern Tasmania and the Melbourne Zone of
Victoria (Seymour and Calver, 1995; VandenBerg et al., 2000). As a result, dominantly deep marine
sedimentation in both regions is largely continuous throughout this cycle, continuing from Benambran
times (e.g., VandenBerg et al., 2000; Fergusson, 2003; Seymour and Calver, 1995). In the Melbourne
Zone, sedimentation appears to shallow upwards and switch from marine to terrestrial, possibly in
response to the start of the Tabberabberan Orogeny, e.g., VandenBerg (2003). VandenBerg (2003)
also indicated a change in sediment transport direction and source in the Emsian with the appearance
of lithic and volcaniclastic detritus derived from the east. This has been taken as evidence for the
incoming arrival of the Benambra Terrane, possibly by strike-slip motion, to something approaching its
present position (VandenBerg et al., 2000; Willman et al., 2002). Prior to this, an open ocean appears
to have existed east of the Melbourne Zone (Whitelaw Terrane), that is, the Benambran and Whitelaw
terranes were previously separated—a conclusion agreed upon by many workers, regardless of
tectonic model (e.g., Gray, 1997; Gray and Foster, 1997, 2004; Fergusson, 2003, Spaggiari et al.,
2004). Similarly, detrital zircon data from sediments of this age in northeastern Tasmania indicate no

32 Geodynamic Synthesis of the Phanerozoic of eastern Australia


apparent sourcing of material from western Tasmania, which led Black et al. (2004) to suggest that
northeastern and western Tasmania were also separate at this time.

Rocks of mid Silurian to Late Devonian age also occur in the Delamerian Orogen. These include
terrestrial to marine sedimentary rocks in western Victoria, largely in the Grampians-Stavely Zone
(VandenBerg et al., 2000), and deep water clastic-dominated sedimentation in Western Tasmania
(Seymour and Calver, 1995, 1998; Figure 2.5, Figure 2.6, Figure 2.9). Sediments in Western Victoria
were apparently deformed between ca. 420–410 Ma (Bindian Orogeny?) and are overlain by post-
deformation Lower Devonian volcanic rocks and intruded by associated plutonism (VandenBerg et al.,
2000). The Bindian Orogeny appears to be absent in western Tasmania, although hiatuses in
sedimentation which may correspond to this orogeny are recorded (Seymour and Calver, 1995, 1998).

Widespread felsic-dominated magmatism occurs across the Western, Central and Eastern Lachlan
within this cycle (Lyons et al., 2000; Collins and Hobbs, 2001, Chappell and White; 1992; Rossiter,
2003; Black et al., 2005; Gray et al., 2003; Bodorkos et al., 2013, 2015; Chisholm et al., 2014b, c;
Figure 2.9). Crystallisation ages largely fall between ca. 430 Ma and 390 Ma but include younger
granites belonging to the Kanimblan cycle (e.g., Chappell and White, 1992; Gray et al., 2003; Black et
al., 2005). The oldest granites (ca. 435–425 Ma) in New South Wales and Victoria are dominantly S-
types in the Central and Eastern Lachlan (e.g., Collins and Hobbs, 2001; Willman et al., 2002; Ickert
and Williams, 2011; Bodorkos et al., 2013, 2015). These include a continuation of the magmatism that
commenced during the Benambran Orogeny (e.g., Collins and Hobbs, 2001). Recently Ickert and
Williams (2011) have confirmed a secular change in granite types, at least for the southern Eastern
Lachlan Orogen. They showed that S-type magmatism in that area largely falls between 435–425 Ma
with I-type magmatism mostly younger (mostly ca. 418 Ma), though they did find older I-type
magmatism contemporaneous with or slightly older than S-type magmatism. This secular change is
not universal, however, and Bodorkos et al. (2013) report younger S-type magmatism (ca. 418–415
Ma) in the Central Lachlan of NSW. Spatial trends in granite ages are also evident the Lachlan
Orogen. In the Eastern Lachlan of New South Wales and Victoria (Kuark and Mallacoota zones), ages
appear to decrease eastwards, with the youngest Devonian granites (Lewis et al., 1994; VandenBerg
et al., 2000; Gray et al., 2003) occurring in the Bega and Moruya batholiths (Bega basement terrane of
Chappell et al., 1988). In Victoria, granites appear to mostly decrease in age towards the Melbourne
Zone (VandenBerg et al., 2000; Gray et al., 2003; Rossiter, 2003). Granites east and west of the
Central Victorian Magmatic Province are largely ca. 420 Ma to 380 Ma in age (e.g., Gray et al., 2003).
Willman et al. (2002) suggested that these granites are post-tectonic in the Whitelaw Terrane, but, in
part, syn-tectonic (related to the Bindian Orogeny) in the Benambra Terrane. The youngest (mostly
post-Tabberabberan, ca. 385–350 Ma) rocks occur within the Middle Devonian to Early Carboniferous
Central Victorian Magmatic Province (Chappell et al., 1988; VandenBerg et al., 2000; Rossiter, 2003;
Figure 2.5), corresponding to the Melbourne and Bassian terranes of Chappell et al. (1988). A similar
diachronous trend is evident in Tasmania, where granite ages record a well-documented, pronounced
westward younging in crystallisation age from ca. 400–375 Ma, pre-, syn-, and post-Tabberabberan
granites in the northeast, to post-Tabberabberan granites, ca. 370–350 Ma in western Tasmania
(Black et al., 2005; Figure 2.6, Figure 2.9).

The large area of magmatism in the Lachlan Orogen, and the associated thermal requirements, are
problematical and has led many authors to speculate on possible tectonic scenarios. These include
multiple subduction zones (e.g., Collins and Hobbs, 2001; Soesoo et al., 1997; Gray and Foster,
1997), delamination (Collins and Vernon, 1994), mantle plumes (e.g., Wyborn, 1992; Cas et al., 2003),
as well as backarc extension (Collins and Richards, 2008). Both the mantle plume and multiple
subduction models, e.g., the double subduction model of Soesoo et al. (1997), have the advantage of
explaining the wide distribution of Lachlan Orogen granites. The multiple subduction zone models,
unlike the plume models, can explain the diachronous trends of the granites, but are, however, largely

Geodynamic Synthesis of the Phanerozoic of eastern Australia 33


not consistent with the bimodal or dominantly felsic nature of the magmatism. Cas et al. (2003)
suggested magmatism effectively ceased (went into hiatus) just before the Tabberabberan Orogeny.
Similarly, it is difficult to see how the plume model could explain the geographic spread and trends in
ages of magmatism.

The ca. 390–380 Ma east-west contractional Tabberabberan Orogeny (e.g., Gray and Foster, 1997,
2004; Spaggiari et al., 2003, 2004; Black et al., 2005), effectively cratonised the whole Lachlan
Orogen and resulted in much of the current configuration. This deformation, the first to affect most of
the Lachlan Orogen, is suggested to have been responsible for the final amalgamation of the terranes
and zones of the Lachlan, e.g., Whitelaw (Western Lachlan) and Benambran (Central and Eastern
Lachlan) terranes (Gray and Foster, 1997, Soesoo et al., 1997; VandenBerg et al., 2000; Willman et
al., 2002; Spaggiari et al., 2003, 2004), and western and eastern Tasmania (Seymour and Calver,
1995; Black et al., 2005). Interpretations for the drivers of this east-west contraction are varied, but
largely reflect the different interpreted tectonic models for the region. Gray and co-workers (e.g., Gray,
1997; Gray and Foster, 1997, 2004; Soesoo et al., 1997; Spaggiari et al., 2003, 2004) suggested that
this collisional event occurred ca. 400–390 Ma (Spaggiari et al., 2003), and was (at least partly)
related to the closure of a marginal basin (effectively the Melbourne Zone) and an end to double
divergent subduction (e.g., Gray and Foster, 1997; Soesoo et al., 1997). Conversely, Willman et al.
(2002) and Cayley et al. (2002) suggested the Tabberabberan Orogeny was responsible for ending
the relative strike-slip movement of the Whitelaw and Benambra terranes, and reflected docking of the
two terranes. In Tasmania, the Tabberabberan deformation (ca. 388 Ma; Black et al., 2005) may relate
to docking of northeastern Tasmania to western Tasmania, based on detrital zircon evidence (Black et
al., 2004). In the recent continental indenter model of Moresi et al. (2014) the Tabberabberan Orogeny
marks the end of orocline formation, and related relative movement of crustal blocks, and sees a
return to a simpler linear continental arc. It is also possible that the Tabberabberan deformation may
also relate (in part) to docking of the Calliope-Gamilaroi arc in the New England Orogen at this time
(Flood and Aitchison, 1992; Murray et al., 2003). In reality, it is possible that all of these models may
be partly correct, as each model better explains some, but not all, aspects of the Tabberabberan
geology. Despite this apparent complexity, it would appear that by the end of the Tabberabberan
Orogeny eastern Australia falls within a relatively simple overall backarc environment behind a west-
dipping slab and subduction zone located to the east (e.g., Gray, 1997; Glen et al., 1998; Cayley et al.,
2002; Gray and Foster, 2004; Glen, 2005; Moresi et al., 2014).

2.1.5.2 Geological history and Time-Space plot explanation (Figure 2.4 to Figure 2.9)

• Deposition of widespread, largely deep marine to shallow marine, sediments (including


carbonates) in dominantly north-south oriented troughs, platforms, shelves and other rift zones
(e.g., Cowra, Tumut, Rast, Jemalong, Hill End troughs, Cobar Basin, Walter Range, Mumbil
Shelves, Cowombat Rift, e.g., Pogson and Watkins, 1998; Meakin and Morgan, 1999; Lyons et al.,
2000; VandenBerg et al., 2000; VandenBerg, 2003; Colquhoun et al., 2005; Glen, 2005), across
the Lachlan Orogen in NSW and eastern parts of Victoria. Sedimentation was accompanied by
widespread, largely S-type, but also I-type and possibly A-type (Colquhoun et al., 2005), volcanism
and plutonism. Extension developed within and between the Ordovician turbidites and the
Macquarie Arc (Lyons et al., 2000; Glen et al., 2007b). Glen (2005) suggested that volcanism is
absent from those basins north of the Lachlan Transverse Zone.
• In Victoria, most deposition of this age was in the Melbourne Zone, continuing from pre-
Benambran times (VandenBerg et al., 2000; Fergusson et al., 2003). This consists of (latest
Ordovician–) Silurian to Middle Devonian dominantly deep marine turbiditic sedimentation of the
Murrindindi Supergroup (VandenBerg et al., 2000; VandenBerg, 2003), which is dominated by
mudstone-siltstone but is marked by episodic influxes of coarser clastic turbidites. Sedimentation

34 Geodynamic Synthesis of the Phanerozoic of eastern Australia


appears to shallow upwards (VandenBerg et al., 2000), with carbonates occurring towards the top
and, locally preserved, very thick terrestrial sedimentation (Cathedral Group) at the very top of the
Supergroup. VandenBerg (2003) also indicated a change in sediment transport direction and
source in the Emsian with the appearance of lithic and volcaniclastic detritus derived from the east.
• In western Victoria, largely in the Grampians-Stavely Zone, deposition of terrestrial to marine
sediments of the Grampians Group (VandenBerg et al., 2000; VandenBerg, 2003) occurred on top
of rocks of the Delamerian Orogen. Part of this sedimentation appears to be contemporaneous
with the Yalmy Group in eastern Victoria, that is, it is partly syn-Benambran in age (VandenBerg et
al., 2000; VandenBerg, 2003). The Grampians Group was apparently deformed ca. 420–410 Ma
and is overlain by post-deformation Early Devonian volcanic rocks (VandenBerg et al., 2000;
VandenBerg, 2003) and associated plutonism. This deformation has been equated with the
Benambran Orogeny (VandenBerg et al., 2000) but appears to be very similar in age to the
Bindian Orogeny in eastern Victoria. Miller et al. (2005) document a low strain southeast-northwest
contraction occurring at this time (ca. 420–414 Ma) in both the Grampians Group and in the
Stawell Zone.
• Marine turbiditic sedimentation in the Mathinna Supergroup in northeastern Tasmania continued
from the previous cycle, probably up to the Early Devonian, although age constraints for this
succession are poor (Seymour and Calver, 1995). The group was deformed in the Early Devonian
(Reed, 2001). Detrital zircon data from the Mathinna Supergroup indicates no sourcing of material
from western Tasmania (Black et al., 2004).
• Following the switch to deeper water clastic-dominated sedimentation (e.g., Eldon Group) around the
end of the Ordovician (Seymour and Calver, 1995, 1998), deposition in Western Tasmania continued
intermittently until the Middle Devonian. Both the Benambran and Bindian orogenies appear to be
absent within this terrane. Hiatuses are recorded in several zones (e.g., Sheffield Element), however,
which may correspond to one or both of these orogenies (Seymour and Calver, 1995, 1998).
• Widespread felsic-dominated magmatism, including the majority of granites in the orogen, occurred
across the Western, Central and Eastern Lachlan within this cycle (Meakin and Morgan, 1999; Lyons
et al., 2000; Collins and Hobbs, 2001, Chappell and White; 1992; Gray et al., 2003; Rossiter, 2003;
Black et al., 2005; Ickert and Williams, 2011; Bodorkos et al., 2013, 2015). Ages largely fall between
ca. 435 Ma and 390 Ma, though magmatism continued until the early Carboniferous (ca. 360–350
Ma), e.g., Chappell and White (1992), Gray et al. (2003), Black et al. (2005). Magmatic ages vary
according to both type and location. The oldest granites are dominantly S-types in the Central and
Eastern Lachlan (e.g., Collins and Hobbs, 2001; Willman et al., 2002; Ickert and Williams, 2011;
Bodorkos et al., 2013, 2015). Early Silurian magmatism of this age appears to be absent from the
Western Lachlan (Whitelaw Terrane) in Victoria (VandenBerg et al., 2000; Willman et al., 2002; Gray
et al., 2003), although some may occur in the Glenelg Zone (VandenBerg et al., 2000). Youngest
granites are found in the Eastern Lachlan of New South Wales and Victoria (Kuark and Mallacoota
zones), in the Melbourne Zone of Victoria (VandenBerg et al., 2000; Gray et al., 2003; Rossiter,
2003), and in western Tasmania (Black et al., 2005; Figure 2.9).
• Deformation related to the Bindian Orogeny appears to be largely confined to the southern parts of
the Central and Eastern Lachlan (e.g., Gray, 1997; Gray and Foster, 1997; Willman et al., 2002;
Gray et al., 2003; Glen, 2005). In eastern Victoria, ages of ca. 418–410 Ma (e.g., Gray and Foster,
1997; Gray et al., 2003) are largely expressed as angular unconformities, especially above the
early rift basins of the Tabberabberan cycle (VandenBerg et al., 2000; VandenBerg, 2003).
According to VandenBerg et al. (2000), there are no significant effects in western Victoria, although
Miller et al. (2005) have documented deformation of this age (ca. 420–414 Ma) in the Stawell
Zone, which they correlated with similar aged deformation in the Grampians Group in western
Victoria. In New South Wales, the Bindian Orogeny, where observed, appears to be of similar age

Geodynamic Synthesis of the Phanerozoic of eastern Australia 35


to that recorded in Victoria, e.g., Glen (2005) summarised evidence for deformation ages of ca.
418 Ma to 410 Ma. The deformation, however, is not recorded everywhere in New South Wales;
for example, apparently continuous sedimentation during this period is recorded in the Hill End,
Cowra and Rast Troughs and Walter Range Shelf (Pogson and Watkins, 1988; Meakin and
Morgan; 1999; Colquhoun et al., 2005). Elsewhere in New South Wales, the deformation may
correspond to breaks in sedimentation, e.g., between early and late Silurian sedimentation in the
Central Lachlan (e.g., Lyons et al., 2000), although no significant unconformities are recorded in
those areas. Importantly, there are unconformities of this age, e.g., in the Kandos and Queens
Pinch groups, which have been suggested to relate to local effects such as uplift associated with
granite intrusion (Meakin and Morgan, 1999). The Bindian Orogeny is also apparently not recorded
in Tasmania (e.g., Black et al., 2005) although lode gold mineralisation possibly of this age is
recorded in the Beaconsfield deposit (Bierlein et al., 2005).
• Renewed extension and rifting, in the Early Devonian, and deposition of deep to shallow marine
sedimentation (including carbonates) in both new and existing basins and rifts in the Central and
Eastern Lachlan in both Victoria and New South Wales, e.g., Hill End Trough, Buchan Rift (Meakin
and Morgan, 1999; Lyons et al., 2000; VandenBerg et al., 2000, Willman et al., 2002; Colquhoun et
al., 2005; Glen, 2005). This was accompanied by widespread bimodal and felsic volcanism, e.g.,
Gregra Group, Snowy River Volcanic Group (Pogson and Watkins, 1998; Meakin and Morgan, 1999;
VandenBerg et al., 2000; VandenBerg, 2003; Cas et al., 2003), as well as associated intrusions.
• During the late Early Devonian to Middle Devonian (ca. 400–380 Ma), the Lachlan Orogen
underwent approximately east-west contractional deformation and associated, mostly low-grade
metamorphism, related to the Tabberabberan Orogeny (Seymour and Calver, 1995, 1998; Gray,
1997; Gray and Foster, 1997, 2004; Meakin and Morgan, 1999; Lyons et al., 2000; VandenBerg et
al., 2000; Willman et al., 2002; Gray et al., 2003; Spaggiari et al., 2003; Black et al., 2005; Glen,
2005). In Victoria, the Tabberabberan deformation was responsible for the deformation and uplift of
the Melbourne Zone (Gray, 1997; VandenBerg et al., 2000; Willman et al., 2002; Spaggiari et al.,
2003). Black et al. (2005) record an age of ca. 388 Ma for this event in Tasmania. Reed (2001)
documented two apparent phases of Tabberabberan deformation in northeast Tasmania, with
opposite sense of vergence, and suggested that northeastern Tasmania better correlated with the
Tabberabbera Zone of Victoria.

2.1.6 Late Devonian to early Carboniferous (ca. 380 to ca. 350 Ma)

2.1.6.1 Geological and tectonic summary

The post-Tabberabberan time period in the Lachlan Orogen is marked by widespread extension, rifting
and accompanying basin formation, as well as significant extrusive and intrusive magmatism,
occurring within a now largely cratonic land mass (VandenBerg et al., 2000; Lyons et al., 2000; Glen,
2005; Figure 2.9). As outlined by Willman et al. (2002), for example, the post-Tabberabberan
sedimentary and volcanic rocks overlie major faults and interpreted suture zones belonging to the
Tabberabberan Orogeny. Like the earlier orogenic cycles, the post-Tabberabberan extension is
thought to reflect behind-arc processes following renewed roll-back, with the main subduction zone
and arc now found within the New England Orogen (e.g., Glen, 2005; Collins and Richards, 2008).

Rocks of this age include Late Devonian sediments accompanied by A-type and bimodal extrusive
and intrusive magmatism (e.g., Meakin and Morgan, 1999; Lyons et al., 2000; Lewis et al., 1994;
Wormald et al., 2004; Figure 2.4, Figure 2.5, Figure 2.9), probably related to initiation of post-
Tabberabberan extension and associated rifting (e.g., Lewis et al., 1994; Meakin and Morgan, 1999;
Lyons et al., 2000). This was followed by Lachlan-wide (New South Wales and Victoria, but not

36 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Tasmania) late Middle Devonian to Early Carboniferous, clastic, mostly continental, sedimentation,
including red beds, of the ‘Lambie facies’, reflecting continuing, more widespread extension (Lewis et
al., 1994; Warren et al., 1995; Meakin and Morgan, 1999; Lyons et al., 2000; VandenBerg et al., 2000;
Cas et al., 2003; Glen, 2004; Figure 2.9). These are also thought to be related to widespread post-
Tabberabberan extension (e.g., Lewis et al., 1994; Meakin and Morgan, 1999; Lyons et al., 2000). The
sediments are mostly quartz-rich (e.g., Meakin and Morgan, 1999; VandenBerg et al., 2000), and Glen
(2005) suggested they may have been sourced largely from central Australia, related to uplift
associated with the Alice Springs Orogeny. VandenBerg et al. (2000) suggest sediments of this age in
eastern Victoria were probably locally derived from ‘Tabberabberan highlands’.

Middle Devonian to earliest Carboniferous intrusive I-, S- and A-type magmatism between ca. 380 Ma
to 350 Ma (Chappell and White, 1992; Gray et al., 2003; Wormald et al., 2004; Black et al., 2005;
Bodorkos et al., 2010), occurred throughout the Eastern Lachlan, in the Central Victorian Magmatic
Province and across Tasmania (Figure 2.9). The youngest granites are found in the Melbourne Zone
of Victoria (VandenBerg et al., 2000; Gray et al., 2003; Rossiter, 2003), and in western Tasmania
(Black et al., 2005; see Tabberabberan section).

Extension was terminated by the early Carboniferous (ca. 360–340 Ma) contractional east-west
Kanimblan Orogeny (e.g., Gray, 1997; Meakin and Morgan, 1999; VandenBerg et al., 2000; Gray et
al., 2003; Glen, 2005). This orogeny folded and inverted Kanimblan cycle and older rocks. It occurred
across the Lachlan Orogen, into the Delamerian Orogen (e.g., Gilmore et al., 2007), but is best
expressed in the Eastern Lachlan (Gray, 1997; Gray et al., 2003; Glen, 2005).

There is little evidence for arc-related magmatism during this period in the Lachlan, with most
evidence clearly indicating that the arc was located further east in the New England Orogen (e.g.,
Meakin and Morgan, 1999; Glen, 2005; Collins and Richards, 2008).

2.1.6.2 Geological history and Time-Space plot explanation (Figure 2.4 to Figure 2.9)

• Middle Devonian sedimentation and locally voluminous, bimodal or felsic I- and/or A-type
magmatism within the Eastern Lachlan in New South Wales, e.g., Dulladerry Rift, Rocky Ponds
Group, Boyd Volcanic Complex, Mount Wellington Volcanic Group (Lewis et al., 1994; Warren et
al., 1995; Pogson and Watkins, 1998; Meakin and Morgan, 1999; Lyons et al., 2000).
• Late Middle to Late Devonian–Early Carboniferous clastic, shallow-marine to mostly continental,
sedimentation, including red beds, of the ‘Lambie facies’ occur across the Lachlan Orogen in
Victoria and New South Wales, e.g., Mulga Downs, Harvey, Catombal, Merimbula Groups, Avon
Supergroup, Combyingar Formation (Lewis et al., 1994; Warren et al., 1995; Meakin and Morgan,
1999; Lyons et al., 2000; VandenBerg et al., 2000; Cas et al., 2003; Glen, 2004; Wormald et al.,
2004). In places the sediments are accompanied by, locally voluminous, bimodal or felsic I- and/or
A-type magmatism, e.g., Mount Wellington Volcanic Group (VandenBerg et al., 2000; Cas et al.,
2003; Wormald et al., 2004). Angular unconformities and folding occur throughout this succession
in Victoria (e.g., VandenBerg et al., 2000). These may relate to local deformation or perhaps
extension. Sedimentary rocks of this age are rare in Tasmania (Seymour and Calver, 1995, 1998).
• Post-tectonic Late Devonian to earliest Carboniferous felsic-dominated intrusive and extrusive
magmatism occurred across the Lachlan Orogen during this cycle (Chappell and White; 1992;
Rossiter, 2003; Black et al., 2005; Gray et al., 2003; Bodorkos et al., 2010; Figure 2.9). Granites
include I-, S- and, relatively common, A-types (Collins et al., 1982; Chappell et al., 1988; Rossiter,
2003; Wormald et al., 2004). Ages mostly fall between ca. 380 Ma to 360 Ma, but magmatism
continued locally until the early Carboniferous (ca. 360–350 Ma), e.g., Chappell and White (1992),
Gray et al. (2003), Wormald et al. (2004), Black et al. (2005), Bodorkos et al. (2010). The youngest

Geodynamic Synthesis of the Phanerozoic of eastern Australia 37


magmatic rocks occur in the Central Victorian Magmatic Province (VandenBerg et al., 2000;
Rossiter, 2003), and western Tasmania (Black et al., 2005).
• Latest Devonian to early Carboniferous east-west shortening and associated low-grade regional
metamorphism of the Kanimblan Orogeny in New South Wales and Victoria (e.g., Gray et al.,
2003; Glen, 2005). This is best expressed in the Eastern Lachlan, with folding and inversion of the
Middle to Late Devonian basins (Pogson and Watkins, 1998; Meakin and Morgan, 1999; Lyons et
al., 2000; VandenBerg et al., 2000; Gray et al., 2003; Glen, 2005), although it extends west to the
Delamerian Orogen (e.g., Gilmore et al., 2007). Glen (2005) suggests an age of ca. 340 Ma for this
deformation in New South Wales. In Victoria the timing is poorly constrained, but must be post-
earliest Carboniferous (e.g., VandenBerg et al., 2000). Gray (1997) and Gray and Foster (1997)
suggests ages of ca. 360–340 Ma for the Kanimblan Orogeny.

2.1.7 Middle Carboniferous to earliest Triassic (ca. 350 Ma to 250 Ma)

2.1.7.1 Geological and tectonic summary

The Kanimblan Orogeny was the terminal event in the Lachlan Orogen (e.g., Pogson and Watkins,
1998), and subsequent geology appears to relate more to the New England Orogen to the east. From
the latest Carboniferous to Permian, Eastern Australia was dominated by tectonic extension and
rifting, and many intracratonic basins were initiated in this time, as was the backarc Sydney-
Gunnedah-Bowen basin system (see Thomson Orogen section). Within Victoria and New South
Wales, the only significant magmatic event was the Carboniferous (ca. 340–310 Ma) intrusive
magmatism recorded as a north-northwest belt in the northeastern Lachlan (e.g., Pogson and
Watkins, 1998; Meakin and Morgan, 1999; Figure 2.4, Figure 2.9) and east of Canberra (e.g., Bryan et
al., 1966; Rose, 1966; Bodorkos et al., 2010). This magmatism includes mostly felsic I-type granites of
the Bathurst Batholith and the Gulgong Suite (Bathurst Terrane of Chappell et al., 1988). They are of
similar age and geochemistry to volcanic and intrusive rocks in the New England Orogen (e.g.,
Chappell et al., 1988; Pogson and Watkins, 1998), and may relate to continental arc formation,
although Meakin and Morgan (1999) suggest emplacement in an extensional environment,
presumably behind the continental arc of the New England Orogen. The eastern part of the Lachlan
Orogen is overlain by the latest Carboniferous to Triassic Sydney-Gunnedah-Bowen basin system,
which initially developed as a backarc rift behind the New England Orogen (e.g., Korsch et al., 2009a;
see New England Orogen section). The basin rocks overlie the Carboniferous granites of the Bathurst-
Gulgong area. Of similar age to the Sydney-Gunnedah-Bowen Basin are the Parmeener Supergroup
sediments of the Tasmania Basin in Tasmania (Figure 2.6, Figure 2.9), which developed over the
suture (Tamar Fracture) between western and northeastern Tasmania (Seymour and Calver, 1995,
1998). These consist of a lower succession of glacial and marine sediments and an upper succession
(Late Permian and younger) of non-marine sediments including coal measures (Seymour and Calver,
1995, 1998). Remnants of possibly more widespread Permian glacial and marine sedimentation are
also recorded (outcrop and sub-surface, e.g., beneath the Murray Basin) in Victoria and southern New
South Wales (O’Brien et al., 2003; Figure 2.9).

38 Geodynamic Synthesis of the Phanerozoic of eastern Australia


2.2 North Queensland region
by DC Champion

2.2.1 Introduction
The north Queensland region contains the Mossman Orogen (Hodgkinson and Broken River
provinces), the northern extension of the Thomson Orogen (Charters Towers, Barnard, Greenvale and
perhaps Iron Range provinces), which occur marginal to the outcropping Paleoproterozoic to
Mesoproterozoic Etheridge, Savannah and Croydon provinces of the North Australian Element (also
North Australian Craton; Figure 2.10, Figure 2.11). The whole region has been the site of extensive
Paleozoic (intrusive and extrusive) magmatism belonging to the Macrossan (Cambrian–Ordovician),
Pama (Silurian–Devonian) and Kennedy (Cambrian–Permian) Igneous Associations (also called
igneous provinces), and overlain by Paleozoic and younger basins, including the widespread
Mesozoic rocks of the Carpentaria, Eromanga and Laura basins (see Jell, 2013). For this reason, the
geological synthesis for this area is discussed in terms of geological regions. These largely
correspond to one or more provinces but include (especially magmatic and cover) rocks of other ages
(compare Figure 2.10 and Figure 2.11). The Georgetown and Coen regions (comprising the
Paleoproterozoic to Mesoproterozoic Etheridge, Savannah and Croydon provinces of the North
Australian Element) are also included in this discussion of the north Queensland region. This is
because a considerable part of the Paleozoic tectonic regime (Tasman Orogen) and related crustal
development in north Queensland was focused across this old continental margin. The time-space plot
for the north Queensland region is shown in Figure 2.12.

2.2.2 Late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma):
Rodinian break-up

2.2.2.1 Geological and tectonic summary

Rocks of this age in north Queensland are dominated by metasedimentary sequences. They are best
represented in the Greenvale Province and Charters Towers region, but also occur in the Georgetown
and Coen regions and in the Barnard Province (Figure 2.12, Figure 2.13). Most sedimentary
successions thought to be early Cambrian or older, however, have poor age constraints (mostly based
on maximum depositional ages or interpreted metamorphic ages), and recent dating (e.g., Kositcin et
al., 2015b) indicates younger components may be present in at least some. The tectonic environment
for this period is typically interpreted as a passive margin, related to (but post-dating) Rodinian break-
up. Fergusson et al. (2007a, b) suggested the presence of Late Neoproterozoic rifting—ca. 600 Ma—
in the region, based on the presence of common 600–500 Ma detrital zircons and on correlation with
other regions, e.g., the Thomson Orogen. The recent results of Fergusson et al. (2007a) also indicate
that the Tasman Line, as previously defined, along the eastern margin of the Greenvale Province,
does not represent Rodinia breakup, but rather younger (Benambran?) west-directed thrusting of
younger rocks over Mesoproterozoic basement.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 39


Figure 2.10 Distribution of geological provinces in northern Queensland. Province names and boundaries are as
defined by Jell (2013). The magmatic Macrossan (Cambrian–Ordovician), Pama (Silurian–Devonian) and
Kennedy (Carboniferous–Permian) Igneous Associations are not shown. In the Time-Space plots for northern
Queensland, provinces (Figure 2.11) are grouped into regions (see Figure 2.12)

2.2.2.2 Geological history and Time-Space plot explanation (Figure 2.10 to Figure 2.13)

Greenvale region (Greenvale and Broken River provinces)

• Metasedimentary-dominated Halls Reward Metamorphics (Withnall et al., 1997a). The age is


uncertain. Nishiya et al. (2003) suggested a Cambrian or Neoproterozoic age, based on
interpreted Cambrian metamorphic ages (ca. 520–510 Ma) and Neoproterozoic detrital zircon
ages. More recently, Kositcin et al. (2015b) found Ordovician maximum depositional ages of ca.
470 Ma, suggesting the unit may be, at least in part, significantly younger.
• Mafic to ultramafic magmatic rocks of the Boiler Gully and Gray Creek Complexes, intimately
associated with the Halls Reward Metamorphics. The unit may be tectonically emplaced and/or
intrusive. Withnall et al. (1997a) suggested it was intrusive, at least in part. Ages are poorly
constrained. Preliminary Nd-Sm data from mafic to ultramafic rocks yielded an apparent age of 466
± 37 Ma (D Huston and R Maas, unpublished data).
• Late Neoproterozoic-Early Paleozoic sedimentation (dolomitic carbonate and quartzofeldspathic
sediments) of the Oasis Metamorphics. Most probably shallow-marine environment formed as part
of the passive Gondwanan margin (Withnall et al., 1997a; Fergusson et al., 2007a). The Oasis
Metamorphics include tholeiitic mafic igneous rocks (intrusive/extrusive; e.g., Withnall et al.,
1997a). Age constrained to between 540–520 Ma (youngest detrital zircon population) and ca. 485

40 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Ma (overprinting metamorphism; Fergusson et al., 2007a). Recent zircon and titanite dating of
calc-silicate rocks suggests that at least part of the Oasis Metamorphics is in fact late
Paleoproterozoic in age (Cross et al., 2015).

Georgetown region

• Terrestrial sedimentary rocks of the Inorunie Group (Withnall et al., 1997a). Ages are poorly
constrained, and the group may be as old as Mesoproterozoic, the preferred age of Withnall et al.
(1997a).

Figure 2.11 Distribution of geological regions as used for the north Queensland synthesis discussions and Time-
Space Plot (Figure 2.12). Geological regions are largely based on the geographic subdivisions used in Bain and
Draper (1997). Regions have been used here to group geological provinces and because of the widespread
Macrossan (Cambrian–Ordovician), Pama (Silurian–Devonian) and Kennedy (Carboniferous–Permian) igneous
associations which occur across north Queensland and within all of the sedimentary/metamorphic provinces.

Barnard region

• Barnard Metamorphics (metasedimentary rocks, amphibolite, metavolcanics) and the intrusive(?)


Babalangee Amphibolite (Bultitude et al., 1997). Ages are poorly constrained but for the Barnard
Metamorphics must be older than Early-Middle Ordovician (ca. 465–490 Ma), the reported range
of ages for cross-cutting granites (Bultitude et al., 1997; Fergusson and Henderson et al., 2013).
The Babalangee Amphibolite may be of this age, as Richards et al. (1966) reported a 642 Ma K-Ar
age for the unit, though the significance of this age is uncertain, and the unit may be significantly

Geodynamic Synthesis of the Phanerozoic of eastern Australia 41


younger. Kositcin et al. (2015a) recorded a ca. 452 Ma age, which they interpreted as
metamorphic, for an amphibolite unit, of uncertain protolith, within the Barnard Metamorphics.

Coen region

• Sefton Metamorphics – poorly age-constrained metasedimentary rocks (Blewett et al., 1997) and
mafic magmatic rocks. Detrital zircons indicate the succession is younger than ca. 1200 Ma
(Blewett et al., 1997). Recent detrital zircon geochronology by Kositcin et al. (2015b) and Cross et
al. (2015b), indicate late Neoproterozoic ages, suggesting that these rocks may represent a
northern continuation of the Thomson Orogen as suggested by Purdy et al. (2013), and as shown
in Figure 1.1.

Charters Towers region

• Widespread remnants of largely marine metasedimentary rocks (Charters Towers Metamorphics,


Argentine Metamorphics, Running River Metamorphics, Cape River Metamorphics; Hutton et al.,
1997). These units also include mafic magmatic rocks, which appear to include both alkaline and
tholeiitic compositions (Hutton et al., 1997; Fergusson et al., 2007b). Ages are not well constrained
except for the Argentine Metamorphics which Fergusson et al. (2007b) constrained to have been
deposited between ca. 500 Ma and ca. 480 Ma. The Cape River Metamorphics give similar
minimum ages, based on intrusive contacts (> ca. 490 Ma; Hutton et al., 1997). Cross et al. (2015)
record a maximum depositional age of ca. 511 Ma for the Charters Towers Metamorphics.

2.2.3 Early to latest Cambrian (ca. 515 to 490 Ma)

2.2.3.1 Geological and tectonic summary

Rocks of this age are poorly represented in north Queensland, partly reflecting the geochronological
uncertainty of many of the units. The best evidence appears to be in the Greenvale Province (reported
metamorphic ages of ca. 520–510 Ma; Nishiya et al., 2003; Figure 2.13) and in the Charters Towers
region (with ages of metamorphism of between ca. 495 Ma and ca. 475 Ma; Fergusson et al.; 2007a).
Potential Delamerian Orogeny deformation may occur in the Georgetown and Coen regions but
geochronological control is missing. A number of deformations that could be interpreted as
Delamerian in age have been shown recently, at least partly, to represent post-Delamerian extension
(Fergusson et al., 2007a, b).

2.2.3.2 Geological history and Time-Space plot explanation (Figure 2.10 to Figure 2.13)

Greenvale region

• The Halls Reward Metamorphics have Cambrian metamorphic ages of ca. 510–500 Ma (Nishiya et
al., 2003), consistent with Delamerian deformation. Fergusson et al. (2007a, 2013) and Fergusson
and Henderson et al. (2013) did not, however, record evidence for Delamerian deformation in the
nearby Oasis Metamorphics. Deformation in this unit appears to post-date the Delamerian
Orogeny, and be related to extension. Fergusson and Henderson et al. (2013) indicate an Early
Ordovician (~475 Ma) age for this deformation and metamorphism; while Cross et al. (2015) record
a younger ca. 456 Ma metamorphic age.
• Rocks of the Boiler Gully and Gray Creek Complexes appear to have similar deformation histories
as the Halls Reward Metamorphics, and so the deformation observed in these complexes is
inferred to be Delamerian in age.

42 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Georgetown region

• Poorly age-constrained, widespread, north-south contractional(?), deformation and associated


metamorphism (largely retrogressive) in the eastern Georgetown region (Withnall et al., 1997a;
Fergusson et al., 2007a). According to Withnall et al. (1997a) deformation is constrained to
between ca. 970 Ma and early Paleozoic.

Barnard region

• The Barnard Metamorphics record a deformation and metamorphic event (locally high-grade) not
observed in the Early Ordovician granites that intrude the metamorphic rocks (Garrad and
Bultitude, 1999).

Coen region

• Poorly age-constrained deformation and associated metamorphic (greenschist or lower grade)


events. According to Blewett et al. (1997) this deformation is constrained to between ca. 1550 Ma
and Devonian (granite emplacement) for most of the Coen Region, but between ca. 1130 Ma and
Carboniferous for the Iron Range region (Sefton Metamorphics). If the suggested Thomson orogen
affiliation for the Sefton Metamorphics is correct, the metamorphism and deformation there is
probably no older than Delamerian. The major fabric in the Sefton Metamorphics is east-west
(Blewett et al., 1997). Metamorphism is strongest near the Devonian granites (R. Blewett, pers.
comm., 2008), suggesting that the metamorphism may be this age.

Charters Towers region

• Fergusson et al. (2007b) infer a ca. 495 Ma contractional deformation event with associated
metamorphism in the Argentine Metamorphics (largely overprinted by slightly younger extension
and granite intrusion). Hutton et al. (1997) and Fergusson et al. (2007b) document similar
deformation in the Charters Towers, Running River, and Cape River Metamorphics.

2.2.4 Latest Cambrian to early Silurian (ca. 490 Ma to ca. 430 Ma)

2.2.4.1 Geological and tectonic summary

This time period in north Queensland is dominated by three general supracrustal successions, and a
major magmatic province:

• Lower to Middle Ordovician volcanic- or volcaniclastic-dominated assemblages with a calcalkaline


signature, interpreted as backarc successions, e.g., Seventy Mile Range Group in the Charters
Towers Province, and the Balcooma Metavolcanic Group and part of the Lucky Creek
Metamorphic Group in the Greenvale Province (Figure 2.12, Figure 2.13).
• Deep water (turbiditic), dominantly quartz-rich sediments, locally with tholeiitic magmatic rocks,
e.g., part of the Lucky Creek Metamorphic Group (Figure 2.12, Figure 2.13). Similar rocks occur in
the Judea Formation though the bimodal volcanics include rocks with calcalkaline signatures
(Withnall et al., 1997b).
• Calcalkaline volcanic and carbonate-dominated successions interpreted, in part, as island or
continental arc successions, e.g., the Lucky Creek Assemblage in the Broken River region
(Henderson et al., 2011, 2013).
• Mafic to felsic magmatic rocks of the Macrossan Igneous Association of Bain and Draper (1997).
These rocks are widely distributed throughout north Queensland but are best represented in the
Charters Towers region. Magmatic ages range from ca. 490 Ma to ca. 455 Ma (e.g., Hutton et al.,

Geodynamic Synthesis of the Phanerozoic of eastern Australia 43


1997; Hutton in Fergusson and Henderson et al., 2013; Figure 2.12, Figure 2.13). Dominated by I-
type and mantle-derived magmatism, though some S-type magmatism has been recorded. The
magmatism includes isotopically-primitive arc- or backarc-related rocks, such as in the Greenvale
region and eastern parts of the Georgetown region (Champion, 2013). Younger syn-Benambran
Orogeny felsic magmatism (ca. 430 Ma), belonging to the Pama Igneous Association, also occurs,
largely in the Georgetown and Greenvale regions.

Rocks of this cycle have long been interpreted as a dismembered continental margin (e.g., Henderson,
1987). Many authors have suggested backarc, continental or island-arc affinities (e.g., Withnall et al., 1991,
1997b; Henderson, 1986; Stolz, 1994; Henderson et al., 2011), suggesting an environment not dissimilar to
Lachlan Orogen rocks of the same age (e.g., Gray and Foster, 2004; Glen, 2005). Like the Lachlan
Orogen, north Queensland rocks of this age are either quartz-rich sediments or calcalkaline volcanics.
Notably, however, unlike the Lachlan Orogen, there are north Queensland units, such as the Judea
Formation (Withnall and Lang, 1993), which contain both quartz-rich marine sediments and calc-alkaline
volcanics, suggesting proximity between arc-related volcanism and craton-derived sedimentation.

Many of the north Queensland rocks were deformed in the Early Silurian (Figure 2.12) by a shortening
event coupled with metamorphism—referred to as the Benambran Orogeny by Fergusson et al.
(2007a). Evidence for this deformation is found in the Georgetown region where it is constrained by
ca. 430 Ma magmatic ages for syn-deformational I-type magmatism of the Pama Igneous Association
(Withnall et al., 1997a; Fergusson et al., 2007a). A similar deformation is recorded in the Charters
Towers region, possibly ca. 440 Ma (Fergusson et al., 2007b). Deformation of this age appears to be
largely absent in both the Hodgkinson and Broken River provinces, although Fergusson et al. (2007a)
suggested that island-arc terranes within the Camel Creek Subprovince (Broken River Province)—the
Everetts Creek Volcanics and Carriers Well Formation—were accreted at this time. Based on
geological relationships, Henderson et al. (2011) have suggested that this accretion occurred in the
late Early Silurian, just prior to deposition of the Crooked Creek Conglomerate at ca. 435 Ma.
Contractional deformation is present within the Hodgkinson and Broken River provinces but appears
to be earlier, probably Late Ordovician (Figure 2.12). Garrad and Bultitude (1999) have suggested
there may be a time break between Ordovician and Silurian rocks in the Hodgkinson Province that
corresponds to uplift (Benambran Orogeny) in the Georgetown region to the west. Recent work by Ali
(2010) on rocks of the Balcooma Metavolcanic Group has demonstrated Benambran deformation
produced continual crustal thickening (and increasing temperatures and pressures) in the period ca.
443–425 Ma, consistent with the timing of the accretion model of Henderson et al. (2011). Ali (2010)
also indicated substantial Early Devonian exhumation of these rocks.

Fergusson et al. (2007a, b) documented extensional deformation coupled with low-P high-T
metamorphism, greenschist to amphibolite-facies, in both (interpreted) backarc successions in the
Greenvale and Charters Towers regions (Figure 2.12). They dated these events at ca. 475 Ma and
480 Ma, but suggested extension was in operation from ca. 490 Ma to 460 Ma. This is supported by
metamorphic ages of ca 467 Ma on the Charters Towers Metamorphics (Cross et al., 2015), ca. 455
Ma for the Oasis Metamorphics (Cross et al., 2015) and ca. 452 Ma for the Barnard Metamorphics
(Kositcin et al., 2015a). The metamorphism and extension was largely synchronous with granite
emplacement, and most of the Macrossan Province magmatism is of this age. Importantly, the
interpreted backarc volcanic rocks in the Charters Towers region and Greenvale Province have quite
different orientations, ~east-west versus north-northeast–south-southwest, respectively (e.g., Bain and
Draper, 1997). This clearly suggests either some later relative movement between the regions, either
due to deformation (e.g., Bell, 1980; Fergusson et al., 2007a), and/or perhaps the volcanism formed
independently on different crustal fragments. Regardless, given a backarc origin for the volcanic rocks
in the southern Charters Towers region, it is possible that Macrossan Province magmatism in the
northern part of that region represents the actual magmatic arc (e.g., Henderson, 1980).

44 Geodynamic Synthesis of the Phanerozoic of eastern Australia


2.2.4.2 Geological history and Time-Space plot explanation (Figure 2.10 to Figure 2.13)

Greenvale region

• Metasedimentary to metavolcanic Lucky Creek Group and Balcooma Metavolcanic Group. Both
groups include calcalkaline volcanics though the Lucky Creek Group also contains tholeiitic rocks
(Withnall et al., 1997a). The Balcooma Metavolcanic Group is dated at ca. 470 Ma to 480 Ma (e.g.,
Withnall et al., 1991). Age is uncertain for the Lucky Creek Group, but it is thought to be the same
age, at least in part, as the Balcooma Metavolcanic Group. Withnall et al. (1991) originally
suggested the calcalkaline volcanics were part of a volcanic arc, while Fergusson et al. (2007a)
suggested a backarc environment. The latter is probably more consistent with similar
interpretations for the Charters Towers area (Draper and Bain, 1997). Recent dating by Kositcin et
al. (2015b) suggests that, at least part of, the Halls Reward Metamorphics are of this age.
Preliminary Sm-Nd data from mafic to ultramafic rocks of the Gray Creek Complex suggest a
similar age of 466 ± 37 Ma (D Huston and R Maas, unpublished data).
• Ordovician extension and related metamorphism and granite magmatism (Macrossan Igneous
Association) dated at ca. 485 Ma to 477 Ma (Fergusson et al., 2007a), to ca. 457 Ma (Cross et al.,
2015).
• Benambran age east-west shortening and metamorphism, and associated syn-deformation I-type
plutonism (Withnall et al., 1997a; Fergusson et al., 2007a); granite magmatism (Dido Supersuite)
has been dated at ca. 430–425 Ma (Withnall et al., 1997a). Ali (2010) recorded metamorphic ages
of ca. 443–425 Ma for the Balcooma Metavolcanic Group, and demonstrated that these rocks
recorded increasing temperatures and pressures through this period, i.e., crustal thickening.

Georgetown region

• Benambran east-west shortening, metamorphism, and syn-tectonic I-type granite magmatism


dated at ca. 435–417 Ma (Withnall et al., 1997a; Murgulov et al., 2007, 2009; Bultitude et al., in
Henderson et al., 2013). A second deformation event, tentatively dated at ca. 400 Ma (see
Section 2.2.5), may also form part of the Benambran event (Withnall et al., 1997a).

Barnard region

• Macrossan Province S- and I-type magmatism, dated at ca. 485 Ma and 455 Ma (Garrad and
Bultitude, 1999; R Wormald, James Cook University, pers. comm. June 2011, cited in Fergusson
and Henderson et al., 2013).
• Regional deformation and associated low to, possibly high-grade metamorphism with poor age
constraints (Bultitude et al., 1997). The deformation event affects the local granites, implying a
maximum age of ca. 460 Ma (Garrad and Bultitude, 1999). Kositcin et al. (2015a) recorded a ca.
452 Ma metamorphic age on an amphibolitic layer within the Barnard Metamorphics.

Coen region

• This region appears to record no definitive evidence of a Silurian deformation event. A poorly age-
constrained deformation event with associated metamorphism (greenschist or lower), between ca.
1130 Ma and the Devonian (Blewett et al., 1997), may be Silurian in age. A later deformation event
is also associated with Pama Igneous Association magmatism, though unlike elsewhere in
northern Queensland, this magmatism, and hence deformation also, is younger—ca. 407 Ma
(post-Benambran) (Blewett et al., 1997).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 45


Figure 2.12. Late Neoproterozoic to Cretaceous time-space plot for the north Queensland region, covering the Mossman Orogen, parts of the Thomson Orogen, and the
Proterozoic basement of the North Australia Element to the west. Refer to text for data sources and discussion. Regions are as outlined in Figure 2.10 and Figure 2.11.

46 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Charters Towers region

• The latest Cambrian(?) to largely Early Ordovician Seventy Mile Range Group, consisting of a
lower succession of marine metasedimentary rocks (with little volcanic input) and an upper
assemblage dominated by calcalkaline mafic to felsic volcanics and volcaniclastics (e.g.,
Henderson, 1986; Hutton et al., 1997). The group is commonly interpreted as forming in a backarc
environment (e.g., Henderson, 1986; Stolz, 1994).
• Deformation and associated metamorphism related to both extension (post-Delamerian) and
younger north-south Benambran contractional deformation (Berry et al., 1992; Hutton et al., 1997;
Fergusson et al., 2005, 2007a, b; Cross et al., 2015).
• Magmatic rocks of the Macrossan Igneous Association. In the Charters Towers region these
include granites of the widespread I-type Hogsflesh Creek and Lavery Creek supersuites, as well
as mafic intrusives. Hutton et al. (1997) suggested that the intrusives fall into 2 ages: Late
Cambrian to Early Ordovician, and mid Ordovician. The mafic intrusives are, in general, poorly
constrained and may be as young as Devonian. The majority of granites post-date the Benambran
deformation (Fergusson et al., 2005).

Hodgkinson region

• Marine, quartz-rich, turbiditic sediments with metabasalt of the Mulgrave Formation, preserved in
fault-bounded blocks along the Palmerville Fault, thought to be Early Ordovician in age (Bultitude
et al., 1997; Garrad and Bultitude, 1999).
• Limestone, and quartzofeldspathic deep water sediments, of Late Ordovician age, preserved in
fault-bounded blocks along the Palmerville Fault (Bultitude et al., 1997; Garrad and Bultitude,
1999). Dacitic volcanic clasts, in conglomerate (Mountain Creek Conglomerate), have been dated
at ca. 455 Ma (Garrad and Bultitude, 1999). Garrad and Bultitude (1999) suggested correlation of
these units with the limestone and volcanics of the Carriers Well Formation and Everett Creek
Volcanics (Broken River region).
• Middle(?) Ordovician east-west shortening. This is only evident in the Early Ordovician(?) Mulgrave
Formation (Garrad and Bultitude, 1999), and so is constrained to be older than Late Ordovician.

Graveyard Creek and Camel Creek regions

• Marine, quartz-rich, turbiditic sediments, commonly with tholeiitic metabasalt, in both the
Graveyard Creek and Camel Creek subprovinces (Withnall and Lang, 1993; Withnall et al., 1997b).
Ages are generally poorly constrained but thought to be Ordovician (Withnall and Lang, 1993;
Withnall et al., 1997b), though may extend into the Silurian. The Judea Formation appears to be of
Early Ordovician age (Withnall and Lang, 1993). Like the Lucky Creek Group in the Greenvale
region, the Judea Formation contains both tholeiitic and calcalkaline igneous signatures, and
includes felsic volcanic rocks.
• Calcalkaline volcanics, volcaniclastics and limestone in both the Graveyard Creek and Camel
Creek subprovinces (Arnold and Fawckner, 1980; Withnall and Lang, 1993; Withnall et al., 1997b).
Ages are variable; Early Ordovician in the Graveyard Creek Subprovince, and Late Ordovician in
the Camel Creek Subprovince. The calcalkaline rocks of the latter subprovince—the Lucky Creek
Assemblage of Henderson et al. (2013)—have been suggested by Fergusson et al. (2007a) and
Henderson et al. (2011, 2013) to represent island-arc remnants. A number of units appear to
contain both quartz-rich marine sediments and calcalkaline arc material (Withnall and Lang, 1993),
suggesting if an arc was present then it was probably proximal to the continent.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 47


48 Geodynamic Synthesis of the Phanerozoic of eastern Australia
Figure 2.13 Generalised distribution of rocks in the north Queensland region, by time period. A: late
Neoproterozoic to latest Cambrian (ca. 600 Ma to ca. 490 Ma), B: latest Cambrian to early Silurian (ca. 490 to ca.
430 Ma), C: Middle Silurian to Late Devonian (ca. 430 Ma to ca. 380 Ma), D: Late Devonian to early
Carboniferous (ca. 380 Ma to ca. 350 Ma), E: Middle Carboniferous to late Permian (ca. 350 Ma to ca. 270 Ma),
F: late Permian to middle Triassic (ca. 270 Ma to ca. 220 Ma)

Geodynamic Synthesis of the Phanerozoic of eastern Australia 49


• Minor sodic I-type granitic magmatism in the Graveyard Creek Subprovince, most probably
Ordovician in age (Withnall and Lang, 1993). Henderson et al. (2013) report unpublished ages of
482–488 Ma for one of these units, the Saddington Tonalite. These units are all likely to be of very
similar ages given their similar primitive isotopic signatures (Champion, 2013); the latter also
support a primitive arc origin.
• As summarised by Arnold and Fawckner (1980) and Withnall and Lang (1993), the Graveyard
Creek and Camel Creek subprovinces appear to have different structural histories. The Judea
Formation (Graveyard Creek Subprovince) records subhorizontal, mélange-type deformation and
low grade metamorphism, that is no younger than Late Ordovician (Withnall and Lang, 1993;
Withnall et al., 1997b) and may be syndeformational (e.g., Arnold and Fawckner, 1980). Withnall et
al. (1997b) also record mélange-style deformation and low-grade metamorphism in the Camel
Creek Subprovince. The age, however, is poorly constrained and Withnall et al. (1997b) suggested
much of this deformation occurred in the Devonian. This deformation may equate with similar aged
deformation in the Hodgkinson Province (Figure 2.13).

2.2.5 Middle Silurian to Late Devonian (ca. 430 Ma to ca. 380 Ma)

2.2.5.1 Geological and tectonic summary

This time period in north Queensland essentially equates to the Mossman Orogen as recently defined
by Henderson et al. (2013). The period is characterised by extensive, probably related, sedimentation
in the Hodgkinson and Broken River provinces (along the eastern and southeastern margins of the
Proterozoic Georgetown region), and also in the Charters Towers region (Figure 2.13). Sedimentation
includes marine siliciclastic sediments and carbonates, along with locally abundant, tholeiitic mafic
volcanics (Arnold and Fawckner, 1980). Sediment provenance is dominantly cratonic (e.g., Bultitude et
al., 1997) but does include Middle Ordovician volcaniclastic material. For example, Garrad and
Bultitude (1999) record dacitic clast ages of 465 Ma in a conglomerate from the Hodgkinson Province.
The geodynamic environment for this sedimentation is controversial. Models include both backarc or
forearc deposition, along with rifted continental margin (see summaries in Arnold and Fawckner, 1980;
Garrad and Bultitude, 1999; Henderson et al., 2013). Arnold (in Arnold and Fawckner, 1980),
Henderson et al. (1980) and Henderson (1987), amongst others, suggested that the Hodgkinson and
Broken River Province sedimentation was part of a forearc and accretionary wedge. This model,
however, has difficulties explaining tholeiitic volcanism, especially in the Hodgkinson Province. The
latter is more consistent with rift or backarc models (e.g., Fawckner in Arnold and Fawckner, 1980;
Bultitude et al., 1997), though could also reflect accreted oceanic crust. Resolution of the geodynamic
environment is critical to understanding the tectonics of the widespread Pama Igneous Association
(previously Pama Province) which comprises all the Silurian to Devonian magmatism in north
Queensland (Bain and Draper, 1997) and forms an essential component of the Mossman Orogen
(Henderson et al., 2013). The Pama Igneous Association forms an extensive quasi-continuous but
distinctly curved belt around the Hodgkinson and Broken River provinces, from Charters Towers in the
south where it strikes east-west, to a pronounced north-south striking belt in Cape York (Figure 2.13).
In forearc models, this belt is interpreted as the magmatic arc (e.g., Henderson, 1987), although, as
pointed out by numerous authors, the chemistry of this magmatism, especially in the Coen region, is
not consistent with a magmatic arc (e.g., Blewett et al., 1997). Pama Igneous Association magmatism
in the region is diachronous. Magmatic ages range from ca. 435–420 Ma (syn-Benambran and
younger) in the Georgetown region, to ca. 425–405 Ma (and younger) in the Charters Towers region,
to ca. 410–395 Ma in the Coen region (see Bultitude et al. (pages 280–295) in Henderson et al.,
2013). As pointed out by Champion and Bultitude (2003), these age differences are also matched by
changes in geochemical signature. The early Pama magmatism (in the Georgetown region) does have

50 Geodynamic Synthesis of the Phanerozoic of eastern Australia


geochemical signatures more consistent with arc magmatism. It is possible, therefore, that the
Hodgkinson (and Broken River) Province was in a forearc environment early (ca. 430 Ma), and then
evolved in a backarc environment (ca. 420–400 Ma and younger?). More recent geochronology
(Carson et al., 2011) shows that the Pama Igneous Association extends considerably westward,
matching the wide east-west extents seen in similar aged magmatism further south in the Lachlan and
Thomson orogens (e.g., Chappell and White, 1992; Greenfield et al., 2010; Fergusson and Henderson
et al., 2013)

The Tabberabberan Orogeny as defined in the Lachlan Orogen is constrained at ca. 380 Ma (e.g.,
VandenBerg et al., 2000; Gray and Foster, 2004; Glen, 2005). Deformations around this age also
occur in north Queensland, and are best represented in the Broken River Province, Charters Towers
region and possibly Hodgkinson Province, where they include time breaks and slight angular
unconformities (Withnall and Lang, 1993; Figure 2.12). Henderson (1987) suggested that deformation
resulted in the cessation of deep-marine sedimentation in the Camel Creek Subprovince, in the Middle
Devonian, caused the angular unconformity observed in the Graveyard Creek Subprovince. Garrad
and Bultitude (1999) record east to north-east thrusting and north-northwest-trending shear zones in
the Hodgkinson Province, which they suggested are Late Devonian in age, possibly related to basin
inversion. Although of a slightly younger age, this event may equate with the Tabberabberan Orogeny
in the Lachlan Orogen. Given the strong commonalities between the Broken River and Hodgkinson
provinces, it is probable that deep-water marine sedimentation ceased simultaneously in both regions.
Arnold and Fawckner (1980) suggested syn-deformational mélange formation in the Hodgkinson
Province occurred through the Tabberabberan Cycle. Additional complications and complexities are
evident in recent dating of the Mount Formartine Granite in the eastern Hodgkinson Province (just
north of Cairns). Kositcin et al. (2015a, b) report ca. 375 Ma ages for these granites, which have been
interpreted as syn-tectonic, e.g., Henderson et al. (2013), suggesting these rocks and the enclosing
Hodgkinson Formation, at least in that area, are recording the Tabberabberan Orogeny. Complications
to this interpretation have recently emerged as younger Late Devonian–early Carboniferous detrital
zircon ages have been found in other parts of the Hodgkinson Formation (Adams et al., 2013; Cross et
al., 2015b; Kositcin and Bultitude, 2015).

There are also older deformation events, at ca. 410–400 Ma, recorded in the Georgetown, Coen and
Charters Towers regions (Figure 2.12). The Coen and Charters Towers deformation coincides with
Pama Province magmatism in those regions. Bultitude et al. (1997) record a change in sedimentation
in the Hodgkinson Province in the late Lochkovian. Similarly, Withnall et al. (1997b) record a hiatus in
sedimentation at this time in the Graveyard Creek Subprovince. Both suggested these changes were
related to hinterland uplift. Notably, sedimentation appears to recommence at this time in the Charters
Towers region. These ca. 410–400 Ma events may correspond to the Tabberabberan Orogeny,
though more likely relate to the Bindian Orogeny.

2.2.5.2 Geological history and Time-Space plot explanation (Figure 2.10 to Figure 2.13)

Georgetown (and Greenvale) region

• Isolated pockets of, largely Lower Devonian, marine, and locally terrestrial, sediments and
limestone, e.g., Conjuboy Formation (Withnall et al., 1997b).
• East-west shortening and greenschist metamorphism (Withnall et al., 1997a), tentatively dated at
ca. 400 Ma, but may form part of the 430 Ma Benambran event. This age does, however, coincide
with lode Au mineralisation (410–400 Ma; Withnall et al., 1997a), and with extensive magmatism to
the north in the Coen region.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 51


Barnard region

• Regional deformation and associated low to potentially high-grade metamorphism (Bultitude et al.,
1997). This poorly constrained deformation event is bracketed between ca. 460 Ma and Permian
(Garrad and Bultitude, 1999).

Coen region

• Voluminous, S-type-dominated, mostly felsic, magmatism of the Pama Province. Dominated by the
S-type Kintore Supersuite, but includes subordinate I-type magmatism (see Bultitude et al. (pages
280–295) in Henderson et al., 2013). Ages constrained between ca. 410 Ma and 395 Ma (Blewett
et al., 1997).
• East-west shortening deformation and low-P high-T metamorphism (to upper amphibolite-facies)
with some contemporaneous Pama Province magmatism – ca. 407 Ma (Blewett et al., 1997).
• East-west shortening and thrusting, which Blewett et al. (1997) equated with inversion of the
Hodgkinson Province.

Charters Towers region

• Largely mixed siliciclastic marine sediments and carbonate successions of the Wilkie Gray and
Fanning River groups (Hutton et al., 1997). This sedimentation essentially continues through to the
Early Carboniferous. There is a reported unconformable contact and a significant change in
sedimentation above the Fanning River Group (e.g., Hutton et al., 1997), although Henderson (in
Fergusson and Henderson et al., 2013), have recently questioned the validity of this interpretation.
• Extensive magmatic rocks of the Pama Igneous Association. In the Charters Towers region, these
include the widespread I-type Millchester Supersuite in the Ravenswood Batholith as well as I-
types in the Reedy Springs Batholith. Ages are best constrained in the Ravenswood Batholith,
where they range from ca. 425 Ma to ca. 405 Ma (Hutton et al., 1997). Similar ages appear to
occur in the Reedy Springs and Lolworth batholiths, although the latter also include widespread
younger (ca. 380 Ma) granites which include S-types (Hutton et al., 1997; Bultitude et al. (pages
280–196) in Henderson et al., 2013)).
• Northeast to east-northeast faulting, with associated lode gold vein formation, ca. 400 Ma (e.g.,
Hutton et al., 1997).

Hodgkinson region

• Early Silurian to Early Devonian limestone, tholeiitic basalt and siliciclastic marine turbiditic
sediments of the Chillagoe Formation (Arnold and Fawckner, 1980; Garrad and Bultitude, 1999).
• Extensive marine, largely siliciclastic, turbiditic sediments––largely in the Hodgkinson Formation
(Arnold and Fawckner, 1980; Bultitude et al., 1997; Garrad and Bultitude, 1999). The latter also
includes minor tholeiitic metabasalts (Arnold and Fawckner, 1980). Upper and lower age limits are
not well constrained but are at least Early to Late Devonian in age (Garrad and Bultitude, 1999).
The latter is supported by Kositcin et al. (2015b) who record ca. 375–378 Ma ages for syn-tectonic
granites of the Mount Formartine Supersuite which intrude the Hodgkinson Formation in the
eastern part of the province. Recent detrital zircon work by Adams et al. (2013), Kositcin and
Bultitude (2015) and Cross et al. (in prep) however, suggests sedimentation, at least in the central
part of the formation, may extend into the earliest Carboniferous, with detrital ages of ca. 360 Ma.
• No unequivocal deformation of Tabberabberan age appears to have affected the Hodgkinson
Province as sedimentation continued throughout the Late Silurian and Devonian, although Arnold
and Fawckner (1980) suggested syn-deformational mélange formation. Also, Garrad and Bultitude
(1999) record east to north-east thrusting and north-northwest-trending shear zones in the

52 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Hodgkinson Province, which they suggested were of Late Devonian age, possibly related to basin
inversion, and the Tabberabberan Orogeny.

Camel Creek and Graveyard Creek regions

• In the Camel Creek Subprovince, continuation of marine, quartzose, turbiditic sedimentation,


locally with tholeiitic metabasalt and limestone, possibly to as young as Early Devonian in the
Kangaroo Hills Formation (Arnold and Fawckner, 1980; Withnall and Lang, 1993).
• Largely mixed siliciclastic marine (to locally fluviatile) sediments and carbonate successions
(Withnall and Lang, 1993; Withnall et al., 1997b) in the Graveyard Creek Subprovince. This
sedimentation appears to have continued through the Silurian to the Early Devonian, although
Withnall and Lang (1993) record time breaks between the Graveyard Creek Group and Shield
Creek Formation and the latter and the Broken River Group (Lochkovian and early Emsian).
• Southeast-northwest deformation, and accompanying greenschist to amphibolite-facies
metamorphism, is recorded in the Graveyard Creek Subprovince (e.g., Arnold and Fawckner,
1980). This deformation produced mild angular unconformities, and is constrained to be older than
Early Frasnian (Withnall et al., 1997b). The Camel Creek Subprovince records two events—east-
directed thrusting and folding, and southeast-northwest deformation with accompanying
greenschist-facies metamorphism (Arnold and Fawckner, 1980; Withnall and Lang, 1993). As
suggested by Withnall et al. (1997b), ages for these events are constrained between early
Devonian and early Carboniferous, but both deformations may equate with that recorded in the
Graveyard Creek Subprovince. Arnold and Fawckner (1980) suggested that this deformation was
late Devonian in age. Henderson (1987) suggested that that the southeast-northwest deformation
affected the two subprovinces, shutting off deep marine sedimentation in the Camel Creek
Subprovince, and producing the angular unconformity in the Graveyard Creek Subprovince.

2.2.6 Late Devonian to early Carboniferous (ca. 380 to ca. 350 Ma)

2.2.6.1 Geological and tectonic summary

This time period in north Queensland produced largely non-volcanic (cratonic provenance), terrestrial
and lesser marine sedimentation across all regions, best preserved in the lower successions of the
Bundock, Clarke River and Burdekin basins of the Broken River and Charters Towers regions
(Figure 2.12, Figure 2.13). Henderson et al. (1998), based on their correlation with the Drummond
Basin to the south, suggested that these three basins formed in response to back-arc processes,
behind the New England Orogen to the east. Sedimentation, with minor andesitic volcanism, is also
recorded in the Georgetown and Coen regions (Withnall et al., 1997a), and minor volcaniclastic input
is recorded in several of the regions during the Late Devonian. Subsequent deformation is minor in
nearly all areas, with the exception of the Hodgkinson Province where significant east-west shortening
and further basin inversion occurred (Garrad and Bultitude, 1997). This deformation and time period
immediately pre-dates the commencement of the voluminous and very widespread extrusive and
intrusive magmatism of the Kennedy Igneous Association (previously Kennedy Province; Champion
and Bultitude, 2013a, b).

The tectonic environment during the Late Devonian to early Carboniferous in northern Queensland is
poorly constrained. The possible continuation into this time period of the largely turbiditic
sedimentation in the Hodgkinson Province (Adams et al., 2013) suggests the persistence of a similar
geodynamic regime, i.e., deposition in either a backarc or forearc environment (e.g., Arnold and
Fawckner, 1980; Garrad and Bultitude, 1999; Henderson et al., 2013).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 53


2.2.6.2 Geological history and Time-Space plot explanation (Figure 2.10 to Figure 2.13)

Georgetown region (and Greenvale Province)

• Sporadic belts and blocks of terrestrial and marine sediments, e.g., in the Gilberton Formation
(Gilberton Basin; Withnall et al., 1997a; Withnall and Lang, 1993). Rocks are mostly non-volcanic
but do include locally abundant andesite lavas (Withnall et al., 1997a).
• Poorly constrained, very open, north-south deformation, thought to be mid Carboniferous, pre ca.
335 Ma (Withnall et al., 1997a).

Barnard Province

• Regional deformation and associated low to potentially high-grade metamorphism (Bultitude et al.,
1997). This poorly constrained deformation event (or events) is bracketed between ca. 460 Ma and
Permian (Garrad and Bultitude, 1999).

Coen Region
2
• Widespread terrestrial sediments (>1000 km ), including coal, in the Pascoe River Beds (Blewett et
al., 1997; McConachie et al, 1997). Volcanic detritus has been recorded in the succession, and it
appears to include a lower volcanic unit of uncertain age (McConachie et al, 1997).
• Poorly constrained minor deformation, thought to be Carboniferous or Permian in age (Blewett et
al., 1997). McConachie et al. (1997) indicate that the Pascoe River Beds are deformed, suggesting
that deformation was bracketed by the early Carboniferous age of the units and the ca. 285 Ma
age of the granite intrusions.

Charters Towers Region

• Terrestrial and marine sedimentation, including volcaniclastics in the Dotswood and Keelbottom
groups of the Burdekin Basin, with a number of transgressive-regressive cycles in the upper part
(early Carboniferous), e.g., Hutton et al. (1997).
• No apparent deformation recorded.

Hodgkinson Province

• Localised(?) continuation of the extensive marine, largely siliciclastic, turbiditic sediments of the
Hodgkinson Formation (Arnold and Fawckner, 1980; Bultitude et al., 1997; Garrad and Bultitude,
1999). As noted previously, there is evidence that sedimentation, at least in the central part of the
Province, may extend into the early Carboniferous, where detrital ages of 360 Ma have been
reported (Adams et al., 2013). Sedimentation may have been localised given that these
Carboniferous ages were not found elsewhere in the province by Adams et al. (2013). This
appears to be supported by recent 375–378 Ma ages for granites of the Mount Formartine
Supersuite (Kositcin et al., 2015a, b) in the eastern part of the province. These new ages are older
than the 357 Ma age reported for the same unit by Zucchetto et al. (1999).
• Locally distributed marine and terrestrial to marginal marine sediments, e.g., the Quadroy
Conglomerate, thought to be mostly early Carboniferous in age (Garrad and Bultitude, 1999).
• Significant early to mid-Carboniferous deformation (prior to late Carboniferous magmatism), which
according to Davis (1994) and Garrad and Bultitude (1999) produced significant east-west
shortening. Bultitude et al. (1997) included north-south shortening during this period. Garrad and
Bultitude (1999) suggested the latter was Permian in age. The north-south orientation, however,
suggests the deformation is likely Carboniferous in age, equating to the Alice Springs Orogeny.

54 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Broken River Province

• In both subprovinces, there are well developed successions of terrestrial and lesser marine
sedimentation (Bundock and Clarke River basins). These are mostly of cratonic provenance,
though do have some volcaniclastic input (Withnall and Lang, 1993, Withnall et al., 1993). This
style of sedimentation changed in the Visean with the commencement of widespread felsic-
dominated volcanism.
• No deformation recorded but there are apparent (mid-Tournaisian to early(?) Visean) time breaks
which coincide with a significant change in sedimentation style, namely the onset of regional
Kennedy Igneous Association magmatism (intrusive and extrusive) during the Visean, evident in
both the Bundock and Clarke River basins.

2.2.7 Middle Carboniferous to earliest Triassic (ca. 350 Ma to 250 Ma)

2.2.7.1 Geological and tectonic summary

This period in north Queensland is characterised by the commencement of the widespread and
voluminous extrusive and intrusive magmatism of the Kennedy Igneous Association (Bain and Draper,
1997; Champion and Bultitude, 2013a; previously Kennedy Province), plus associated, mostly minor
sedimentation (Figure 2.12, Figure 2.13). As documented by numerous authors (e.g., Richards et al.,
1966) this magmatism is crudely diachronous, commencing earlier in the Georgetown, Broken River
and Charters Towers regions (Figure 2.12, Figure 2.13), in the Visean (ca. 340–335 Ma), and
younging to mid-Permian in the Hodgkinson Province. There are also accompanying changes in
geochemistry. Magmatism in the mid to Late Carboniferous is almost exclusively I-type in nature, with
some of it mantle-derived. In the Early Permian, magmatism switched to A- and I-type in the
Georgetown and western Hodgkinson regions, and to S- and I-type in the central and eastern
Hodgkinson (Champion and Bultitude, 2003, 2013a, b).

There are at least two and probably three, deformations recorded during this time period. There is a
widespread but minor north-south contraction found in the Broken River, Hodgkinson and Georgetown
regions (Withnall et al., 1997a, b; Bultitude et al., 1997; Figure 2.12). This is thought to be mid to late
Carboniferous in age, and commonly equated to the Alice Springs Orogeny. In addition, there is a
more significant deformation event that occurred in the early Permian. This penetrative east-west
shortening deformation is best documented in the Hodgkinson Province (e.g., Davis, 1994; Davis et
al., 1996; Garrad and Bultitude, 1999; Figure 2.12). A late Permian east-west deformation event is
also recorded in the Hodgkinson and Barnard provinces (Garrad and Bultitude, 1999). The latter
deformation is equated with the Hunter-Bowen Orogeny.

The tectonic regime for this time period is not well understood. Magmatism of the Kennedy Igneous
Association, although dominated by crustal input, is generally thought to be broadly arc-related,
possibly back-arc (e.g., Champion and Bultitude, 2003, 2013a). This is consistent with the New
England Orogen, which at this time, was characterised by a convergent margin environment,
alternating between extension and contraction (e.g., Van Noord, 1999). This link is supported by
similar magmatic ages recorded in both the Kennedy Igneous Association and the northern New
England Orogen (compare Champion and Bultitude (2013a) and Bultitude et al. (pages 360–369) in
Purdy et al. (2013)).

2.2.7.2 Geological history and Time-Space plot explanation (Figure 2.10 to Figure 2.13)

Georgetown (and Greenvale) region

Geodynamic Synthesis of the Phanerozoic of eastern Australia 55


• Mid-Carboniferous to Early Permian voluminous extrusive and intrusive magmatism of the Kidston
and Tate subprovinces of the Kennedy Igneous Association. Magmatism ranges in age from ca.
340 Ma to ca. 275 Ma (e.g., Withnall et al., 1997a; see Champion and Bultitude, 2013a). It is
mostly felsic, but includes minor mafic and intermediate products (Champion and Chappell, 1992).
Magmatism is dominantly I-type but includes widespread but generally small volume A-types (e.g.,
Withnall et al., 1997a, Champion and Bultitude, 2003). The I-type magmatism shows a broad
decrease in age from ca. 340 Ma in the southwest to ca. 280 Ma in the northeast, e.g., Richards et
al. (1966), Champion and Bultitude (2003). The A-types, however, are typically younger, with ca.
290–275 Ma ages (Withnall et al., 1997a; Champion and Bultitude, 2003), and appear to occur
across the region. Magmatism of the Kennedy Igneous Association is extensively mineralised,
mostly associated with I-type granites (Champion and Heinemann, 1994).
• Minor, dominantly terrestrial, sedimentation occurs associated with volcanic complexes.
• A poorly defined, east-west deformation event, thought to be late Permian in age and equated with
the Hunter-Bowen Orogeny, has been documented (Withnall et al., 1997a).

Barnard Province

• Permian S-type, and less common I-type, Kennedy Igneous Association magmatism (Bultitude et
al., 1997; Garrad and Bultitude, 1999; Champion and Bultitude, 2013a, b).
• Regional deformation and associated low (to high?) grade metamorphism (Bultitude et al., 1997)
related to the Hunter-Bowen Orogeny. This event strongly deforms the early Permian granites and
so must postdate ca. 280 Ma (Garrad and Bultitude, 1999).

Coen Region

• Late Carboniferous to early Permian extrusive and intrusive magmatism of the Jardine
Subprovince, Kennedy Igneous Association, which contain (intermediate to) felsic I- and A-types.
Ages range from ca. 310 Ma (in the Torres Straits; Von Gnielinski et al., 1997) to Middle Permian
(ca. 275 Ma, Blewett et al., 1997).
• Permian terrestrial sediments, including coal, in the concealed Olive River Basin (McConachie et
al., 1997).
• Poorly constrained Carboniferous to Permian(?) deformation and low grade metamorphism
(Blewett et al., 1997). Both east-west and younger north-south deformations occur. The latter is
most consistent with events equated to the Alice Springs Orogeny, and as such is probably early to
mid Carboniferous in age. Younger events may be present, certainly McConachie et al. (1997)
record unconformities above and below the Olive River Basin.

Charters Towers region

• Terrestrial sedimentation and basaltic to felsic volcanics of the Early to mid-Carboniferous


Glenrock Group (Hutton et al., 1997). Of similar age (ca. 340–330 Ma) are rhyolitic volcanism and
felsic granites of the Oweenee Supersuite (Hutton et al., 1997), as well as other unrelated granites.
All these, and subsequent magmatism, belong to the Kennedy Igneous Association (Champion
and Bultitude, 2013a).
• Late Carboniferous to early Permian felsic volcanism and terrestrial sediments. These appear to
unconformably overlie the Glenrock Group and other similar aged volcanics.
• Late Carboniferous to mid-Permian mainly felsic I-type granites and associated (minor) mafic
intrusives. Hutton et al. (1997) indicated magmatism broadly youngs in age from ca. 310 Ma in the
west to as young as ca. 265 Ma in the east. The younger magmatism has transitional A-type
characteristics.

56 Geodynamic Synthesis of the Phanerozoic of eastern Australia


• Hutton et al. (1997) suggested a relatively stable backarc setting for magmatism of the Kennedy
Igneous Association in this region with a dominant extensional environment throughout the cycle.
The potential unconformity between the early to middle Carboniferous volcanics and overlying
rocks may indicate effects related to the Alice Springs Orogeny.

Hodgkinson Province

• Extensive I- and A-type extrusive and intrusive magmatism of the Herberton Subprovince,
Kennedy Igneous Association, and minor terrestrial sediments, concentrated in the western and
southern part of the Hodgkinson Province (Champion and Bultitude, 2003; 2013a). The I-type
magmatism has the greatest age range––ca. 320 Ma to ca. 275 Ma (Garrad and Bultitude, 1999).
The A-type magmatism is Permian in age (ca. 290 Ma to 275 Ma; Garrad and Bultitude, 1999).
The I-type intrusives are extensively mineralised.
• Permian S-type, and less common I-type magmatism of the Daintree Subprovince, Kennedy
Igneous Association, in the eastern half of the Hodgkinson Province (Bultitude and Champion,
1992; Champion and Bultitude, 2003, 2013a, b). The magmatism appears to decrease in age
towards the northeast and east, from ca. 280 Ma to ca. 260 Ma, and possibly younger (Richards et
al., 1966; Garrad and Bultitude, 1999).
• Locally distributed, largely terrestrial sediments and coal measures of Late Permian age, e.g.,
Mount Mulligan Coal Measures (Garrad and Bultitude, 1999).
• Regional (syn-granite emplacement) deformation (at ca. 275–280 Ma), best documented by Davis
(1994). According to Davis (1994) this deformation, which produced north-south fabrics, was
strongly partitioned and so is variably developed across the Province. This age corresponds
closely to that of A-type magmatism in the western Hodgkinson Province and Georgetown region,
which is commonly interpreted as being emplaced in an extensional environment.
• Significant Late Permian to Early Triassic transpressive(?) deformation and greenschist
metamorphism thought to be related to the Hunter-Bowen Orogeny (Garrad and Bultitude, 1999).
Bultitude et al. (1997) document K–Ar deformation ages of ca. 250 Ma in older granites.
• Minor Triassic terrestrial sediments, which, at least locally, overlie older Permian rocks with an
angular unconformity (Garrad and Bultitude, 1999).

Broken River Province

• Visean and younger terrestrial sedimentation and felsic volcanics in the upper parts of the Bundock
and Clarke River groups (Withnall and Lang, 1993; Withnall et al., 1997b). These rocks equate
with the Glenrock Group in the Charters Towers region, though lack the mafic-intermediate rocks
found in the latter.
• The margins of the province shares similar Kennedy Igneous Association magmatism to that
documented for the Hodgkinson and Charters Towers regions (Withnall et al., 1997b).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 57


2.3 New England Orogen
by DC Champion, updating N Kositcin (2009)

2.3.1 Introduction
The New England Orogen (NEO, Figure 2.14, Figure 2.15) is a complex collage of different terranes
(i.e., tectonic elements) on the eastern margin of the Australian continent (Cawood and Leitch, 1985;
Leitch and Scheibner, 1987; Flood and Aitchison, 1988). The NEO has had a complicated evolutionary
history that stretches from the Neoproterozoic to the Late Mesozoic, although most of the terranes are
Silurian to Carboniferous in age (Aitchison et al., 1992a; Aitchison and Ireland, 1995). Terrane
nomenclature for the New England Orogen is largely split along state lines. Donchak et al. (2013)
provide a recent synthesis of the New England Orogen in Queensland where a formalised province
and subprovince nomenclature has been introduced replacing the previous block, terrane and belt
nomenclature. The latter is still used for the New South Wales portion of the orogen, essentially
following the nomenclature used by Korsch et al. (2009c; Figure 2.14).

The major component of the New England Orogen evolved during the Devonian and Carboniferous in
a convergent plate margin tectonic setting related to a west-dipping subduction system (Murray et al.,
1987; Korsch et al., 1990; Fergusson et al., 1993). Parts of the volcanic arc, forearc basin and
accretionary wedge are still preserved in the orogen (Figure 2.15). In the northern NEO, Devonian and
Carboniferous rocks can be subdivided into arc, forearc, and accretionary wedge assemblages,
consisting of the Connors and Auburn Subprovinces in the west (magmatic arc), the Rockhampton
Subprovince (previously Yarrol Belt) in the centre (forearc basin), the accretionary wedge to the east
(Coastal, Yarraman, North D’Aguilar, South D’Aguilar and Beenleigh subprovinces; Murray, 1986;
Figure 2.14, Figure 2.15), and a backarc basin (Drummond Basin). In the southern NEO, the arc is
inferred to have been located to the west of the Tamworth Belt (either concealed beneath the
Gunnedah Basin or Tamworth Belt, or removed by erosion and/or strike-slip faulting; Korsch et al.,
1997). The forearc basin is represented by the Tamworth Belt and Hastings Block, and the
accretionary wedge by the Tablelands Complex (Woolomin, Sandon and Coffs Harbour Associations
of Korsch, 1977, or in this compilation, the Woolomin Province, and the Wisemans Arm, Central and
Coffs Harbour terranes; Figure 2.14). The present boundary between forearc basin strata to the west
and subduction complex assemblages to the east is a major fault zone marked by serpentinite lenses
(the Yarrol Fault in the northern NEO and the Peel-Manning Fault System in the southern NEO;
Murray, 1988; Figure 2.14). The Gympie Province, in the northern NEO, is thought to be an exotic
terrane (e.g., Korsch and Harrington 1987; Murray, 1988).

Numerous tectonic models have been proposed for the development of the NEO (e.g., Leitch, 1975;
Cawood 1982, 1983; Murray et al., 1987). Most infer a long-lived, east-facing convergent plate margin
setting, with progressive accretion of younger rocks at the eastern margin of Gondwanaland. The
orogen was most likely island arc-related from the Cambrian to the Early Devonian, evolving to a
continental margin magmatic arc from the Late Devonian onwards. The subsequent history of the
orogen involved strike-slip faulting and major oroclinal bending. Large amounts of volcanism and
plutonism took place in the Late Permian and Early Triassic (e.g., Korsch et al., 1990).

A brief synthesis, highlighting key points in the evolution of the NEO from the Neoproterozoic through
to the Triassic, is presented in the following sections and accompanying time-space and other plots.
The time-space plots (Figure 2.16 and Figure 2.17) summarise our current state of knowledge of the
stratigraphic history and tectonic evolution of the NEO through time.

58 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 2.14 Distribution of geological provinces, subprovinces and blocks in the New England Orogen used to
construct the time-space plots. Nomenclature for the Queensland part of the orogen has been updated from that
of Kositcin in Champion et al. (2009) to the new definitions of Donchak et al. (2013); nomenclature for the
southern part of the orogen follows Korsch et al. (2009c). WP = Woolomin Province.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 59


Figure 2.15 Broad subdivision of the New England Orogen into regions defined by tectonic setting.

2.3.2 Late Neoproterozoic to latest Cambrian (ca. 600 to ca. 490 Ma)

2.3.2.1 Geological and tectonic summary

The New England Orogen in this period is dominated by Cambrian tectonic blocks, largely of oceanic
fragments, including island arc-related remnants (Figure 2.18). These occur in the southern NEO, as
accreted blocks along the Peel-Manning Fault System (PMFS, e.g., Offler and Shaw, 2006;
Figure 2.14). These blocks are generally thought to provide records of subduction environments
offshore of continental Australia in the late Neoproterozoic and Cambrian, and to have been accreted
to the NEO in the Late Paleozoic, though Glen (2013) has recently suggested an alternative model
that invokes a long lived arc in the NEO that operated from the Cambrian to the Mesozoic.

60 Geodynamic Synthesis of the Phanerozoic of eastern Australia


The only definitive record of Neoproterozoic-Cambrian material in the northern NEO is represented by
the Princhester and related ophiolites, which are ca. 565 Ma in age (Bruce et al., 2000; Figure 2.18).
These were most probably accreted to the NEO in the Late Paleozoic (Murray and Blake, 2005).

2.3.2.2 Geological history and Time-Space plot explanation

Southern NEO (Figure 2.14 to Figure 2.18)

• In the southern New England Orogen, the Peel-Manning Fault System (PMFS; Figure 2.14) is a
major structural element characterised by tectonic blocks in serpentinite mélange (Offler and
Shaw, 2006).
• The PMFS contains fragments of dismembered ophiolite of suprasubduction origin (Aitchison et
al., 1994), the oldest of which is Cambrian in age. Eclogite blocks (Attunga Eclogite), exhumed
along the PMFS were originally dated at 571 Ma by Watanabe et al. (1998); this date has now
been revised to ca. 536 Ma (Fanning et al., 2002). Island arc boninitic volcanism at 536 Ma
associated with subduction and high P-low T metamorphism of MORB-like basalts to eclogite,
therefore, commenced in the Early Cambrian. This record of Cambrian subduction is earlier than
boninitic magmatism recorded in Tasmania (514 ± 5 Ma; Black et al., 1997) and Victoria (519–514
Ma; VandenBerg et al., 2000).
• Similar ages to the Attunga Eclogite have been obtained from plagiogranite (U-Pb zircon 530 ± 6
Ma; Aitchison et al., 1992a) and metadiorite (Sm-Nd isochron 536 ± 22 Ma; Sano et al., 2004) in
schistose serpentinite along the same fault system. The plagiogranite age suggests that the
enclosing ophiolitic low-Ti tholeiitic basalts and boninitic ultramafic rocks are relics of a Cambrian
suprasubduction zone forearc and, thus, of a Cambrian convergent plate boundary (Aitchison et
al., 1994). Chemically, the basalts resemble Cambrian basalts of western Tasmania (Aitchison and
Ireland, 1995). Sano et al. (2004) suggest that the boninitic metadiorite formed in an immature
island arc setting as a result of mixing of a depleted mantle wedge with sediment melt and fluid
and melt derived from MORB.
• The Rocky Beach Metamorphic Mélange (ca. 545 Ma) and the Port Macquarie Serpentinite (~545–
509 Ma) are inferred to be the oldest rocks in the Port Macquarie Block (Figure 2.17, Figure 2.18),
largely dated by analogy with rocks from elsewhere in the NEO (e.g., Och et al., 2007). They
contain a fragmentary late Neoproterozoic to early Paleozoic history that includes subduction and
associated metamorphism under high-P low T conditions, possibly around 536 Ma (Watanabe et
al., 1998; Och et al., 2007). The Port Macquarie Serpentinite is a product of alteration of cumulate
ultramafic rocks of a ca. 530 Ma forearc ophiolite (Och et al., 2007).
• Palaeontological studies have revealed island arc-related activity during the middle to later part of the
Cambrian in the Gamilaroi Terrane (Cawood, 1976; Leitch and Cawood, 1987; Stewart, 1995).
Volcaniclastic rocks from the Murrawong Creek and Pipeclay Creek Formations contain Cambrian
trilobite, brachiopod and conodont faunas (Cawood, 1976; Cawood and Leitch 1985; Stewart, 1995),
and possibly form an ancient basement to the predominantly mid-Paleozoic rocks of the Gamilaroi
Terrane (Stratford and Aitchison, 1997). Cambrian conodonts in the Pipeclay Creek Formation
(Stewart, 1995) indicate the presence of Cambrian sedimentary rocks below the main Devonian-
Carboniferous pile in the Tamworth Belt, supporting the idea of Korsch et al. (1997) that perhaps at
least some of the ophiolites were the floor to the Tamworth Belt. Cawood and Leitch (1985) have
suggested that volcanic clasts in these sediments were derived from a low-K intra-oceanic island arc.

Northern NEO (Figure 2.14 to Figure 2.18)

• In the northern NEO, the rift phase of the Delamerian cycle is represented by an ophiolite
(Northumberland and Princhester serpentinites) in the Marlborough Province of central

Geodynamic Synthesis of the Phanerozoic of eastern Australia 61


Queensland that has a 562 ± 22 Ma Sm–Nd isochron age (Bruce et al., 2000). The ophiolite has
depleted MORB-like trace element characteristics that suggest formation as oceanic crust at a
Neoproterozoic ocean spreading ridge (Bruce et al., 2000). These data are consistent with the
existence of a proto-Pacific Ocean east of the Delamerian Orogen after supercontinent breakup
and is consistent with the inferred presence of old lithosphere under the southern NEO (Glen,
2005). The Marlborough Terrane was emplaced into its present position during the Hunter-Bowen
Orogeny (Holcombe et al., 1997b; Korsch et al., 1997; Murray and Blake, 2005).
• There appears to be no record in the northern NEO of supra-subduction zone ophiolites and
volcanic arc deposits of Cambrian age, similar to those exposed along the PMFS. Age and
geochemical data suggest that the ~562 Ma magmatism in the Princhester Serpentinite must have
occurred outboard of any Early Paleozoic subduction zone along the margin of the Australian
continent (Murray, 2007).

2.3.3 Latest Cambrian to early Silurian (ca. 490–430 Ma)

2.3.3.1 Geological and tectonic summary

The Benambran cycle geology in the New England Orogen is dominated by Ordovician and Silurian
sedimentary rocks of island arc and oceanic affinity and Late Ordovician arc magmatism, the latter
recorded along the PMFS (e.g., Offler and Shaw, 2006; Figure 2.18). Early and Middle Ordovician
blueschist metamorphism resulted from the exhumation of Cambrian to Early Ordovician rocks at Port
Macquarie and along the PMFS.

2.3.3.2 Geological history and Time-Space plot explanation

• Following island arc boninitic volcanism at 536 Ma, exhumation of MORB-like basalts in the Early
Ordovician gave rise to: high-pressure blueschist metamorphism (K/Ar ages of 482–467 Ma for
blueschist rocks within serpentinite-matrix mélange at Glenrock and Pigna Barney; Fukui et al.,
1995); arc-related plutonism, e.g., the Attunga gabbro (U-Pb zircon; 479 ± 11 Ma; Fanning et al.,
2002); and accretionary wedge-related blueschist metamorphism (Fukui et al., 1995; Offler, 1999),
along the PMFS (Offler, 1999). Offler and Shaw (2006) also provide evidence for Late Ordovician
arc magmatism along the PMFS (U-Pb zircon age of 444.7 ± 2.4 Ma for hornblende gabbro of
calcalkaline affinity at Glenrock station; see also Watanabe et al., 1998). Similarly, a U-Pb zircon
age of 436 ± 9 Ma for tonalite of the Pola Fogal Suite in the Pigna Barney Ophiolite Complex
(Kimbrough et al., 1993) may also be related to an Ordovician arc, though may also reflect Pb-loss
from the Cambrian arc (Kimbrough et al., 1993).

• The oldest sedimentary rocks in the southern NEO are siliciclastic sedimentary rocks and
fossiliferous limestones found within the imbricate zone of the PMFS, and include the Trelawney
beds and Haedon Formation in the Gamilaroi Terrane (island arc related environment), and Uralba
beds (probably part of the Gamilaroi Terrane), preserved along the PMFS (Figure 2.18). Corals
and conodonts from these limestones indicate they are of Late Ordovician age (Hall, 1975). A
similar-aged limestone succession lies below the Silverwood Group (Wass and Dennis, 1977;
Silverwood Terrane; Figure 2.16, Figure 2.17). Late Ordovician coral limestone, probably
representing partly accreted seamounts, was also deposited in an ocean basin environment, now
incorporated in the Woolomin Terrane (435–428 Ma; Hall, 1978).

• Middle to Late Ordovician rocks (chert, mudstone, siltstone, tuffaceous sandstone, tuff,
conglomerate, olistostromal rocks and basalt) of the Watonga Formation at Port Macquarie
interpreted to have been deposited on oceanic crust prior to accretion (Och et al., 2007).

62 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Geodynamic Synthesis of the Phanerozoic of eastern Australia 63
64 Geodynamic Synthesis of the Phanerozoic of eastern Australia
Geodynamic Synthesis of the Phanerozoic of eastern Australia 65
66 Geodynamic Synthesis of the Phanerozoic of eastern Australia
Geodynamic Synthesis of the Phanerozoic of eastern Australia 67
Figure 2.16 Late Neoproterozoic to Jurassic time-space plot for part of the northern New England Orogen. Refer
to text for data sources and discussion. Terranes, provinces and regions are as outlined in Figure 2.14 and
Figure 2.15.

68 Geodynamic Synthesis of the Phanerozoic of eastern Australia


• K-Ar ages of ca. 469 Ma for phengite related to retrograde blueschist metamorphism of eclogite
blocks in the Port Macquarie Block has been interpreted to suggest that substantial exhumation of
these rocks had occurred by the Middle Ordovician (Fukui, 1991; Watanabe et al., 1993; Fukui et
al., 1995; Offler, 1999). In addition, glaucophane from blueschist in mélange of the Port Macquarie
Block has a K-Ar isotopic age of 444 Ma (Lanphere, cited in Scheibner, 1985).
• Blake (in Donchak et al., 2013) reports a small fault-bounded(?) block of limestone and
volcaniclastic sediments, close to the Yarrol Fault in the northern NEO. The rocks are thought to
be Ordovician in age, based on fossil assemblages. Blake suggested that these rocks may be
analogous to similar units along the Peel-Manning Fault System in the southern NEO.

2.3.4 Middle Silurian to late Devonian (ca. 430–380 Ma): intra-oceanic island
arc

2.3.4.1 Geological and tectonic summary

Late Silurian to late Devonian arc successions in the Gamilaroi Terrane, Calliope Province, Coffs
Harbour North (Willowie Creek Group), Silverwood (Silverwood Group), and at Alice Creek (Craigilee
Equivalents), are interpreted to have all formed as part of one intraoceanic island arc (e.g., van Noord,
1999; Offler and Murray, 2011; Figure 2.16 to Figure 2.18), though variations in polarity and number of
subduction zones have been suggested (e.g., Aitchinson and Flood, 1995; Murray, 2007; Offler and
Murray, 2011). Island arc magmatism is also recorded along the PMFS (early Silurian) and in the
Marlborough Province (middle Devonian). Ocean basin cherts and basalts of the Woolomin Terrane
and Coastal Block were deposited in this time interval.

2.3.4.2 Geological history and Time-Space plot explanation

Southern NEO (Figure 2.14 to Figure 2.18)

• The Gamilaroi Terrane is made up of Late Silurian–Devonian arc volcanics and volcanogenic
sediments (Flood and Aitchison, 1988, 1992; Aitchison and Flood, 1995; Stratford and Aitchison,
1995; Offler and Gamble, 2002; Offler and Murray, 2011). Diverse suites of volcaniclastic
sediments are intercalated with regionally extensive extrusive and subvolcanic felsic meta-igneous
rocks of low K and calcalkaline affinity (Cawood and Flood, 1989; Aitchison and Flood, 1995; Offler
and Gamble, 2002).
• Geochemical studies suggest that the Late Silurian–Early Devonian successions of the Gamilaroi
Terrane formed in an intraoceanic arc setting that was rifted in the Middle to Late Devonian (Offler
and Gamble, 2002; Offler and Murray, 2011).
• The sedimentary and volcanic rocks of the Gamilaroi Terrane are unconformably overlain to the
south and west by Late Devonian–Carboniferous age rocks of the Tamworth Belt (Aitchison and
Flood, 1992). The Gamilaroi Terrane had been accreted to the Gondwana margin by the Late
Devonian and is constrained by the first appearance of distinctive, westerly-derived (Lachlan
Orogen) quartzite clasts in the uppermost Devonian Keepit Conglomerate (Flood and Aitchison,
1992).
• Two rock successions of Late Silurian–Middle Devonian age, the Silverwood Group and Willowie
Creek beds, have been interpreted as either a southern continuation of the Calliope Province (Day
et al., 1978) or as displaced fragments of the Tamworth Group (Cawood and Leitch, 1985). We
support an island arc rather than a continental margin forearc basin setting for these successions.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 69


• Remnant blocks recording Early Silurian island arc magmatism and metamorphism occur along the
PMFS. This is recorded by a ca. 425 Ma age for hornblende cumulates and diorites from Glenrock
Station (Sano et al., 2004).
• Late Silurian to Devonian sedimentation recorded in the Woolomin Terrane consists of 428–380
Ma chert and basalt deposited in an ocean basin.

Northern NEO (Figure 2.14 to Figure 2.18)

• The Calliope Province (Blake in Donchak et al., 2013) comprises a number of recognised
subprovinces (now present as separate fault blocks and inliers), consisting of Late Silurian to
Middle Devonian shallow marine volcaniclastic sediments with varying amounts of calcalkaline
felsic to mafic volcanic rocks. Geochemical studies suggest formation in either a primitive
continental or island arc setting (Morand, 1993a; Offler and Gamble, 2002; Murray and Blake,
2005; Offler and Murray, 2011). Murray and Blake (2005) suggest that the data support an origin
for the Silurian to Middle Devonian assemblages as exotic oceanic terranes mainly from island
arcs, but may include some backarc basin rocks (see Bryan et al., 2004; Offler and Murray, 2011).
The Mount Morgan Tonalite intruded the succession at 381 ± 5 Ma, and has an island arc or rifted-
arc geochemistry (Golding et al., 1994; Murray, 2003). The Calliope assemblage hosts the Mount
Morgan gold-copper deposit (~380 Ma; Yarrol Project Team, 1997, 2003).
• The Calliope Province rocks are overlain by Late Devonian and younger forearc basin strata of the
Rockhampton Subprovince (Yarrol Province) (see below). If exotic, the Calliope Province must
have reached its present position by the end of the Middle Devonian; the accretion event being
represented by an early Late Devonian unconformity with the overlying Late Devonian
Rockhampton Subprovince (e.g., Korsch et al., 1990). Morand (1993b) suggested the unconformity
did not relate to major orogenesis, and suggested it was more consistent with a continental arc
origin for the Calliope Arc.
• Ocean basin cherts were deposited in the Coastal Subprovince (Wandilla Province) during this
period, with Doonside Formation sedimentation occurring between ca. 418 Ma and ca. 360 Ma.
• Calcalkaline basalt, dolerite, and gabbro in the Neoproterozoic Marlborough Ophiolite
(Marlborough Province) have a Middle Devonian Sm-Nd isochron age (380 ± 19 Ma) and trace
element data suggestive of an intra-oceanic island arc (Bruce and Niu, 2000). These relations
suggest that the arc was probably built on Neoproterozoic oceanic crust, that the arc lay only a
short distance offshore, and was accreted to Gondwana by the early Permian (Bruce and Niu,
2000; Murray, 2007).

2.3.5 Late Devonian to Late Triassic overview (380–220 Ma)


The accretion of the Gamilaroi-Calliope island arc to the Gondwana margin heralded the initiation of
the New England Orogen as an Andean-style continental margin. The subsequent history of the NEO
records an event history distinct from the remainder of the Tasmanides (e.g., Glen, 2005). This is not
to suggest that Kanimblan- or Alice Springs-related deformations are not present, just that they do not
appear to be significant in regards to the formation of the orogen. The main phases in the NEO are:

• Development of the NEO as a continental margin magmatic arc (Connors-Auburn Arc; Keepit-
Connors arc of Cawood et al., 2011) with associated fore-arc and accretionary wedge
development, from the late Middle Devonian to the late Carboniferous;
• An extensional, back-arc phase from the late Carboniferous to late Early Permian, possibly
transitional from the last phase. Extension was accompanied by widespread magmatism and

70 Geodynamic Synthesis of the Phanerozoic of eastern Australia


sedimentation, including initiation of the Sydney-Gunnedah-Bowen basin system. Extension was
terminated by deformation and formation of the Texas and Coffs Harbour Oroclines;
• Renewed continental margin arc magmatism and sedimentation, with associated contraction and
formation of a fold-thrust belt and retroforeland basin. This was related to the main phase of the
Hunter Bowen Orogeny, occurring during the early Permian to Middle Triassic period. This orogeny
effectively cratonised eastern Australia. It was followed by widespread, backarc(?) extensional-
related magmatism and sedimentation.

2.3.6 Late Devonian to late Carboniferous (ca. 380–305 Ma): continental arc

2.3.6.1 Geological and tectonic summary

During the Late Devonian and Carboniferous, eastern Australia was located on a convergent margin
with a westerly dipping subduction zone responsible for the development of a magmatic arc in the
west, flanked by a forearc basin and accretionary wedge in the east (Cawood, 1982; Murray et al.,
1987). Components of the convergent plate margin system in the southern NEO include the
accretionary wedge (Woolomin, Central and Coffs Harbour blocks), forearc basin (Tamworth Belt and
Hastings Block) and magmatic arc (now either buried or removed; Figure 2.15, Figure 2.18). In the
northern NEO (Donchak et al., 2013), a continental margin magmatic arc is represented by the
Connors and Auburn subprovinces (Yarrol Province), the forearc basin is represented by the
Rockhampton Subprovince (Yarrol Province), and the accretionary complex is represented by the
Coastal, Yarraman, North D’Aguilar, South D’Aguilar and Beenleigh subprovinces of the Wandilla
Province (Figure 2.15, Figure 2.18). A possible back-arc basin occurs in the north (Drummond Basin),
but not in the south (Glen, 2005), although the Late Devonian of the Lachlan Orogen is probably the
backarc equivalent.

2.3.6.2 Magmatic arc

There is general agreement that during the Late Devonian the tectonic setting of eastern Australia
changed as a result of the accretion of the intraoceanic Gamilaroi-Calliope island arc to the
Gondwanan continental margin (e.g., Aitchison et al., 1992b; although cf. Morand (1993a) for an
alternate view). This led to the development of an Andean-style continental arc––the Connors-Auburn
continental arc––that remained active until the latest Carboniferous (Roberts et al., 1995, see below).

• The Connors and Auburn subprovinces (Yarrol Province) comprise late Paleozoic granites and
silicic volcanic rocks (e.g., Withnall et al., pages 351–369, in Donchak et al., 2013; Figure 2.14;
Figure 2.16). These are generally considered to represent the magmatic portion of the Connors-
Auburn continental arc (Murray, 1986; Figure 2.15, Figure 2.17, Figure 2.18); though other
interpretations have been proposed, e.g., interpreted as back-arc magmatism by Bryan et al.
(2001). The Connors and Auburn subprovinces are separated by the Gogango Thrust Zone
(Figure 2.14) which represents part of the Permian Bowen Basin succession strongly deformed
and thrust westwards in the late Permian to Early Triassic (Fergusson, 1991). Igneous activity in
both subprovinces commenced at ca. 350 Ma, although the main pulse of granite formation was
between ca. 320 Ma and 300 Ma in the Auburn Subprovince and ca. 330-305 Ma in the Connors
Subprovince (Murray, 2003; Withnall et al., pages 351–369, in Donchak et al., 2013). Murray
(2003) suggested that the older granites in the Auburn Subprovince and the older part of the
Connors Subprovince are subduction-related, and the younger granites with large volumes of
rhyolitic to dacitic ignimbrites and local andesite lavas in the Connors Subprovince spanned the
changeover from subduction to the beginning of extension, at around 305 Ma (see below).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 71


2.3.6.3 Forearc basin

To the west of the Peel and Yarrol Faults, the Tamworth Belt in the southern NEO and the
Rockhampton Subprovince (Yarrol Province) in the northern NEO are usually interpreted to be parts of
a continuous forearc basin (Korsch et al., 1990; Figure 2.15, Figure 2.18). The forearc basin deepened
towards the east and consists predominantly of continental to shallow marine clastic sediments,
derived predominantly from the volcanic arc to the west.

Southern NEO

• Early and Middle Devonian volcaniclastic sediments, volcanics and limestones of the Tamworth
Belt (Figure 2.16) were deposited in a forearc basin to the west of the PMFS. Cawood and Leitch
(1985) interpreted the Tamworth Group as the fill of a forearc basin between the western arc and
the partly uplifted accretionary prism to the east of the PMFS. An arc along the inboard part of the
orogen is not preserved but is inferred from large Late Devonian olistromal blocks of andesitic
volcanic rocks in the inboard part of the forearc basin, i.e., the Tamworth Belt (Brown, 1987).
Indeed, strong evidence for its existence is provided by the composition of sandstones from the
Tamworth Belt and accretionary complex with virtually all of them classified as arc-derived
volcaniclastics (Cawood, 1983; Korsch, 1984), whereas volcanic material is more abundant closer
to the active magmatic arc (McPhie, 1987). The forearc Tamworth Belt was filled by marine strata
that become finer grained and of deeper water character eastwards towards the PMFS, which
marks the preserved outboard edge of the basin (Yarrol Project Team, 1997). While deposition in
the forearc basin was dominated by volcaniclastic sediments with minor interbedded volcanics,
limestones developed in the Early Carboniferous during periods of diminished terrigenous
sedimentation (Roberts and Engel, 1980; Cawood and Leitch, 1985). More detailed descriptions of
sedimentary rocks from the Tamworth Belt are given by Cawood (1983) and Korsch (1984).
• The Hastings Terrane is widely considered to represent a dispersed fragment of the Tamworth
forearc basin, with which it shows an overall similarity throughout its tectonic development
(Cawood and Leitch, 1985; Roberts et al., 1993). Units interpreted to part of the Tamworth Belt
form part of the Texas and Coffs Harbour Oroclines. Components of the Oroclines considered to
be part of the forearc basin succession include rocks of the Emu Creek Block and Carboniferous
outcrops at Mount Barney and Alice Creek (Figure 2.16 to Figure 2.18). We have also assigned a
forearc environment for deposition of the Touchwood Formation in the Port Macquarie Block.
• Volcanism and deposition in the Tamworth Belt appear to have largely ceased at ~305 Ma in
response to the eastward migration of the subduction zone (Roberts et al., 2004). Mechanisms to
account for this migration have included slab breakoff (Caprarelli and Leitch, 2001) and rollback
(Jenkins et al., 2002).

Northern NEO

• The Rockhampton Group in the Rockhampton Subprovince is considered to have been deposited
in a forearc basin (Murray et al., 2003; although cf. Bryan et al., 2001) and covered the former Late
Devonian arc, as well as older basin strata to the east (Figure 2.18). Late Devonian strata consist
of volcaniclastic sandstone and conglomerate derived from the andesitic arc to the west,
interbedded with sediments and some limestone (Yarrol Project Team, 1997).
• On a regional scale, the Late Devonian rocks of the Rockhampton Subprovince are considered to
represent a transitional phase in the change from an intraoceanic setting (Calliope Arc), epitomised
by the Middle Devonian Capella Creek Group, to a continental margin setting in the northern NEO
in the Carboniferous (Murray and Blake, 2005).

72 Geodynamic Synthesis of the Phanerozoic of eastern Australia


2.3.6.4 Accretionary wedge

The Woolomin, Wisemans Arm, Central and Coffs Harbour blocks in the southern NEO and the
Coastal, Yarraman, North D’Aguilar, South D’Aguilar and Beenleigh subprovinces in the northern NEO
are interpreted as a once continuous accretionary wedge that grew ocean-wards by accreting trench-
fill volcaniclastic turbidites (derived from a magmatic arc) and minor amounts of oceanic crust (basalt,
chert, mudstone) (Figure 2.15, Figure 2.18; Korsch et al., 1990).

Southern NEO

• The mid-Devonian to Carboniferous accretionary wedge in New South Wales is made up of the
Woolomin Province (chert, basalt – mostly accreted ocean floor, very rare volcaniclastic
sandstone, mudstone, turbidite – trench fill), the Coffs Harbour Block (mainly trench fill turbidite),
the southern Beenleigh Subprovince and successions at Wisemans Arm.
• The accretionary wedge consists largely of deep marine sedimentary rocks, including metabasalt-
chert-argillite associations (Cawood, 1982). Radiolaria indicate a mid-Silurian age for basalt, Late
Silurian-Frasnian age for chert and a Fammennian age for siliceous chert and overlying
volcanogenic sandstone of the westernmost Woolomin Province (Aitchison et al., 1992b). The
Central Block consists of deformed and dismembered volcanogenic siltstone and sandstone,
basalt, chert and minor conglomerate (Cross et al., 1987).

Northern NEO

• Mid-Devonian to Carboniferous accretionary wedge rocks in the northern NEO (Figure 2.18) occur
in the Wandilla Province and the Coastal Subprovince in the north, and in the North and South
D’Aguilar, Yarraman and Beenleigh subprovinces in the south. The Beenleigh Subprovince
consists of greywacke, argillite and early Carboniferous chert (Aitchison, 1988). Similar strata
occur in the South D’Aguilar Subprovince (Holcombe et al., 1997b). In contrast, the North
D’Aguilar Subprovince is a composite terrane, containing high-level accretionary wedge rocks as
well as ophiolitic components of an accretionary wedge that were subducted to depths of ~20 km
(~6 kbars) before 315 Ma, and were subsequently exhumed in the lower plate below a latest
Carboniferous, low angle normal fault (Little et al., 1992, 1995; Holcombe et al., 1997b). Similar-
aged accretionary wedge successions are recorded in the Yarraman Subprovince and
Marlborough Province.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 73


74 Geodynamic Synthesis of the Phanerozoic of eastern Australia
Figure 2.17 Time-space plot of the NEO by tectonic setting. Refer to text for data sources and discussion. Terranes and provinces are as outlined in Figure 2.14 and
Figure 2.15. Unit names are as outlined in Figure 2.16.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 75


2.3.7 Late Carboniferous to late Permian (ca. 305 to 265 Ma): back-arc
extension and orocline formation

2.3.7.1 Geological and tectonic summary

Along the Gondwana margin, a transition took place from active accretion in the mid-Carboniferous to
widespread extension through the latest Carboniferous and Early Permian (Leitch, 1988; Holcombe et
al., 1997a). This transition has been interpreted in terms of eastward retreat of the subducting slab,
and migration of the volcanic arc offshore (Holcombe et al., 1997a). Thus, by the Late Permian, much
of the NEO was the site of numerous backarc extensional basins filled with marine and terrestrial
deposits and mafic-silicic volcanics. Two recent models for the late Paleozoic evolution of the southern
NEO (Caprarelli and Leitch, 1998, 2001; Jenkins et al., 2002) involve alternate mechanisms, namely
slab breakoff, and arc rollback, respectively, to explain the easterly migration of the subduction zone,
generation of granitic suites, and large scale volcanism and deformation.

In the early Permian (~295–280 Ma), the initiation of the Bowen-Gunnedah-Sydney basin system
occurred to the west of the continental margin magmatic arc in a backarc tectonic setting (Korsch et
al., 1993; Figure 2.18). Also around this time, but probably after extension ceased, oroclinal bending of
the forearc and accretionary wedge successions (~285–265 Ma) produced the Texas and Coffs
Harbour Oroclines (Korsch and Harrington, 1987; Murray et al., 1987). Rosenbaum et al. (2012) have
suggested orocline formation may have occurred in a number of phases, with an earlier subduction-
related phase of curvature during slab rollback at ca. 300–285 Ma, and a later tightening related to the
Hunter-Bowen Orogeny.

2.3.7.2 Geological history and Time-Space plot explanation

NEO

• In the southern NEO, the cessation of subduction just before the end of the Carboniferous and a
postulated eastward migration of the subduction zone was followed by emplacement of early S-
type granites at ~300 Ma (e.g., Collins et al., 1993; Figure 2.18). Approximately contemporaneous
cessation of deposition in the Tamworth forearc basin at ca. 305 Ma was followed by major strike-
slip faulting and anticlockwise rotation of crustal blocks (Roberts et al., 1995, 1996). A change
from contraction to extension in the NEO led to the region being located in a backarc environment
(Roberts et al., 1996).

• From the latest Carboniferous through the Early Permian, deposition of bimodal volcanics,
volcaniclastic and siliciclastic sedimentary rocks in an extensional continental backarc setting
occurred throughout most of the NEO (e.g., Connors and Auburn subprovinces, Gogango Thrust
Zone, Rockhampton Subprovince, Tamworth Belt, Hastings Block, Peel-Manning FS, Silverwood
Province, Marlborough Province, Central Block N, North D’Aguilar and South D’Aguilar
subprovinces, Nambucca Block and Port Macquarie; Figure 2.16 to Figure 2.18). These rocks
overlie the older forearc and accretionary wedge successions.
• Latest Carboniferous and Early Permian extension is also recorded by the emplacement of
granites into, and formation of low-angle extensional faults in, the former accretionary wedge
(Glen, 2005). Late Carboniferous to early Permian granitoids of the New England Batholith
primarily intruded into the accretionary wedge assemblage, with minor intrusion into the Tamworth
Belt. This eastward shift in magmatism, at ca. 300 Ma, into the former accretionary wedge, is
interpreted to reflect the initial outboard retreat of the arc in the southern NEO (Collins et al., 1993).
This magmatism is mainly represented by the S-type Hillgrove and Bundarra Granite Suites (Shaw
and Flood, 1981). The S-type granitoid suites were emplaced before tectonism associated with the

76 Geodynamic Synthesis of the Phanerozoic of eastern Australia


contractional late Permian Hunter-Bowen Orogeny (see below), whereas I-type suites largely
intruded in the late Permian–Triassic (Shaw and Flood, 1981).
• Both Caprarelli and Leitch (1998, 2001) and Jenkins et al. (2002) proposed that Late
Carboniferous volcanism ceased at around 305 Ma, prior to the commencement of either slab
breakoff or rollback. Intrusion of the Hillgrove Suite at 302 Ma (Collins et al., 1993; Kent, 1994) was
interpreted by Caprarelli and Leitch (1998) as a response to the melting of sediments by upwelling
of asthenosphere, following slab breakoff, with intrusion occurring in a contractional environment.
Jenkins et al. (2002), however, suggested that rollback by 300 Ma had caused magmatism to
move eastward into the accretionary wedge, resulting in generation of the Hillgrove Suite in an
extensional environment. We prefer the latter interpretation.
• The Nambucca rift basin formed during backarc extension during the Early Permian. Sediments in
the Nambucca Basin were deposited unconformably onto the forearc successions of the Hastings
Block (Roberts et al., 1993), the implication being that the Hastings Block was in its current
position prior to early Permian sedimentation in the Nambucca Block (Johnston et al., 2002). This,
however, may not necessarily be the case; the Permian rocks of the Nambucca Block are highly
deformed and they may have been caught between the Coffs Harbour Block moving south and the
Hastings Block moving north during the period of oroclinal bending (R.J. Korsch, pers. comm.
2008).
• Tholeiitic and alkaline dolerite dykes with enriched geochemical signatures intruded the ophiolite
succession of the Marlborough Province in the Early Permian (293 ± 35 Ma, Bruce and Niu, 2000).
This magmatism is characterised by compositions typical of intraplate basalts, and a continental
intraplate origin is preferred by Bruce and Niu (2000).
• Extensional events in the Early Permian caused subsidence of the Sydney-Gunnedah Basin,
abundant basaltic and rhyolitic volcanism and onlap of the forearc basin (Leitch, 1988; Scheibner
and Basden, 1998; Roberts and Geeve, 1999) by the Permian sediments (Roberts et al., 2004).
• Volcanics in the Connors Subprovince, e.g., the Camboon Volcanics, are widely considered to
represent a superimposed, early Permian subduction-related episode, although associated forearc
basin or accretionary wedge elements have not been recognised (Holcombe et al., 1997a; Withnall
et al., 1998). The main units identified in the Connors Subprovince are the Camboon Volcanics,
Lizzie Creek Volcanic Group and the Nogo and Narayen beds (Figure 2.16, Figure 2.17), which we
have interpreted as an extensional continental backarc setting. Holcombe et al. (1997a) also have
suggested that the Connors Province represents an extensional event that is not necessarily
related in any aspect to active subduction.
• The time of oroclinal bending that produced the Texas-Coffs Harbour Oroclines has been the
subject of much debate. Some authors consider that it occurred in the late Carboniferous (310–
3early to mid-Permian (280–265 Ma; Korsch and Harrington, 1987; Cawood et al., 2011). Offler
and Foster (2008) suggest that development of the Texas and Coffs Harbour Oroclines took place
between 273–260 Ma. We consider oroclinal bending most likely took place after early Permian
backarc extension but prior to the main phase of the Hunter-Bowen Orogeny. Rosenbaum et al.
(2012) have suggested that orocline formation may have occurred in two or three phases: an
earlier subduction-related phase of curvature during slab rollback at c. 300–285 Ma, subsequent
dextral transpression-related bending, and possibly a final stage involving tightening during the
Hunter-Bowen Orogeny.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 77


78 Geodynamic Synthesis of the Phanerozoic of eastern Australia
Figure 2.18 Generalised distribution of rocks in the New England Orogen, and surrounds, by time period. A– late
Neoproterozoic to latest Cambrian (ca. 600 Ma to ca. 490 Ma), B – latest Cambrian to early Silurian (ca. 490–430
Ma), C – middle Silurian to late Devonian (ca. 430–380 Ma), D – Late Devonian to early Carboniferous (ca. 380
Ma to 350 Ma). E – early Carboniferous to late Permian (ca 350 to ca. 270 Ma), F – Late Permian to late Triassic
(ca. 270 Ma to ca. 220 Ma).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 79


Gympie Island Arc

• An inferred Early Permian backarc setting for the NEO and Sydney-Bowen Basin requires that an
arc environment existed outboard at this time (Figure 2.18). The early Permian Highbury Volcanics
of the Gympie Province have a chemical signature indicative of an arc setting (Sivell and
McCulloch, 1997). The submarine and subaerial island arc tholeiites, basaltic tuff breccias and
lavas that constitute the Highbury Volcanics are unconformably overlain by andesites and dacites
of the Rammutt Formation (Sivell and McCulloch, 2001). The primitive chemical and isotopic
character of these units implies a juvenile island-arc terrain, isolated from the influence of
continental crust (Sivell and Waterhouse, 1988). This is consistent with their intimate association
with primitive oceanic backarc (Cedarton and Cambroon) basalts (Sivell and McCulloch, 1997),
which suggests that a well-developed backarc separated the primitive Gympie island arc from the
NEO in the Early Permian.
• The intraoceanic Gympie island arc was located east of the continental margin of Gondwana in the
early part of the Permian. Detrital zircon data from Gympie (Korsch et al., 2009c) suggest that the
Gympie arc was attached to eastern Australia at the end of the Permian or start of the Triassic.

2.3.8 Late Permian to mid Triassic (ca. 265 Ma to ca. 230 Ma): Hunter-Bowen
Orogeny

2.3.8.1 Geological and tectonic summary

The late part of the Permian (~265–262 Ma) saw another change in the dynamics of the subduction
system, when a continental margin magmatic arc was re-established along the Paleo-Pacific
continental margin of Australia and the backarc changed from an extensional to a contractional regime
(Korsch and Totterdell, 1995). This led to the formation of a retroforeland fold-thrust belt west of the
magmatic arc, which was better developed in the northern NEO than the southern NEO (Korsch et al.,
1997). This contractional regime resulted in the development of a major retroforeland basin phase in
the Bowen-Gunnedah-Sydney basin system that continued until the Middle Triassic (Korsch and
Totterdell, 1995).

Thus, this period is characterised by renewed arc magmatism, the foreland basin stage of
development of the Bowen-Gunnedah-Sydney basin system and the Hunter-Bowen Orogeny
(Figure 2.18).

The Late Triassic (ca. 230 Ma) saw a switch in geodynamics back to an extensional, probably backarc
environment. This resulted in a change in plutonism to A-type granites, bimodal volcanism and
development of extensional basins with coal-bearing successions. This also marked the timing of
effective cratonisation of eastern Australia.

2.3.8.2 Geological history and Time-Space plot explanation

• A change from extensional to contractional tectonism began in the latest Early Permian, at ca. 270
Ma (Roberts et al., 1996). The Bowen-Gunnedah-Sydney basin system developed as a foreland
basin phase in a backarc setting that persisted until the Middle Triassic (Korsch et al., 2009a).
Foreland basin sedimentation is also recorded in the Gogango Thrust Zone, Rockhampton
Subprovince, Emu Creek Block, and in the Gympie Province (Figure 2.16 to Figure 2.18).
• The late Permian saw the renewed onset of subduction-related magmatism with voluminous Late
Permian to Early (and Middle) Triassic intrusive and extrusive magmatism occurring throughout the
NEO (Gust et al., 1993; Holcombe et al., 1997b; Van Noord, 1999; Withnall et al., pages 351–369,

80 Geodynamic Synthesis of the Phanerozoic of eastern Australia


in Donchak et al., 2013; Purdy, pages 344–399, in Donchak et al., 2013; Figure 2.18). Examples of
arc-related volcanism include late Permian to early Triassic deposition of Mount Wickham Rhyolite
(Connors Province) and Mount Eagle Volcanics (Gogango Thrust Zone) in a continental margin
magmatic arc setting. Continental margin arc-related rocks also occur at Emu Creek, Silverwood,
and Central Block North, with Triassic continental margin arc volcanic rocks in the Gympie
Province (Figure 2.17). Volcanism is predominantly andesitic (Holcombe et al., 1997b).
• Widespread intrusive magmatism also occurred at this time. These granites extend from the New
England Batholith (Shaw and Flood, 1981), in the south, to at least Rockhampton in the north (e.g.,
Murray, 2003; Withnall et al., pages 351–369, in Donchak et al., 2013; Purdy, pages 399–344, in
Donchak et al., 2013; Figure 2.18). About half of the exposed granites in the northern NEO have
K–Ar ages between 270 Ma and 230 Ma (Gust et al., 1993; Murray, 2003), although most of these
are reset ages. The granitoids are widely distributed and are predominantly of intermediate to
felsic, I-type, composition, and most seem to belong to the Clarence River Supersuite or
equivalents (Bryant et al., 1997; Murray, 2003). They are compositionally very similar to the earlier
(late Carboniferous to early Permian) magmatism (Murray, 2003). Some A-type magmatism is also
apparently present (Murray, 2003).
• The widespread calcalkaline magmatism strongly supports active subduction during this time with
compositional changes probably related to changes in arc configuration. Gust et al. (1993), for
example, suggested it may reflect changes in the angle of the subducting slab. As documented by
Gust et al. (1993), magmatism appears to wane in the Middle Triassic, although age control is
poor.
• This time interval encompasses the Late Permian to Middle Triassic orogenic event (which
includes deformation, metamorphism, magmatism, and foreland sedimentation) known as the
Hunter-Bowen Orogeny (Murray, 1997a, b; Holcombe et al., 1997b; Roberts et al., 2006; Korsch et
al., 2009b). In the New England Orogen, this phase of deformation is characterised by
retrothrusting driven by subduction further to the east. Subsidence of the Bowen and Gunnedah
basins during the foreland basin phase was driven by thrust loading due to westward-propagating
thrust sheets from the New England Orogen (Korsch et al., 2009b). The orogeny covers a period of
about 35 m.y. from ~265 Ma to ~230 Ma, (Holcombe et al., 1997b; Korsch et al., 2009b). The
foreland basin phase of sedimentation associated with the Hunter-Bowen Orogeny was punctuated
by a series of discrete contractional events (Figure 2.16), frequently producing unconformities
which were very short-lived (Korsch et al., 2009b). Contractional events of the Hunter-Bowen
Orogeny thrust the Tamworth Belt westwards over the eastern edge of the Sydney-Gunnedah
Basin and Lachlan Craton (Korsch et al., 1997; Roberts et al., 2004) in the Late Permian and
earliest Triassic.
• The Gympie region appears to have been the site of a continent-island arc collision and recent
detrital zircon age spectra from the Rammutt and Keefton formations (Korsch et al., 2009c) provide
some constraints on the accretion of the island arc component of the Gympie Province to the
eastern Australian margin. Zircon age data of Korsch et al. (2009c) suggest that the Gympie
Province came into contact with, and was sourced from, the accretionary wedge prior to deposition
of the Keefton Formation (at ca. 250 Ma).
• In contrast to the intermediate-dominated composition of granites of Early and Middle Triassic age
in the northern NEO, the late Triassic period (ca. 230–220 Ma) is characterised by intrusions of
dominantly silicic granite composition associated with the development of volcanic complexes of
rhyolite and minor mafic lavas (Stephens et al., 1993; Holcombe et al., 1997b), including A-type
magmatism (D. Champion, pers. comm., 2008). The change from backarc contraction and thrust
loading to backarc extension was probably at ~230 Ma (R.J. Korsch pers. comm. 2008).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 81


• Permian-Triassic plutonism was followed by widespread Late Triassic extension, characterised by
the development of small, elongate basins, associated with bimodal volcanics and interbedded
coal-bearing successions (e.g., Ipswich, Tarong, Lorne, Clarence-Morton basins; Holcombe et al.,
1997b). Plutonism waned at this stage, and was restricted to small plutons near the coast.
Therefore, following the Hunter-Bowen Orogeny, it appears that the New England Orogen reverted
to a retreating subduction boundary system during the Middle–Late Triassic (Jenkins et al., 2002).
This includes the Middle to Late Triassic deposition of sediments of the Esk Trough, also most
probably in an extensional continental backarc setting (R.J. Korsch, pers. comm., 2008).

2.4 Thomson Orogen, cover basins, and Koonenberry Belt


by DC Champion, modified from E Mathews, DC Champion, N Kositcin and C Brown (2009)

2.4.1 Introduction
The regions covered in this section include the Thomson Orogen – specifically the Anakie Inlier and
Fork Lagoons Province (Queensland), the Louth-Bourke region (New South Wales), the northeastern
part of the Delamerian Orogen - the Koonenberry Belt (New South Wales), and various
(Neoproterozoic to) Paleozoic basins (Georgina, Warburton, Cooper, Adavale, Drummond, Galilee,
Bowen-Gunnedah-Sydney, Surat) mostly overlying the Thomson Orogen or its contacts with either the
Delamerian or the New England orogens (refer to Figure 1.1 and Figure 1.2). The Warburton Basin is
covered in more detail in this edition as it has important correlations with the rocks of the Thomson
Orogen. The Neoproterozoic to Devonian Georgina Basin, to the northwest of the Thomson Orogen, is
also included in this discussion. These two basins, along with the Koonenberry region, provide
information on the geology at the western margin of the Thomson Orogen. Although the Charters
Towers, Greenvale, and Barnard provinces are now included in the Thomson Orogen (Fergusson and
Henderson et al., 2013), in this report they are chiefly discussed in the Mossman Orogen section, with
which they share many similarities.

The Thomson Orogen is perhaps the least understood of the eastern Australian orogens, largely
because of the thick succession of overlying basin sediments. Similarly, apart from the southern and
northern boundaries, the full extent of the Thomson Orogen is poorly defined. The northern and
western boundaries of the Thomson Orogen (and the western boundary of the Lachlan Orogen) are
largely defined by the enigmatic Tasman Line, which is traditionally considered to separate
Proterozoic rocks from Phanerozoic eastern Australia (see Direen and Crawford, 2003). Apart from the
type area in the north, however, where it is clearly evident based on geophysical anomalies that
truncate the Proterozoic Mount Isa Province, for much of its length it is nebulous and difficult to identify
unequivocally. This has resulted in numerous ‘Tasman Line’ interpretations in terms of both location
and tectonic significance (see Direen and Crawford, 2003). For this reason, and the extensive basin
cover, we have not attempted to fully delineate the western boundary of the Thomson Orogen north of
the Koonenberry region (Figure 1.1, Figure 1.2, Figure 2.20). This is in the region of the Warburton
Basin and it is possible it is closely related to the Thomson Orogen. The eastern boundary with the
New England Orogen is also poorly defined, because of overlying basin cover (i.e., Bowen Basin;
Figure 1.2). We have based our eastern boundary on Nd isotopic data (Champion, 2013), which
suggests that the boundary is east of that shown by Glen (2005). The Thomson-Lachlan seismic
survey images the southern boundary with the Lachlan Orogen, and it is defined by the major planar,
north dipping Olepoloko Fault (Glen et al., 2007c, 2013).

82 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Little is known about the age or nature of the basement to the Thomson Orogen (Figure 2.20). Most
information is based on outcropping parts of the orogen, such as the Anakie, Charters Towers and
Greenvale provinces in Queensland. Since the pioneering study of Murray (1994), both the
Queensland and New South Wales geological surveys have increasingly targeted available drill core
from holes which have intersected the orogen. Brown et al. (2012), for example, document nearly
1400 drill holes, in their recent database compilation for the Queensland part of the undercover
Thomson and surrounds. Both surveys have reported an increasing amount of geochronological data
in recent years (e.g., Black, 2005, 2006, 2007; Draper, 2006, Bodorkos et al., 2013; Glen et al., 2013;
Fraser et al., 2014; Kositcin et al., 2015a; Cross et al., 2015b). Additional new information is also
becoming available with recent work, including seismic acquisition and interpretation in the New South
Wales part of the orogen (e.g., Glen et al., 2013), complementing earlier seismic surveys in the
Queensland part of the orogen (e.g., Korsch et al., 2012). Originally, the Thomson Orogen was
defined by Kirkegaard (1974) as being Cambrian to Carboniferous in age and, hence, largely
equivalent to the Lachlan Orogen. New seismic data across the southern part of the Thomson Orogen
show the two orogens, at least in that transect area, have different lower crustal characters, with the
Thomson possessing thicker crust (Moho at 48 km) than the Lachlan (Moho at 32 km; Glen et al.,
2007c, 2013). The thinner, more reflective crust of the Lachlan confirms a major difference in the
crustal character between the two orogens (Glen et al., 2007c, 2013). In addition, Thomson basement
rocks, at least locally (Charters Towers, Anakie provinces; Figure 2.19, Figure 2.20), consist of
Neoproterozoic (inferred) and early Cambrian metasedimentary rocks that contain evidence of the
Delamerian Orogeny (Withnall et al., 1995; Fergusson et al., 2007c; Fergusson and Henderson et al.,
2013), making them different to the majority of the Lachlan Orogen. Dissimilar lithologies suggest that
the depositional setting of the Thomson Orogen was not related to that of the Lachlan Orogen and the
two orogens preserve different histories prior to at least the Middle to Late Ordovician (e.g.,
Kirkegaard, 1974; Draper, 2006; Glen et al., 2013; Fergusson and Henderson et al., 2013).

2.4.2 Late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma):
Rodinian breakup

2.4.2.1 Geological and tectonic summary

Rocks of this age are best represented in the Anakie Province, but also occur in the Charters Towers
and Greenvale regions, the Koonenberry Belt, the Georgina Basin and possibly the central Thomson
Orogen (Figure 2.7, Figure 2.19, Figure 2.20). All these regions, except for the central (and southern)
Thomson Orogen, have reasonable age constraints. The tectonic environment of this period is
typically interpreted as passive margin and/or rifting related to Rodinian breakup of the Precambrian
supercontinent (e.g., Crawford et al., 2003a; Fergusson et al., 2007c, 2009; Greene, 2010; Greenfield
et al., 2011). The late Neoproterozoic to early Cambrian (ca. 580 Ma to 520 Ma) saw deposition of
marine sediments, as well as, local magmatism, in the Charters Towers and Greenvale provinces, in
the Anakie Province (Anakie Metamorphic Group) and in the Koonenberry Belt, Georgina Basin, and,
possibly the Warburton Basin and southern Thomson Orogen (Withnall et al., 1995; Hutton et al.,
1997; Fergusson et al., 2001; 2007c, 2009; Gilmore et al., 2007; Greene, 2010; Greenfield et al.,
2010, 2011; Sun and Gravestock, 2001; Glen et al., 2013; Figure 2.19, Figure 2.20). Tholeiitic and/or
alkaline magmatism, consistent with rifting and/or passive margin volcanism is present within the
Bathhampton Metamorphics in the Anakie Province (mafic schist; Withnall et al., 1995; Fergusson et
al., 2009), within the Cape River Metamorphics (amphibolite; Withnall et al., 1997b; Hutton et al.,
1997), and the ca. 585 Ma alkaline Mount Arrowsmith Volcanics of the Grey Range Group (alkali
basalt, trachybasalt, trachyte, submarine and subaerial lavas, pyroclastics and related intrusives;
Crawford et al., 1997; Black, 2007; Greenfield et al., 2010, 2011; Figure 2.19, Figure 2.20). The

Geodynamic Synthesis of the Phanerozoic of eastern Australia 83


available geochronology, therefore, strongly suggests that a number of parts of the Thomson Orogen,
e.g., Anakie Metamorphic Group, do indeed correlate with latest similar-aged Neoproterozoic-
Cambrian units in the Delamerian Orogen, such as the Adelaide Fold Belt (South Australia) and the
Koonenberry Belt (western New South Wales) (Fergusson et al., 2001). Models for eastern Australia
at this time commonly invoke a rifted to passive margin setting (e.g., Glen, 2005, 2013) and the
Thomson, especially the north-eastern Thomson is no different, e.g., Fergusson et al. (2001, 2009).
General models for Rodinia breakup (e.g., Li et al. 2008) invoke major rifting at ca. 830 Ma and after.
This is well represented in central and southern Australia, e.g., within the Centralian Superbasin
(Walter et al., 1995; Greene, 2010) and within the Adelaide Fold Belt and associated basins
(Figure 2.7). In both these regions basin formation is thought to be related to a plume at ca. 830 Ma––
the remnants of which are represented by the Gairdner Dyke Swarm––and a number of successive rift
events are evident. Preiss (2000), for example, documents five periods of rifting (and mafic
magmatism), beginning at ~827 Ma and ending at ca. 700 Ma. Greene (2010) also discusses similar
rifting events in the Georgina Basin. These older rocks are not recorded in the Koonenberry or
Thomson regions, with the best evidence being for ca. 580 Ma rifting and mafic magmatism (e.g.,
Crawford et al., 1997), which may also be present in the Georgina Basin (Greene, 2010). Fergusson
et al. (2009) suggested the apparent absence of pre 600 Ma rifting in the Thomson orogen may be
explained by rifting off of a microcontinent fragment (removing the record of the older rifting).
Greenfield et al. (2011), in their tectonic model of the Koonenberry region, also invoke rifting of a
Proterozoic block at this time.

2.4.2.2 Extent of pre-Delamerian orogeny basement in the Thomson Orogen

Given the evidence for pre-Delamerian Orogeny rocks within the northeastern parts of the Thomson
Orogen, the major question becomes how far do they extend to the southeast, in the under-cover
(beneath the Eromanga and other basins) portion of the orogen? Lithologies intersected in drill core
beneath these basins include metasediments, volcanic rocks and intrusive rocks (Murray, 1994), each
of which provide constraints on the age and types of basement.

Metasedimentary rocks within the Thomson Orogen have been tentatively correlated with late
Neoproterozoic to Middle Cambrian rocks in the Anakie Province and the Charters Towers region
(e.g., Draper, 2006). Murray (1994) documented (non-SHRIMP) ages from basement cores in the
Thomson Orogen that suggest that Neoproterozoic to Middle Cambrian rocks may be prevalent. In
addition the extension of the Warburton Basin into the western part of the Thomson Orogen
(Figure 3.6) also implies older basement rocks may exist. Direct evidence for possibly older
metasedimentary rocks comes from detrital zircon geochronology, both by maximum depositional
ages and provenance studies (e.g., Fergusson et al., 2001, 2007c; Kositcin et al., 2015a; Cross et al.,
2015b, in prep). Difficulties are present in both approaches. As documented by Kositcin et al. (2015a)
maximum depositional age data not only show that some sedimentary rocks have pre-Delamerian
maximum ages, but that there is also mounting evidence for multiple ages of sedimentary sequences,
often in close proximity. This makes it difficult to not only confidently predict the presence of any old
(middle Cambrian) basement in the undercover Thomson Orogen, but also its extent, even where
geochronology is available. This is exacerbated by the overlap of maximum depositional ages with
inferred ages of the Delamerian Orogeny in this region (both ca. 500–495 Ma), making it difficult to
ascertain whether the rocks in question are actually pre- to syn-orogeny or are in fact post-orogeny.

Detrital zircon ages obtained by Fergusson et al. (2001, 2007c) show that two major rock successions
were sourced from the east Gondwana margin. The apparently older succession, i.e., late
Neoproterozoic in age (ca. 600 Ma), generally contain abundant ca. 1200 Ma detrital zircons
suggesting that a Grenville-aged orogenic belt may have existed in northeastern Australia to provide
the major sediment source (Fergusson et al., 2001). These rocks also contain only minor 1870–1550

84 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Ma zircons, indicating at best only minor input from cratonic older Proterozoic terranes, such as Mount
Isa and Georgetown (Fergusson et al., 2007c). The younger succession, which includes the Wynyard
Metamorphics (upper part of the Anakie Metamorphic Group, Anakie Province) and Cambrian–
Ordovician upper Argentine Metamorphics (Charters Towers region), is early Palaeozoic in age, and is
dominated by the 600–500 Ma Gondwana signature, not dissimilar to that seen in the Lachlan Orogen
and relatively widespread in the Thomson Orogen (e.g., Kositcin et al., 2015a; Cross et al., 2015b).
These results also highlight further complexity in determining basement to the Thomson Orogen. In
particular, the documented close spatial association of metasedimentary rocks with perhaps two ages,
but certainly with two distinct detrital zircon provenances (and maximum depositional ages), e.g.,
Anakie Metamorphic Group (Fergusson et al., 2001) and Argentine Metamorphics (Fergusson et al.,
2007c), is problematic.

Indirect evidence for older (pre-Delamerian) basement is also provided by both extrusive and intrusive
felsic magmatic rocks of Ordovician age. Glen et al. (2013) have suggested that mafic and
intermediate volcanics in the southern Thomson Orogen may be Neoproterozoic in age. The age
constraints of these, however, are not definitive and more work is required to confirm these ages.
Better evidence is available in the central Thomson Orogen, where metasedimentary rocks are
overlain by apparently subhorizontal Early Ordovician volcanic rocks (and intruded by Ordovician
granites; Figure 3.6). This was interpreted by Draper (2006) to show that the sedimentary rocks were
deformed (thickened) and metamorphosed prior to this (Draper, 2006), i.e., during the Delamerian
Orogeny, an interpretation that may or may not be supported by recent geochronology which suggests
a 496 ±17 Ma maximum depositional ages for the underlying sediment (Kositcin et al., 2015a).
Regardless, and more importantly, the very presence of the early Ordovician felsic magmatism
suggests the country rocks they occur within have been through a cycle of prior crustal thickening.
Their relatively widespread presence (Figure 3.6) is not consistent with models advocating post-
Delamerian sedimentation directly onto sea floor (e.g., Glen, 2013), though do not rule out such a
model if sedimentation occurred prior to the Delamerian Orogen or the latter is much younger in this
region. The extent of the early Ordovician (pre-470 Ma) granites and felsic volcanic rocks can be used
to provide inferential evidence for the minimum extent of older (pre-Delamerian) basement in parts of
the non-outcropping Thomson Orogen (Figure 3.6). It is perhaps the best evidence for older basement
given the equivocal nature of the detrital zircon data for the metasedimentary rocks.

2.4.2.3 Geological history and Time-Space plot explanation (Figure 2.19, Figure 2.20)

Anakie Province (Anakie region)

• The Anakie Metamorphic Group consists of widespread remnants of largely marine


metasedimentary rocks (Withnall et al., 1995; Fergusson et al., 2007c; Fergusson and Henderson
et al., 2013). Detrital zircon and monazite ages from the Anakie Metamorphic Group suggest a
range of ages for this group, although not all units have been dated. These include maximum
deposition ages of 1300–1000 Ma and greater than 600 Ma for the Bathampton Metamorphics
(Fergusson et al., 2001; Fergusson and Henderson et al., 2013), suggesting deposition continued
until the late Neoproterozoic. Detrital zircon and monazite ages from the Wynyard Metamorphics
show three age components at ca. 580–570 Ma, ca. 540 Ma, and ca. 510 Ma (Fergusson et al.,
2001, 2007c). Sediments probably relate to, or were derived from, Rodinian rifting along the
Gondwanan passive margin (Fergusson et al., 2001, 2007c). The Anakie Province is thought to
extend undercover to at least the Nebine Ridge to the south and it is possible that correlates of the
Anakie Metamorphic Group extend to there (e.g., Withnall et al., 1995).
• Both tholeiitic and alkaline magmatism, consistent with rifting and/or passive margin volcanism, are
present within the Anakie Metamorphic Group in the Anakie Province (mafic schist; Withnall et al.,
1995, Fergusson et al., 2009).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 85


Figure 2.19 Late Neoproterozoic to Cretaceous time-space plot for the Thomson Orogen, Koonenberry Belt, and cover basins. Refer to text for data sources and discussion. Orogens, regions and basins are as shown in Figure 1.1 and Figure 1.2.

86 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Central Thomson Orogen

• Ages for metasedimentary rocks within the central Thomson Orogen are poorly constrained though
have been tentatively correlated with late Neoproterozoic to Middle Cambrian rocks in the Anakie
Province and the Charters Towers region (Draper, 2006). Overlying apparently subhorizontal
volcanic rocks have Early Ordovician ages (Draper, 2006) and so the underlying metasediments
have been interpreted to have been deformed and metamorphosed prior to this (Draper, 2006).
Early Ordovician ages have also been recorded for granites (Draper, 2006), providing a minimum
age for the metasediments. Murray (1994) documents a number of K-Ar ages (ca. 540–480 Ma)
from basement cores in the Thomson Orogen.
• The Warraweena Volcanics, comprising basalts and basaltic-andesites with calcalkaline affinities
occur in the southern Thomson Orogen (Burton et al., 2008; Glen et al., 2013). Glen et al. (2013)
suggested, on the basis of U-Pb zircon data, that the unit may be Neoproterozoic (ca. 582 Ma).
The age is not well constrained, however, as younger minor peaks are also present, e.g., at 504
Ma. In addition the zircon data (in Glen et al., 2013) shows a pattern with many peaks more
suggestive of some sedimentary input. This interpretation seems consistent with the range of
epsilon Hf seen in the zircons, particular the younger grains (<700 Ma) which have evolved
signatures (Glen et al., 2013). In addition, Fraser et al. (2014) report a maximum deposition age of
ca. 438 ± 9 Ma for a siltstone near to the Warraweena Volcanics, suggesting a younger age for this
unit. Accordingly, the 580 Ma age advocated by Glen et al. (2013) is probably best thought of as a
maximum age. The unit appears to be overlain by sediments of the Paka Tank Trough and
possibly also intruded by Silurian–Devonian granites (Figure 3 of Glen et al., 2013), suggesting the
unit falls somewhere between Neoproterozoic and Silurian in age.

Koonenberry Belt

• Sediments and volcanics of the late Neoproterozoic Grey Range Group (continental shelf to deep
marine) and interpreted equivalents, the Adelaidean Farnell Group, west of the Bancannia Trough,
formed during intracontinental rifting associated with Rodinian break-up (Crawford et al., 1997;
Gilmore et al., 2007; Greenfield et al., 2010, 2011).
• The Mount Arrowsmith Volcanics (Grey Range Group) (alkali basalt, trachybasalt, trachyte,
submarine and subaerial lavas, pyroclastics and related intrusives) are dated at ca. 585 Ma
(Crawford et al., 1997; Black, 2007). Ultramafic and mafic rocks (peridotites and gabbros) were
also formed at this time although their exact age and relationship to the Mount Arrowsmith
Volcanics is unknown (Crawford et al., 1997). Greenfield et al. (2011) indicate that basaltic rocks
with chemical affinities to the Mount Arrowsmith Volcanics occur within the Farnell Group and
possible equivalent rocks also occur to the southwest in the Loch Lilly–Kars Belt.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 87


88 Geodynamic Synthesis of the Phanerozoic of eastern Australia
Figure 2.20 Generalised distribution of rocks in the Thomson Orogen, Koonenberry region and cover basins, by time
period. A– late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma), B – early to latest Cambrian (ca. 515
Ma to ca. 490 Ma), C – latest Cambrian to early Silurian (ca. 490–430 Ma), D – middle Silurian to late Devonian (ca.
430–380 Ma), E – Late Devonian to early Carboniferous (ca. 380 Ma to 350 Ma). F – early Carboniferous to late
Permian (ca 350 to ca. 270 Ma), G – Late Permian to late Triassic (ca. 270 Ma to ca. 220 Ma).

2.4.3 Early to latest Cambrian (ca. 515 Ma to 490 Ma)

2.4.3.1 Geological and tectonic summary

The break-up of the Rodinia Supercontinent and the associated extension of eastern Australia
between the Late Neoproterozoic (ca. 600 Ma) and the Early Cambrian ended with the development of
subduction and subsequent contractional orogenesis—the Delamerian cycle and orogeny. In southern
Australia, the Delamerian Orogeny commenced at ca. 515 Ma and ended at ca. 490 Ma (e.g.,
Seymour and Calver, 1995; Foden et al., 2002a, 2006; Figure 2.7). In the Thomson region, the best,
and earliest, evidence for a subduction-related setting is in the Koonenberry region—the Mount Wright
Arc of Greenfield et al. (2011)—exemplified by the the calc-alkaline suite of the Mount Wright
Volcanics but also including Cambrian sediments of the Teltawongee Group (continental slope
turbidites), Ponto Group (fine-grained marine clastic rocks and tuffs) and volcanic and sedimentary
rocks of the Gnalta Group (shallow marine/shelf sediments; Crawford et al., 1997; Gilmore et al., 2007;
Greenfield et al., 2010, 2011; Figure 2.19, Figure 2.20).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 89


Greenfield et al. (2010, 2011) indicate that the Mount Wright Volcanics have both arc and intra-plate
chemical characteristics, and suggested arc-backarc affinities for the Gnalta Group. They invoked a
continental arc and a west-dipping subduction zone, with the back-arc located largely within the
Bancannia Trough, and the Ponto Group forming in a fore-arc setting. In contrast, Crawford et al.
(1997) suggested the Mount Wright Volcanics formed in an immature continental rift, evolving to later
tholeiitic magmatism with more extension. The presence of volcanics rocks of the Bittles Tank
Volcanics in the Ponto Group with both tholeiitic and alkaline chemistries appears to lend support to
the interpretation of Crawford et al. (1997), though Greenfield et al. (2011) cite examples of similar
rocks in modern fore-arc environments. The best age constraint for the Delamerian cycle and orogeny
is provided by the ca. 516 Ma Williams Peak Granite which is thought to be pre- to synkinematic
(Black, 2007). Greenfield et al. (2011) note that Teltawongee Group sediments disconformably overlie
metasediments of the Neoproterozoic Kara Group which they interpreted to indicate a lack of any
tectonic disturbance between deposition of the two groups. This has implications for models, such as
Gibson et al. (2011), that advocate collision of an island arc and a change in subduction polarity, in
southeastern Australia and Tasmania, prior to the Delamerian Orogeny—either such a scenario did
not occur, or if it did, then it did not reach this far north.

The Mount Wright Arc may have continued northward into the Warburton Basin region (Figure 2.7,
Figure 2.19). The lower part of the basin contains the early Cambrian, ca. 517 Ma (and younger?)
Mooracoochie Volcanics (Gatehouse, 1986; Boucher, 2001). The Mooracoochie Volcanics are thought
to be correlatives of the Cambrian volcanics in the Koonenberry Belt (Gravestock and Gatehouse,
1995), and were emplaced along what has been called the Gidgealpa Volcanic Arc (Gatehouse, 1986;
Boucher, 2001). Middle Cambrian volcanism is also recorded by a ca. 510 Ma rhyolitic ignimbrite (DIO
Adria Downs-1 well), north of the Queensland-South Australia border (Draper, 2006), suggesting a link
with magmatism in the Warburton Basin and Koonenberry region. There is no evidence for any arc-
related magmatism within the south-eastern Georgina Basin at this time. Mafic magmatism of this age
does occur within the Queensland part of the Georgina Basin, e.g., the Colless Volcanics (Jell in
Withnall and Hutton et al., 2013) but this is related to the Kalkarindji Large Igneous Province (Glass
and Phillips, 2006) which is contemporaneous with the calc-alkaline magmatism in eastern Australia.

Rocks of this age do not appear to have been recorded in the south-western Thomson Orogen, i.e.,
next to the Koonenberry region. Similarly, there is little or no evidence for arc-related magmatism
elsewhere in the Thomson Orogen at this time, though the lack of exposure and uncertainty of ages
means such rocks could have been present (though it is noted that there is also little or no evidence
for such rocks further north in northern Queensland region either). It is possible, for example that the
Warraweena Volcanics in the south-eastern Thomson Orogen (Burton et al., 2008; Glen et al., 2013)
may be this age.

The possible presence of an arc in the Warburton-Koonenberry region at this time, well west of the
Anakie Province, is problematical. Withnall et al. (1995) suggested this may have either reflected a
very wide Delamerian Orogen or subsequent (post-Delamerian) extension and rifting of the Anakie
Province eastwards. The latter does not appear consistent with the discovery of both granites
intruding, and subhorizontal early Ordovician volcanic rocks overlying, deformed and metamorphosed
metasedimentary rocks in the central Thomson Orogen (Draper, 2006; see below).

The cycle ended with the Delamerian Orogeny which deformed and metamorphosed rocks of the
Anakie Province, Charters Towers region, Koonenberry Belt and central Thomson Orogen, and
generally resulted in downwarping of the Warburton Basin (Murray and Kirkegaard, 1978), and
discontinuities in the Georgina Basin (Webby, 1978). It is best recorded in basement rocks of the
Anakie and Koonenberry regions. Based on comparative evidence from the Anakie Province, Draper
(2006) suggested that deformation of subsurface metasedimentary rocks in the eastern Thomson

90 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Orogen was likely to have occurred during the Delamerian Orogeny. The contractional event was
predominantly east-west in the Thomson Orogen, and northwest-southeast in the Koonenberry Belt
(Gilmore et al., 2007; Greenfield et al., 2010). Uplift and erosion of the Warburton Basin was
coincident with the Delamerian Orogeny (locally recognised as the Mootwingee Movement;
Gravestock and Gatehouse, 1995), producing a brecciated interval and depositional hiatus (Boucher,
2001; Harvey and Hibburt, 1999). This event also deformed the Adelaide and Kanmantoo orogens,
and broadly similar ages have been recognised in Koonenberry, Charters Towers and South Australia,
although deformation appears to have been shorter-lived in northern Australia (Foden et al., 2006;
Black, 2007; Fergusson et al., 2007a, b).

2.4.3.2 Geological history and Time-Space plot explanation (Figure 2.19, Figure 2.20)

Anakie Province

• The Anakie Metamorphic Group was complexly deformed and metamorphosed, to greenschist and
amphibolite facies, in the Middle Cambrian, ca. 500 Ma (Withnall et al., 1995; 1996; Green et al.,
1998; Fergusson et al., 2007c; Fergusson and Henderson et al., 2013). Withnall et al. (1996)
suggested that this deformation represented the northern continuation of the Delamerian Orogen.
Offler et al. (in Fergusson and Henderson et al., 2013) report medium pressures (0.79–0.64 GPa)
and high temperatures (640–580 °C) for this deformation.
• Syn- or post-orogenic magmatism occurs within the Anakie Province (Crouch et al., 1995, Withnall
et al., 1995). Magmatism includes the syn-deformation S-type Mooramin Granite (ca. 473–463 Ma;
Richards et al., in Fergusson and Henderson et al., 2013) and the ca. 473 Ma (L.P. Black, pers.
comm., cited in Fergusson and Henderson et al., 2013) I-type Coquelicot Tonalite. The ages
suggest these units are post-Delamerian.

Thomson Orogen

• Middle Cambrian volcanism is recorded at ca. 510 Ma (Draper, 2006) in basement rhyolitic
ignimbrite beneath the Poolowanna Trough and Eromanga Basin (DIO Adria Downs-1 well, north
of the Queensland-South Australia border). The exact relationship of these volcanics to the
Mooracoochie Volcanics (trachyte and dacites) in the Warburton Basin is unresolved due to
overlying sedimentary basin cover.
• The Warraweena Volcanics, comprising basalts and basaltic-andesites with calcalkaline affinities,
occur in the southern Thomson Orogen (Burton et al., 2008; Glen et al., 2013). These rocks have
poor age constraints (Neoproterozoic to Silurian; see Section 2.4.2).

Koonenberry Belt

• Rocks of the Neoproterozoic Grey Range Group are overlain by Cambrian sediments of the
Teltawongee Group (continental slope turbidites), Ponto Group (fine-grained marine clastic rocks
and tuffs) and volcanic and sedimentary rocks of the Gnalta Group (shallow marine/shelf
sediments) which includes the calcalkaline suite of the Mount Wright Volcanics (Crawford et al.,
1997; Gilmore et al., 2007; Greenfield et al., 2010, 2011). Tuffs in the Ponto Group have been
dated at ca. 512 Ma and 508 Ma (Black, 2005). These ages are similar to those reported for
calcalkaline volcanism (andesite, basalt, rhyolite and dacite) within the Mount Wright Volcanics
(Black, 2007). Based primarily on geophysical interpretation, but also on some drill hole
intersections, Greenfield et al. (2011) suggested that the Mount Wright Volcanics are much more
extensive, and probably also occur within both the Bancannia Trough and in the Loch Lilly–Kars
Belt. Volcanism in the Mount Wright Volcanics and Gnalta Group is linked to the Delamerian
Orogeny.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 91


• Contractional, south-southeast–north-northwest deformation and associated low-grade regional
metamorphism was recognised by Greenfield et al. (2010, 2011) in the Koonenberry region.
Deformation resulted in tight, west-vergent folding and thrusting with a component of sinistral
strike-slip movement (e.g., Gilmore et al., 2007). The Williams Peak Granite provides the best
evidence for the onset of the Delamerian Orogeny. It has been dated at ca. 516 Ma and indicates
intrusion was pre- to synkinematic (Black, 2007). The end of Delamerian deformation is marked by
a felsic intrusives cross-cutting the Delamerian foliation, and has been dated at ca. 497 Ma (Black,
2005).

Warburton Basin

• Porphyritic trachyte and dacitic lavas from the Mooracoochie Volcanics have been dated at ca. 517
Ma (Malgoona-1 well; U-Pb zircon; Boucher, 2001). This date correlates with the U-Pb age of ca.
510 Ma for an unnamed rhyolitic ignimbrite beneath the Cooper Basin (Adria Downs-1 well;
Draper, 2006), and suggests these two units are equivalents (Draper, 2006). The Mooracoochie
Volcanics are also thought to be correlatives of the Cambrian volcanics in the Koonenberry Belt
(Gravestock and Gatehouse, 1995).
• Delamerian uplift and erosion (i.e., Mootwingee Movement of Gravestock and Gatehouse, 1995)
produced a depositional hiatus adjacent to the Australian craton prior to the onset of marine
sedimentation (e.g., Dullingari Group).

2.4.4 Latest Cambrian to early Silurian (ca. 490 Ma to 430 Ma)

2.4.4.1 Geological and tectonic summary

Middle Cambrian, Ordovician and early Silurian rocks in the Thomson Orogen and surrounds fall into
two end-members:

• Sequences dominated by shelf and deltaic to deep marine sedimentation (Koonenberry,


Warburton Basin, Georgina Basin, Southern Thomson) with local, tholeiitic mafic magmatism (e.g.,
Warburton Basin), and;
• Sequences characterised by (interpreted) backarc volcanism, plus or minus sediments (Fork
Lagoons, Greenvale, Charters Towers), and regions dominated by felsic intrusive and extrusive
rocks (the Central Thomson; Figure 2.7, Figure 2.19, Figure 2.20).

In the southern Thomson, Koonenberry, Warburton and Georgina basin regions, the Delamerian
Orogeny was followed by a return to marine basin sedimentation. Progressive marine incursion
caused an expansion of basins across central Gondwana as a shallow epieric sea (the Larapintine
Sea; Webby, 1978), believed to link the Warburton, Amadeus, Georgina and perhaps the Canning
basins (e.g., Gravestock and Gatehouse, 1995; Haines and Wingate, 2007; Maidment et al., 2007).
This seaway also connected basinal environments in the Koonenberry region. This is particularly
evident in commonalities between the Warburton Basin and Koonenberry rocks of this age. In both
regions there is an evident change in sedimentation from terrestrial or shallow marine to deeper
marine basinal sediments, though the polarity varies: southwest to northeast in the Koonenberry
region (Greenfield et al., 2010), and northwest to southeast in the Warburton Basin (e.g., Sun and
Gravestock, 2001), i.e., sediments were deposited in shelf and trough environments in both the
Koonenberry region and the Warburton Basin, presumably related to the Larapintine Sea. Importantly,
in the Koonenberry region, the east-west change in sedimentation straddles the boundary between the
Delamerian and Thomson orogens, with the deep water sediments of the Warratta Group occurring in
both (Greenfield et al., 2010). The Warburton Basin sediments are also interpreted to continue north-

92 Geodynamic Synthesis of the Phanerozoic of eastern Australia


eastwards into a substantial part of the Thomson Orogen in Queensland (Figure 2.20), and Murray
(1994) suggested that many of the basement cores drilled in that area of the Thomson Orogen
resembled the Dullingari Group of the Warburton Basin. Murray (1994) further suggested that perhaps
most of the siliciclastic sediment basement in the Thomson could be correlated with the Warburton
Basin, i.e., were Ordovician in age (similar to those in the Lachlan Orogen also), though, how the
Ordovician volcanics in the central and northern Thomson fit this scenario is uncertain. Certainly, the
geological evidence suggests that the Koonenberry belt and Warburton Basin are closely related and
perhaps, at least parts of, the Thomson Orogen and the Warburton Basin are as well.

Early–Middle Ordovician extension accompanying sedimentation, has been recorded in numerous


places. These include the granulite-facies metamorphism at mid-crustal depths between the Amadeus
and Georgina basins–called the Larapinta Event, e.g., Maidment et al., 2007), and extensional
deformation recorded in the Charters Towers (Fergusson et al., 2005, 2007b; Cross et al., 2015a) and
Anakie (Fergusson et al., 2005; Richards et al., in Fergusson and Henderson et al., 2013; Wood and
Lister, in Fergusson and Henderson et al., 2013) regions.

This extension is also evident within the magmatic record. In the Thomson Orogen, magmatism took
place in the Early Ordovician, Middle Ordovician and Middle Silurian (Murray, 1994; Draper, 2006).
This magmatism forms part of the widespread Macrossan Igneous Association in northern
Queensland (e.g., Hutton et al., 1997; see Section 2.2.4; Figure 2.19, Figure 2.20). Mafic to felsic
magmatism of the Macrossan Igneous Association is best developed in the Charters Towers region,
which is dominated by I-type and mantle intrusive and extrusive magmatism with ages from ca. 490
Ma to ca. 455 Ma (e.g., Hutton et al., 1997; Hutton in Fergusson and Henderson et al., 2013). The
magmatism can be subdivided into Early to Middle and Middle to Late Ordovician. Lower Ordovician
magmatism is present as felsic volcanic rocks and granites in the central part of the orogen (Draper,
2006) and as mafic, intermediate and felsic volcanism and intrusives in the northern parts, specifically
in the Greenvale and Charters Towers provinces. The volcanics rocks in the latter regions have been
interpreted as being back-arc in origin (Henderson, 1986; Stolz, 1994; Fergusson et al., 2007a). Upper
Ordovician volcanism also includes rocks in the Forks Lagoon area (Anakie Province), interpreted as
either backarc or island arc (Withnall et al., 1995), and volcanic rocks of the Lucky Springs
Assemblage in the Broken River Province, interpreted as an island arc assemblage (Henderson et al.,
2011, 2013). Intrusive rocks of this age are widespread. Possible backarc and arc-related volcanism––
the Warraweena Volcanics, e.g., Burton et al. (2008)––also occur in the southern Thomson Orogen, in
the Bourke region. The age of these rocks is uncertain and may include Neoproterozoic to Silurian–
Devonian components (e.g., Glen et al., 2013). Burton et al. (2008) suggested an Ordovician age,
based on chemical similarities with the Macquarie Arc. The Warraweena Volcanics and the Mount
Dijou samples are characterised by both calcalkaline to shoshonitic, and alkaline (with OIB-like
patterns), compositions (Watkins, 2007; Burton et al., 2008). The calcalkaline compositions appear to
be similar to the Ordovician Lachlan Macquarie Arc (Burton et al., 2008), and Watkins (2007)
suggested the possible presence of a contemporaneous oceanic arc in the Thomson Orogen. Watkins
(2007) appears to suggest a south-dipping slab, effectively putting the alkaline rocks, which occur
south of the calcalkaline rocks (Figure 2.20) in a backarc position. Notably, Glen et al. (2013) report
detrital zircons from a volcaniclastic sandstone in the nearby Silurian–Devonian Louth Volcanics which
have Ordovician ages (ca. 487–441 Ma) and relatively primitive Hf isotopic signatures suggesting a
provenance comprised of primitive magmatic rocks, not unlike the Macquarie Arc magmatism. Such a
provenance may have either been present within the southern Thomson region at that time or
alternatively was external to it; a question that may be resolved with further geochronology.

The Benambran cycle ended with the Early Silurian Benambran Orogeny. This orogeny is well
represented in the northern parts of the Thomson Orogen, e.g., Charters Towers, Greenvale regions
(Fergusson et al., 2005, 2007b), probably the Anakie region (e.g., Withnall et al., 1995), and the

Geodynamic Synthesis of the Phanerozoic of eastern Australia 93


Koonenberry Belt (Greenfield et al., 2010). Based on metamorphic ages (e.g., Draper, 2006), it is
presumed that the Benambran deformed other rocks of the Thomson basement, but the full areal
extent of Benambran deformation in the Thomson Orogen is yet to be resolved. The Benambran
Orogeny also caused uplift and associated regression of the Larapintine Sea (Gravestock and
Gatehouse, 1995). Henderson et al. (2011) have indicated that the Benambran Orogeny in northern
Queensland is related to accretion of the island arc (represented by the Lucky Springs Assemblage) in
the Broken River Province, broadly contemporaneous with accretion of the Macquarie Arc in the
Lachlan Orogen.

2.4.4.2 Geological history and Time-Space plot explanation (Figure 2.19, Figure 2.20)

Anakie region

• Late Ordovician marine metasedimentary rocks including carbonates and associated mafic to
intermediate volcanic rocks of the Fork Lagoons beds (Withnall et al., 1995; Fergusson and
Henderson et al., 2013). Intrusion of gabbro was comagmatic with basalt and andesite extrusion;
the geochemistry of these rocks suggests tholeiites of either island arc or backarc origin (Withnall
et al., 1995). The Fork Lagoons beds also contain structurally emplaced serpentinite (Withnall et
al., 1995). The metasedimentary rocks of the Fork Lagoons beds were apparently sourced from
both cratonic and volcanic provenances (e.g., Fergusson et al., 2007c), which led Withnall et al.
(1995) to suggest an arc setting not too far from a continent. Withnall et al. (1995) also suggested
that the relationship between the Fork Lagoons beds and the older Anakie Metamorphic Group
was similar to that observed between early Paleozoic and Neoproterozoic rocks in the Georgetown
region.
• Minor I-type (ca. 471 Coquelicot Tonalite; L.P. Black, pers. comm., cited Fergusson and
Henderson et al., 2013) and S-type (ca. 460 Ma Mooramin and ca. 443 Ma Gem Park granites;
Richards et al., in Fergusson and Henderson et al., 2013; Wood and Lister, in Fergusson and
Henderson et al., 2013) magmatism is present, intrusive into the Anakie Metamorphic Group
(Fergusson and Withnall, in Fergusson and Henderson et al., 2013). The Gem Park Granite is
synchronous with the Benambran Orogeny, while the Mooramin Granite relates to extensional
deformation(?) (Richards et al., in Fergusson and Henderson et al., 2013).
• Poorly constrained northwest-southeast contractional deformation in the Fork Lagoons beds may
relate to either Benambran or Tabberabberan deformation (Withnall et al., 1995).

Koonenberry Belt

• Cessation of Delamerian contraction, in the late Cambrian to Early Ordovician, resulted in uplift
and erosion to produce local extensional basins, with deposition of shelf, to deltaic to deep-water
sediments of the Kayrunnera, Mutawintji and Warratta groups, respectively (Greenfield et al.,
2010). Warratta Group sediments occur within both the Delamerian and Thomson orogens
(Greenfield et al., 2010).
• Contractional deformation associated with the Benambran Orogeny occurred in the Late
Ordovician to early Silurian (Gilmore et al., 2007). According to Gilmore et al. (2007) deformation
was most intense east of the Koonenberry Fault. These authors record a vergence change across
this fault––from east-vergent in the east, to west-vergent west of the fault.

Bourke-Louth regions

• Possible backarc and arc-related volcanism––the Warraweena Volcanics, e.g., Burton et al.
(2008)––and associated sedimentation, have been recognised in the Bourke (Thomson Orogen)
and nearby Mount Dijou (Lachlan Orogen) regions (northwest NSW) (Watkins, 2007). The ages of

94 Geodynamic Synthesis of the Phanerozoic of eastern Australia


these rocks are uncertain. Burton et al. (2008) suggested an Ordovician age, based on chemical
similarities with the Macquarie Arc. More recently, Glen et al. (2013) have suggested a
Neoproterozoic age though the zircon evidence for this is equivocal and may well represent a
detrital age, i.e., a maximum depositional age. Fraser et al. (2014) report a maximum deposition
age of ca. 438 ± 9 Ma for a siltstone near to the Warraweena Volcanics, though the actual
significance of this age is uncertain. The Warraweena Volcanics and the Mount Dijou samples are
characterised by both calcalkaline to shoshonitic, and alkaline (with OIB-like patterns),
compositions (Watkins, 2007; Burton et al., 2008). The calcalkaline compositions appear to be
similar chemically to the Ordovician Lachlan Macquarie Arc (Burton et al., 2008).

Thomson Orogen

• Felsic volcanism and granite intrusion occurred in the eastern and central Thomson Orogen
(Murray, 1994), and has been dated as Early Ordovician to Middle Silurian (Draper, 2006).
Porphyritic rhyolite, rhyolitic tuff and crystal tuff from Thomson basement beneath the Eromanga
(GSQ Maneroo-1), Drummond (BEA Coreena-1) and Adavale (Carlow-1) basins and have been
dated as Early Ordovician (473 Ma, 478 Ma, and 484 Ma; respectively; Draper, 2006; McKillop et
al., 2007).
• Felsic magmatism occurred in the western Thomson Orogen beneath the Eromanga and Cooper
basins (AMX Toobrac-1; DIO Ella-1; Murray, 1994). Recent age dating (reported in Draper, 2006)
indicates Middle Ordovician (469 Ma) and Middle Silurian (428 Ma) ages for these granites. Similar
Early to Middle Ordovician ages of the volcanics imply that the volcanic and plutonic rocks are
closely related and may be comagmatic during extension and crustal thinning during the
Ordovician (Draper, 2006).
• Bultitude and Cross (in Fergusson and Henderson et al., 2013) found a late Ordovician age (ca.
456 Ma) for the Granite Springs Granite in the Eulo Ridge region, southern Thomson Orogen.

Warburton and Georgina basins

• Following the Delamerian Orogeny and basin rifting and volcanism, shallow shelf, slope and basin
sediments of the Kalladeina Formation (carbonates) and deep water sediments of the Dullingari
Group (shale, siltstone and chert) were deposited along the margins of the Australian craton
(Gatehouse and Cooper, 1986; Boucher, 2001).
• Rift-related, continental/intraplate mafic volcanics (Jena Basalt) and related agglomerate were
extruded in the lower part of the Dullingari Group in the Middle Cambrian (Meixner et al., 1999,
2000; Boucher, 2001).
• Red beds of the Innamincka Formation were deposited in a shallow marine epicontinental sea and
although they have been drilled extensively, their age is unconstrained (Gatehouse, 1986); Sun
and Gravestock (2001) suggested they were no older than Early Ordovician.
• Sediments underwent folding and erosion during the Early Silurian Benambran/Alice Springs
Orogeny (1)(Gatehouse, 1986). Cambrian and Late Ordovician rocks were probably
metamorphosed in the Silurian (Draper, 2006), although further work is necessary to determine the
precise age.
• Sedimentation in the southern Georgina Basin throughout the latest Cambrian to early Silurian was
dominated by carbonates and carbonate-bearing sequences of the Cockroach Group. More
clastic-dominated sequences of the Toko Group appear in the Middle and Late Ordovician (e.g.,
Draper, in Fergusson and Henderson et al., 2013).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 95


2.4.5 Middle Silurian to late Devonian (ca. 430 Ma to 380 Ma)

2.4.5.1 Geological and tectonic summary

During this period, both terrestrial and marine sedimentation and associated extrusive and intrusive
magmatism occurred within two episodes, largely in response to extension following the Benambran
and Bindian orogenies (e.g., Olgers, 1972; Neef and Bottrill, 1991; Murray, 1994, 1997a; Withnall et
al., 1995; McKillop et al., 2007; Gilmore et al., 2007; Greenfield et al., 2010; Glen et al., 2013;
Figure 2.19, Figure 2.20). Post-Benambran, Late Silurian to Early Devonian deposition and
magmatism is largely confined to the northern and southern parts of the Thomson Orogen and is
much more extensive in the adjoining Lachlan and Mossman orogens. These rocks are recorded in
the Koonenberry region and Darling Basin (Neef and Bottrill, 1991; Cooney and Mantaring, 2007), and
in the southern Thomson Orogen (Glen et al., 2013). They consist of marine and terrestrial sediments,
of cratonic and/or volcanic provenance, and are associated with mafic and felsic volcanic rocks (Neef
and Bottrill, 1991; Glen et al., 2013). Felsic intrusives of this age are relatively abundant in the
southern Thomson (Greenfield et al., 2010; Bultitude and Cross, in Fergusson and Henderson et al.,
2013; Glen et al., 2013) again similar to the neighbouring Lachlan Orogen (e.g., Ickert and Williams,
2011). They also occur in the Charters Towers and Greenvale provinces and neighbouring
Georgetown region (e.g., Hutton et al., 1997) where they form part of the extensive Silurian–Devonian
Pama Igneous Association of the Mossman Orogen (Bain and Draper, 1997; Bultitude et al., in
Henderson et al., 2013; Figure 2.19, Figure 2.20).

Local folding and metamorphism of these successions, probably with associated felsic magmatism,
took place during the Bindian Orogeny (e.g., Thalhammer et al., 1998). The orogeny may have been
diachronous. Deformation in the Koonenberry region is recorded as Late Silurian to Early Devonian
(Gilmore et al., 2007; Greenfield et al., 2010; Cooney and Mantaring, 2007), while it appears to be
Early Devonian in the Thomson Orogen, where it is constrained to be older than ca. 408 Ma and 402
Ma volcanic rocks (Gumbardo Formation) in the post-Bindian Adavale Basin (McKillop et al., 2007).

Renewed extension, following the Bindian Orogeny, produced widespread Early to Middle Devonian,
terrestrial to shallow marine sedimentation and volcanism, in the Adavale Basin in the central
Thomson Orogen (McKillop et al., 2007), Burdekin Basin (Charters Towers region; Hutton et al.,
1997), the Ukalunda Shelf (Anakie region; Henderson et al., in Fergusson and Henderson et al.,
2013), and in the Koonenberry region and southern Thomson Orogen associated with the Darling
Basin (Neef, 2004, Cooney and Mantaring, 2007). Sedimentation of this age also occurs within the
Georgina Basin (Draper, in Fergusson and Henderson et al., 2013). The Adavale Basin is thought to
have formed in a continental setting, either as an intracontinental volcanic rift or an extensional basin
(Murray, 1994; McKillop et al., 2007). Felsic intrusive magmatism accompanied extension in the
Thomson Orogen (e.g., Murray, 1994; Evans et al., 1990; Hutton et al., 1997; Figure 2.19,
Figure 2.20). McKillop et al. (2007) suggest that extension may have been the result of more regional
events such as subduction further to the east in the New England Orogen. A similar scenario has been
invoked by Henderson et al. (in Fergusson and Henderson et al., 2013) for the Ukalunda Shelf
sediments. The Middle Devonian Tabberabberan Orogeny (which has been called the Alice Springs
Orogeny (2) in the Thomson Orogen, e.g., McKillop et al., 2007) is best recorded in the Koonenberry
region and Darling Basin, in general (Cooney and Mantaring, 2007), where it resulted in east-
northeast–west-southwest contractional deformation at ca. 395 Ma (Mills and David, 2004; Neef,
2004). Deformation of this age in the Thomson Orogen is either poorly developed or often difficult to
distinguish from other events (Withnall et al., 1995; Hutton et al., 1997; Fergusson and Henderson et
al., 2013). In the Adavale Basin, the orogeny appears to have resulted in a major hiatus in the late
Middle Devonian and a possible subsequent change to restricted basin conditions (McKillop et al.,
2007). McKillop (in Fergusson and Henderson et al., 2013) equated this event to similar deformation

96 Geodynamic Synthesis of the Phanerozoic of eastern Australia


events elsewhere in the region, e.g., in the Burdekin Basin and Broken River Province in the northern
Thomson Orogen. Notably, Fergusson and Henderson et al. (2013; their Figure 3.126) suggest that
timing of the Tabberabberan Orogeny in the Adavale and Burdekin basins may have been younger,
around 370 Ma, i.e., at the top of the Adavale Basin, and top of the Dotswood Group in the Burdekin
Basin. In the Ukalunda Shelf (Anakie Province), however, they show that the Tabberabberan
deformation is constrained to be older than the overlying Drummond Basin rocks, i.e., predates the ca.
370 Ma mid-Early Famennian (Paleontological) age for the Greybank Volcanics (Mawson and Talent,
2003), although U-Pb zircon ages for volcanic rocks of the lower parts of the basin are younger (ca.
356 Ma to 344 Ma; see Henderson et al., in Fergusson and Henderson et al., 2013). We have followed
McKillop (in Fergusson and Henderson et al., 2013) and have placed the Tabberabberan Orogeny in
the late Middle Devonian.

2.4.5.2 Geological history and Time-Space plot explanation (Figure 2.19, Figure 2.20)

Anakie region and Ukalunda Shelf

• Early to Middle Devonian carbonates and siliciclastic sediments, with associated mafic,
intermediate and felsic volcanic rocks occur as small remnants in the Anakie Province and also as
(uplifted) isolated occurrences associated with younger rock sequences. These have all been
grouped as the Ukalunda Shelf by Henderson et al. (in Fergusson and Henderson et al., 2013)
who suggested that they probably once formed a more continuous sequence deposited on top of
metasedimentary rocks in the Anakie Province, and granites and volcanics of the central Thomson
Orogen basement (Olgers, 1972; Grimes et al., 1986; Murray, 1994; Figure 2.19, Figure 2.20).
Rocks include the Douglas Creek Limestone, Glendarriwell beds (Olgers, 1972; Withnall et al.,
1995), Dunstable Volcanics and Sedgeford Formation (Henderson et al., in Fergusson and
Henderson et al., 2013) and the Ukalunda beds (Withnall et al., 1995). The Ukalunda beds appear
to form basement to (part of) the Drummond Basin; the two units are separated by an angular
unconformity (Grimes et al., 1986; Hutton et al., 1998a; Draper et al., 2004) interpreted as
reflecting the Tabberabberan Orogeny (Fergusson and Henderson et al., 2013). The Theresa
Creek Volcanics, comprising mafic volcanic rocks and associated sedimentary rocks, also occur in
this region. Their timing relative to the Tabberabberan Orogeny in this region is uncertain; they are
lithologically similar to the Middle Devonian Dunstable Volcanics (Withnall et al., 1995) though
dating, ca. 382 Ma (Cross et al., 2015a), suggests they may be slightly younger.

Adavale Basin

• Felsic volcanic and volcaniclastic rocks (acid crystal tuffs, ignimbrite, minor mafic rocks) of the
Gumbardo Formation formed during initial rifting and half graben formation in the late Early
Devonian (McKillop et al., 2007). The volcanics have been dated as late Early Devonian (402 Ma,
408 Ma; Gumbardo-1 and Carlow-1 wells; McKillop et al., 2007), and were deposited in fluvial or
fluvial-lacustrine conditions (McKillop et al., 2007).
• Terrestrial, fluvial-lacustrine and marginal marine sedimentation (Eastwood Formation) continued
until the Late Devonian. A marine transgression produced a phase of shallow marine carbonate
deposition (Log Creek Formation, Bury Limestone) from the Early to Middle Devonian Alice
Springs Orogeny (McKillop et al., 2007).
• The Alice Springs Orogeny (~Tabberabberan Orogeny) is thought to have produced an
unconformity between terrestrial and overlying shallow marine sedimentary rocks, and may have
triggered the onset of restricted basin conditions in the late Middle Devonian (mid Givetian;
McKillop et al., 2007; McKillop, in Fergusson and Henderson et al., 2013). It is also possible that
the Tabberabberan Orogeny may have been slightly younger in this region and affected the whole
basin (see discussion in McKillop, in Fergusson and Henderson et al., 2013, p. 178).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 97


Thomson Orogen

• Middle Silurian and Early Devonian felsic intrusion into Thomson basement rocks. These include
an earlier phase of granites (ca. 428–420 Ma) in the southern Thomson, e.g., the Tibooburra Suite
in the Tibooburra region (Greenfield et al., 2010); in the Hungerford and Eulo regions (Cross et al.,
2015), the Warburton Basin region (Draper, 2006), and the northern and eastern Thomson, e.g.,
Charters Towers and Greenvale regions (Bultitude et al., in Henderson et al., 2013), Anakie region
(Cross et al., 2015). Younger, ca. 415–385 Ma magmatism is more widespread, with occurrences
in the Hungerford and Eulo regions (Bultitude and Cross, in Fergusson and Henderson et al.,
2013; Cross et al., 2015a), Louth-Eumarra shear zone along the Thomson–Lachlan boundary
(Glen et al., 2103), central Thomson (Draper, 2006), Anakie region (Woods and Lister, in
Fergusson and Henderson et al., 2013), and in the northern Thomson (Bultitude et al., in
Henderson et al., 2013).
• Middle Devonian (395–385 Ma) felsic volcanism (rhyolitic ignimbrite) is known from drill core (AAE
Towerhill-1: ca. 385 Ma; Draper, 2006; Thunderbolt 1: ca. 392 Ma; Kositcin et al., 2015a) beneath
the Galilee Basin. The younger age is close to the age of the Tabberabberan Orogeny and has
been suggested to correlate with the Silver Hills Volcanics in the Drummond Basin (Murray, 1994).
If correct, this would suggest the former are post-Tabberabberan Orogeny.

Bowen Basin

• Devonian deep marine (turbidite?) deposition of metasedimentary rocks of the Timburry Hills
Formation (basement to Bowen Basin). The quartz sandstone unit is uniform in character and
suggests a continental source (Murray, 1997a). A Devonian age is constrained by the presence of
plant material and emplacement of the Roma Granites close to the Devonian–Carboniferous
boundary (Murray, 1997a). Kositcin et al. (2015a) report a maximum deposition age of 420 Ma
(latest Silurian) from a sandstone unit, consistent with the inferred Devonian age.

Koonenberry Belt, Darling Basin and southern Thomson

• Andesitic and basaltic volcanics, volcaniclastics and limestones––the Louth Volcanics––were


deposited in the Uppermost Silurian to Lower Devonian in the southern Thomson Orogen (Dadd,
2006; Glen et al., 2013). According to Dadd (2006) the volcanic rocks have a geochemical
signature consistent with either an ocean-island or intracratonic rift origin and distinct from that in
the older Macquarie Arc rocks to the south. Glen et al. (2013) showed that inherited ca 550–400
Ma zircons within this unit were characterised by primitive Hf isotopic signatures, suggesting pre-
existing primitive basement in this region.
• Deposition in the Early Devonian of the nonmarine sediments of the Mount Daubeny Basin derived
from a mixed cratonic and volcanic provenance (Neef and Bottrill, 1991). Both Neef and Botrill
(1991) and Gilmore et al. (2007) record andesitic volcanism (ca. 425 Ma; Black, 2007) and
intrusives within the succession, and Neef and Bottrill (1991) suggested a proximal volcanic source
for the volcaniclastic component of the sediments. The Mount Daubeny Basin has been suggested
to represent an earlier part of the Darling Basin (e.g., Cooney and Mantaring, 2007), with
correlatives also in the northern Lachlan (e.g., Winduck Group) though not including the similar-
aged Cobar Basin. Inferred correlations of both the Darling and Cobar basins continue into the
southern Thomson Orogen, e.g., Middle to Upper Devonian rocks within the Paka Tank Trough
(Glen et al., 2013).
• The Late Silurian to Early Devonian Bindian Orogeny affected the Koonenberry region and largely
resulted in dextral strike-slip movement (Gilmore et al., 2007; Greenfield et al., 2010). Apparently
associated with the deformation are ca. 427 to 420 Ma monzodioritic and I-type granite intrusions
(Black, 2006; Greenfield et al., 2010). These have intruded the Koonenberry region, but are mostly

98 Geodynamic Synthesis of the Phanerozoic of eastern Australia


concentrated in the adjoining Thomson Orogen (Thalhammer et al., 1998). The latter include the
fractionated (I-type) Tibooburra Suite (Greenfield et al., 2010). Thalhammer et al. (1998)
suggested that these granites were emplaced in an intracontinental setting, syntectonic with local
Early Devonian deformation, probably the Bindian Orogeny (Gilmore et al., 2007). High level Late
Silurian–Early Devonian rhyolites were emplaced at ca. 418–414 Ma (Black, 2006; Gilmore et al.,
2007).
• Post-Bindian extension resulted in deposition of the Early and Middle Devonian terrestrial to
shallow marine, quartz-rich sediments of the Wana Karnu Group (Gilmore et al., 2007). This
includes the Snake Cave Sandstone of the Darling Basin, described by Neef and co-workers, e.g.,
Neef (2004). Correlatives, probably of the Mulga Downs Group, also occur in the southern
Thomson Orogen, e.g., Paka Tank Trough (Glen et al., 2013).
• East-northeast–west-southwest contractional deformation of the Tabberabberan Orogeny, ca. 395
Ma (Mills and David, 2004, Neef, 2004; Cooney and Mantaring, 2007).

2.4.6 Late Devonian to early Carboniferous (ca. 380 Ma to ca. 350 Ma)

2.4.6.1 Geological and tectonic summary

During the late Devonian to early Carboniferous terrestrial and marine sedimentation, often with
accompanying volcanism, occurred across the Thomson Orogen and surrounding regions, e.g., the
Darling Basin (Koonenberry and Lachlan regions), upper parts of the Adavale and Burdekin basins, in
the Bundock and Clarke River basins (northern Thomson) and similar aged sedimentation in the
Georgetown and Coen regions, and in the Drummond Basin (Anakie Province) (Figure 2.19,
Figure 2.20). This sedimentation was largely in response to intracratonic extension following the
Tabberabberan Orogeny, but also, as for the Drummond Basin, a result of backarc extension behind a
Late Devonian to early Carboniferous arc in the New England Orogen (e.g., Neef and Bottrill, 1991;
Murray, 1994; Withnall et al., 1995; Henderson et al., 1998; Draper et al., 2004; Gilmore et al., 2007).
The Drummond Basin consists of a thick succession of continental and lesser marine sediments and
volcanics (Olgers, 1972; Hutton et al., 1998a; Henderson et al., 1998; Henderson and Blake, in
Fergusson and Henderson et al., 2013). These have been subdivided into three major
tectonostratigraphic cycles (Figure 2.19), separated by unconformities (e.g., Olgers, 1972). The
lowermost cycle––cycle 1 (latest Devonian to early Carboniferous)––consists of syn-rift related
volcanic rocks and associated marine to terrestrial volcaniclastic sediments (Olgers, 1972; Henderson
et al., 1998). Early Carboniferous Cycle 2 rocks consist of a thick succession of terrestrial (and local
marine) sediments which reflect an abrupt end to volcanism and a switch to a cratonic provenance
(Olgers, 1972). Cycle 3 rocks (also early Carboniferous) reflect a return to volcanism, which continued
episodically, with terrestrial sediments (Olgers, 1972).

Felsic and lesser amounts of intermediate and mafic magmatism accompanied extension episodically
throughout this cycle. This includes intrusive and related extrusive magmatism in the Anakie Province
occurring syn? to post-Tabberabberan Orogeny, but prior to the formation of the Drummond Basin.
During the Late Devonian to early Carboniferous, the Thomson Orogen also experienced regionally
extensive felsic magmatism (e.g., Murray, 1994; Kositcin, 2015a; Figure 2.19, Figure 2.20). Early
Carboniferous granites were emplaced into Thomson Orogen basement and the Warburton Basin
(e.g., Murray, 1994). During initial Late Devonian–Early Carboniferous backarc extension, silicic
magmatism at this time was spread over a broad region in the Drummond Basin and may be related to
episodes of silicic magmatism in the New England Orogen (e.g., Bryan et al., 2004). This magmatism
largely predates the widespread Kennedy Province magmatism in Charters Towers and further north,
although volcanism in cycle 3 of the Drummond Basin appears to correlate with volcanism in the upper

Geodynamic Synthesis of the Phanerozoic of eastern Australia 99


parts of the Burdekin, Bundock and Clarke River basins in the Charters Towers and Broken River
regions of north Queensland (e.g., Henderson et al., 1998; Figure 2.13).

The early to middle Carboniferous Kanimblan Orogeny, or third stage of the Alice Springs Orogeny as
it is often referred to in the Thomson Orogen, produced a major episode of faulting and deformation in
the Koonenberry Belt and in the Darling Basin, slight contraction in the Drummond Basin and regional-
scale folding and subsequent erosion in the Adavale Basin (e.g., Olgers, 1972; Neef, 2004; Cooney
and Mantaring, 2007; Greenfield et al., 2010). This deformation event is suspected to have driven
regional-scale, southward thrusting of the Thomson over the Lachlan Orogen (Korsch et al., 1997).

2.4.6.2 Geological history and Time-Space plot explanation (Figure 2.19, Figure 2.20)

Anakie Province (Anakie region)

• The Middle Devonian (ca. 385–370 Ma) I-type Retreat Batholith intrudes the Anakie Metamorphic
Group (Webb and McDougall, 1968; Olgers, 1972; Withnall et al., 1995), and post-dates regional
metamorphism and probably the Tabberabberan Orogeny.
• Middle Devonian mafic to intermediate volcanics and associated volcaniclastics of the Theresa
Creek Volcanics (Olgers, 1972; Withnall et al., 1995). The volcanics are locally intruded by granites
of the Retreat Batholith and conformably overlie the Douglas Creek Volcanics (Olgers, 1972).
Although they may belong to the previous cycle, Withnall et al. (1995) suggested that the volcanics
were contemporaneous with, and probably genetically related to, intrusive magmatism of the
Retreat Batholith. Cross et al. (2015a) obtained what they described as a maximum depositional
age of ca. 382 Ma for the Theresa Creek Volcanics, consistent with the interpretation of Withnall et
al. (1995).
• Early Late Devonian andesitic volcanism and minor terrestrial to marine sedimentation of the
Greybank Volcanics (Withnall et al., 1995). Withnall et al. (1995) correlated these with the Dee
Volcanics of the New England Orogen.
• Effects of the Kanimblan-Alice Springs Orogeny in the Anakie Province are uncertain. Withnall et
al. (1995) document minor folding and deformation in the Middle and Late Devonian volcanics and
sediments that may be Kanimblan in age. Fenton and Jackson (1989) consider that Carboniferous
deformation produced right-lateral offset and uplift of the Anakie Province as a series of blocks.

Drummond Basin (Anakie region)

• The Drummond Basin was initiated during this cycle, probably in the latest Late Devonian
(Henderson et al., 1998). The basin consists of a thick succession of continental sediments and
volcanics, with minor marine interbeds towards the base of the succession (Olgers, 1972; Hutton
et al., 1998a; Henderson et al., 1998; Figure 2.19, Figure 2.20). The basin unconformably overlies
the Early Devonian Ukalunda beds, the Retreat Batholith, and the older rocks of the Anakie
Province (Olgers, 1972; Withnall et al., 1995). The Drummond Basin has been subdivided into
three major tectonostratigraphic cycles, separated by unconformities (e.g., Olgers, 1972; Hutton et
al., 1998a; Figure 2.19). The lowermost cycle––cycle 1 (latest Devonian to early Carboniferous)––
consists of syn-rift related volcanics rocks (andesitic, dacitic and dominant rhyolitic lava, ignimbrite
and tuff) and associated marine to terrestrial volcaniclastic sediments (Olgers, 1972; Henderson et
al., 1998). Henderson et al. (1998) indicated that volcanism and deposition initiated first in the
north. Cross et al. (2009) report ages of ca. 363 and ca. 360 Ma for the Cycle 1 Silver Hills
Volcanics and Bimurra Volcanics, respectively. Early Carboniferous Cycle 2 rocks consists of a
thick sequence of terrestrial (and local marine) sediments––mostly quartz-rich sandstone,
conglomerate and mudstone––which reflects an abrupt end to volcanism and a switch to a cratonic
provenance (Olgers, 1972). Cycle 3 rocks (also early Carboniferous) reflect a return to volcanism,

100 Geodynamic Synthesis of the Phanerozoic of eastern Australia


reflected by volcaniclastic sediment, tuff and conglomerate (Olgers, 1972). Volcanism continued
episodically throughout this cycle, although the upper parts are dominated by terrestrial sediments
(Olgers, 1972).
• Deposition within the Drummond Basin was terminated by the Kanimblan Orogeny although the
actual effects of the Kanimblan Orogeny are subject to interpretation, varying from being
responsible for uplift and deformation (e.g., Olgers, 1972) through to minimal effect (e.g., Van
Heeswijck, 2004). Deformation appears to record both Kanimblan and the younger Hunter-Bowen
Orogeny (see discussion in Henderson and Blake, in Fergusson and Henderson et al., 2013). The
best evidence for the Kanimblan Orogeny relates to unconformities between the Drummond and
overlying sequences, e.g., Bowen and Galilee basins (Henderson et al., 1998). It is also possible
that some of the observed deformation may have been produced by the Tabberabberan Orogeny,
especially if the latter orogeny was slightly younger in this region and affected the whole basin (see
discussion by McKillop, in Fergusson and Henderson et al., 2013, p. 178), although the
observations of Van Heeswijck (2004) would suggest otherwise.

Thomson Orogen

• Late Middle Devonian (385 Ma) felsic volcanism (rhyolitic ignimbrite; AAE Towerhill-1; Draper,
2006), which has been suggested to correlate with the Silver Hills Volcanics in the Drummond
Basin (Murray, 1994). Kositcin et al. (2015a) obtained a slightly older age (ca. 393 Ma) for a
rhyolitic ignimbrite (from the Thunderbolt 1 drill hole) 80 km south-southeast of the Draper (2006)
sample.
• Widespread Late Devonian to early Carboniferous ca. 370–355 Ma) intrusion of S-type, and more
localised I-type, Roma granites into the Timburry Hills Formation (Murray, 1994; Kositcin et al.,
2015a) in the eastern Thomson Orogen. Mineralogy suggests the granites were intruded into old,
stable continental crust (Murray, 1994). Recent geochronology indicates magmatic ages of 368 Ma
(Kositcin et al., 2015a) from an interpreted S-type granite (Murray, 1994), on the western side of
the Roma Shelf, and 363 Ma (Cross et al., 2015b), from a granite, also interpreted as S-type, on
the eastern side.

Adavale Basin

• Post-Tabberabberan deposition in the Adavale Basin includes latest Middle Devonian to late
Devonian fluvial to marginal marine quartzose sediments and minor limestone and overlying late
Devonian terrestrial red bed rocks (e.g., McKillop et al., 2007; McKillop, in Fergusson and
Henderson et al., 2013). These rocks unconformably overlie Middle Devonian marine sequences.
The unconformity is marked by a hiatus with related widespread aridity and formation of halite
deposits (McKillop et al., 2007).
• Termination of deposition and basin deformation occurred during the Alice Springs Orogeny, with
development of regional-scale folds followed by widespread erosion (McKillop et al., 2007).

Koonenberry Belt and Darling Basin

• Post-Tabberabberan extension resulted in deposition of terrestrial quartz-rich sediments (e.g., the


Ravendale Formation), as part of the Darling Basin (e.g., Neef, 2004; Cooney and Mantaring,
2007).
• These rocks were deformed as part of the Kanimblan Orogeny, which produced regional faulting,
fault reactivation and transpressive deformation (Neef, 2004; Gilmore et al., 2007).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 101


2.4.7 Early Carboniferous to late Permian (ca. 350 Ma to ca. 265 Ma)

2.4.7.1 Geological and tectonic summary

From the late Carboniferous to early Late Permian, the Thomson region was dominated by formation
of a number of large basins (Cooper, Galilee, Bowen basins; Figure 2.19, Figure 2.20). Continental
margin extension in the (latest Carboniferous–) early Permian led to formation of the Bowen Basin in a
backarc setting west of the Camboon Volcanic Arc in the New England Orogen (e.g., Korsch et al.,
2009a; see discussion of geodynamic models for the Bowen Basin by Draper, in Donchak et al.,
2013). By the early Permian, the Bowen Basin, together with the Gunnedah and Sydney basins,
formed the ‘East Australian Rift System’ (Korsch et al., 1998, 2009a). This system went through a
period of extension and deposition of fluvial and lacustrine sediments and volcanics––Supersequence
A of Brakel et al. (2009) and Totterdell et al. (2009)––prior to a short lived period of thermal relaxation
and deposition of marine, deltaic and fluvial sediments––Supersequence B of Brakel et al. (2009) and
Totterdell et al. (2009). The latter was ended by a change to a foreland basin environment, marked by
the Aldebaran Event, a significant mid-Permian deformation event that signifies the start of the Hunter-
Bowen Orogeny (e.g., Korsch et al., 2009a).

Although early Permian extension was located on the continental margin, intracratonic basins such as
the Galilee and Cooper formed on the craton further west at this time (Gravestock and Jensen-
Schmidt, 1998; Gray and McKellar, 2002; Draper and McKellar, 2002; Korsch et al., 1998
Figure 2.20).These basins are also commonly suggested to have been initiated by tectonic extension
plus or minus rifting (e.g., Jackson et al., 1981; Middleton and Hunt, 1989; Evans et al., 1990; Van
Heeswijck, 2004; although see discussions by Gravestock and Jensen-Schmidt (1998), and by
McKellar, in Fergusson and Henderson et al. (2013) for the wide range of models advocated). Relative
tectonic stability in west and central Queensland during the late Carboniferous to Permian led to
widespread terrestrial sedimentation in both the Cooper and Galilee basins (e.g., Draper, 2002a, b).
Both basins experienced similar sediment deposition (Scott et al., 1995). Deposition ended in the
Galilee and Cooper basins at much the same time as in the Bowen Basin, reflecting onset of the
foreland phase in the latter basin (Draper, in Donchak et al., 2013).

Widespread late Carboniferous to early Permian igneous activity of the Kennedy Igneous Association
occurs in the Charters Towers and Cape River regions including the northern Anakie and Drummond
regions (Champion and Bultitude, 2013a; Figure 2.13, Figure 2.20). Magmatism of this age is also
present further south and west in the Anakie Province and in the Warburton Basin (e.g., Gatehouse et
al., 1995; Hutton et al., 1998a; Denaro et al., 2004; Sliwa and Draper, 2005). This activity is the same
age as extensive arc- and backarc-related magmatism in the Bowen Basin and in the New England
Orogen to the east (e.g., Donchak et al., 2013) although there are distinct differences in the nature of
magmatism between the two regions, reflecting the cratonic versus arc-related environments
(Champion and Bultitude, 2013a).

2.4.7.2 Geological history and Time-Space plot explanation (Figure 2.19, Figure 2.20)

Drummond Basin and northern Anakie Province (Anakie region)

• Eruption of the dominantly felsic Bulgonunna Volcanic Group in the late Carboniferous to early
Permian(?) (ca. 305–292 Ma; Black, 1994; Hutton et al., 1998; Cross et al., 2012). It is dominated
by felsic extrusives including rhyolite, ignimbrite and minor related sediments (Olgers et al., 1972;
Hutton et al., 1998a). The group unconformably overlies the Mount Wyatt Formation (northern
Drummond Basin; Olgers et al., 1972). The group forms part of the Burdekin Falls Subprovince of
the Kennedy Igneous Association (Champion and Bultitude, 2013a).

102 Geodynamic Synthesis of the Phanerozoic of eastern Australia


• Late Carboniferous comagmatic, and Permian multi-phase intrusions, smaller plutons and dykes
occur on the margins of the Bulgonunna Volcanic Group, and have also intruded the Drummond
Basin and Anakie Province (Olgers et al., 1972; Hutton et al., 1998a). Intrusive magmatism was, in
part, comagmatic with silicic volcanics of the Bulgonunna Volcanic Group (Oversby et al., 1994;
Hutton et al., 1998a). Recorded ages range from ca. 308 Ma to ca. 287 Ma (Black, 1994; Hutton et
al., 1998a; Cross et al., 2012).

Bowen and Gunnedah basins

• Late Carboniferous to early Permian fluvial and lacustrine sedimentation (e.g., Reids Dome beds)
in half-grabens on the western side of the Bowen Basin (e.g., Brakel et al., 2009) and basaltic,
andesitic to rhyolitic volcanics and fluvial sediments (e.g., Combarngo Volcanics in the west,
Camboon Volcanics, Lizzie Creek Volcanic Group in the east) concentrated on the eastern side of
the Bowen Basin (e.g., Green et al., 1997b; Brakel et al., 2009; Korsch et al., 2009a). Similar rocks
occur in the Gunnedah and southernmost Bowen basins (e.g., Boggabri Volcanics, Goonbri
Formation, e.g., Tadros, 1995; Totterdell et al., 2009). Voluminous volcanic rocks are most
abundant in the eastern parts of the basins and may also be present at depth. Gravity modelling of
the ~1200 km Meandarra gravity ridge, which runs through the Gunnedah and southern Bowen
basins, suggests the ridge comprises thick successions (up to 9 km) of mafic volcanic rocks
(Krassy et al., 2009) interpreted as being rift-related. Ages for the Camboon Volcanics and Lizzie
Creek Volcanic Group range from ca. 297–277 Ma (see compilation by Draper, in Donchak et al.,
2013). Ages for the Combarngo Volcanics are not well constrained but are probably late
Carboniferous to early Permian (Draper, in Donchak et al., 2013). These rocks form
Supersequence A of Brakel et al. (2009) and Totterdell et al. (2009), which they suggested was
related to lithospheric extension, initial rifting and formation of the basin.
• Following mild basin inversion (Cattle Creek Event of Korsch et al., 2009b; Brakel et al., 2009), the
Bowen and Gunnedah basins entered a period of deposition of marine, deltaic and fluviatile
sediments including the formation of the Collinsville Coal Measures (e.g., Draper, in Donchak et
al., 2013; Hamilton, 1993; Tadros, 1995; Shaw, 2002). These form Supersequence 2 of Brakel et
al. (2009) and Totterdell et al. (2009). They include the lower part of the Back Creek Group and are
commonly related to a thermal subsidence phase following initial extension (e.g., Korsch et al.,
2009a). Sedimentation ended with mid-Permian deformation––the Aldebaran Event of Korsch et
al. (2009b) which these authors suggested was related to the end of thermal subsidence and
initiation of a foreland basin environment. This event produced the mid-Aldebaran unconformity
(e.g., Stephens et al., 1996; Korsch et al., 1998; Draper, in Donchak et al., 2013) which Draper
suggested was contemporaneous with the post-Daralingie unconformity in the Cooper Basin and
the mid-Permian hiatus in the Galilee Basin.

Cooper Basin

• Early to Late Permian sediments of the Gidgealpa Group were deposited in glacial, fluvial and
lacustrine environments with the Patchawarra Formation recording the waning stages of early
Permian glaciation (Gravestock and Jensen-Schmidt, 1998; Gray and McKellar, 2002).
• Numerous, minor contractional events occurred during deposition of the Gidgealpa Group, and are
thought to reflect episodes of the Hunter–Bowen Orogeny (e.g., Apak et al., 1997; Gray and
McKellar, 2002). Uplift in the late Early Permian resulted in the formation of a significant hiatus and
unconformity––the post-Daralingie unconformity (e.g., Gravestock and Jensen-Schmidt, 1998).

Galilee and Warburton basins

Geodynamic Synthesis of the Phanerozoic of eastern Australia 103


• Late Carboniferous to early Permian terrestrial (fluvial) sediment deposition of the glacial Joe Joe
Group including the early Permian Aramac Coal Measures (Scott et al., 1995).
• Deposition of the coal measures preceded a significant hiatus in the middle Permian that is much
more significant than that expressed in the neighbouring Cooper and Bowen basins (e.g., McKellar
and Henderson, in Fergusson and Henderson et al., 2013). These authors suggested this hiatus
was due to early Hunter-Bowen Orogeny deformation.
• Middle Carboniferous (323 ± 5 Ma) and possibly earliest Permian (298 ± 4 Ma) felsic granite
intrusion (Big Lake Suite) into the Warburton Basin (dated from Moomba-1 and McLeod-1;
Gatehouse et al., 1995). Emplacement could potentially be syntectonic with the Alice Springs
Orogeny of Central Australia.

2.4.8 Late Permian to late Triassic (ca. 265 Ma to 230 Ma)

2.4.8.1 Geological and tectonic summary

In general, tectonic stability continued throughout the Permian and into the Triassic and led to the
widespread infilling of intracratonic basins. Fluvial and lacustrine systems were associated with
extensive swamps in the Cooper and Galilee basins, which resulted in the continuous deposition of
plant-rich material suitable for coal generation (Cowley, 2007; Figure 2.19, Figure 2.20). In the late
Permian, coastal swamps formed in the subsiding Bowen Basin leading to an accumulation of
extensive coal deposits (Shaw, 2002; Draper, in Donchak et al., 2013). Sedimentation in the Bowen,
Cooper and Galilee basins continued throughout the Permian and into the Triassic, until the Middle
Triassic (Scott et al., 1995; Green et al., 1997b; Draper, 2002a, b; Korsch et al., 2009a, b).

During this time period, sedimentary basins were variably affected by the Hunter-Bowen Orogeny
(e.g., Harrington and Korsch, 1985; Korsch et al., 2009a,b), which is proposed to have extended from
the middle of the Permian to the Middle or Late Triassic (ca. ~265 Ma to ~230 Ma). The New England
Orogen was thrust westward during this event, which resulted in tectonic loading and subsidence in
the Bowen and Gunnedah basins, i.e., a switch from extension and thermal relaxation stages to a
foreland setting stage which continued until the late Middle Triassic (Korsch et al., 2009a, b). In the
mid-Permian, large volumes of volcaniclastic material were shed from the volcanic arc and deposited
in the adjoining Bowen Basin during foreland loading (Green et al., 1997a). This switch to a foreland
setting is also marked by a well-developed mid Permian unconformity, not just in the Bowen and
Gunnedah basins, but also in the Galilee and Cooper basins (e.g., e.g., Gravestock and Jensen-
Schmidt, 1998; McKellar and Henderson, in Fergusson and Henderson et al., 2013). This
unconformity marks a change in both sedimentation style and tectonic regime (e.g., Draper, in
Donchak et al., 2013).

The Hunter-Bowen Orogeny ended in the Middle to Late Triassic and is marked by the Goondiwindi
Event of Korsch et al. (2009a, b). This contractional event resulted in uplift and erosion, and the
cessation of sediment deposition in the Galilee, Cooper and Bowen basins (Apak et al., 1997; Korsch
et al., 1998). This widespread event was felt across eastern Australia and resulted in folding and uplift
of parts of the Drummond Basin (Fenton and Jackson, 1989; Van Heeswijck, 2004).

In the west, restricted magmatic activity occurred in the Cooper Basin and the Bourke-Louth region.
Burton et al. (2007) suspect that the Midway Granite and coeval intrusives, east of Bourke, indicate a
more spatially widespread Middle Triassic magmatic pulse than is currently recognised.

104 Geodynamic Synthesis of the Phanerozoic of eastern Australia


2.4.8.2 Geological history and Time-Space plot explanation (Figure 2.19, Figure 2.20)

Drummond Basin (Anakie region)

• Regional Middle Triassic (255–230 Ma) east-west contraction resulted in folding, thrusting, sinistral
strike-slip movement and erosion (Olgers, 1972; Murray, 1990; Johnson and Henderson, 1991;
Van Heeswijck, 2004).

Cooper Basin

• Deposition of the fluvial to deltaic late Permian coal-bearing Toolachee Formation above the post-
Daralingie unconformity (e.g., Gravestock and Jensen-Schmidt, 1998). This was followed
conformably by the latest Permian to Middle Triassic deposition of the fluvial to lacustrine
Nappameri Group (Gray and McKellar, 2002).
• Major Late Triassic contraction produced widespread uplift and erosion following deposition of the
Nappameri Group, and led to the cessation of sediment deposition (Apak et al., 1997; Korsch et
al., 1998) and formation of a basin-wide erosional unconformity (Gravestock and Jensen-Schmidt,
1998).
• Late Triassic or Early Jurassic mafic volcanism in the southwestern part of the basin (Nappamerri
Trough) produced olivine basalts with ages of 227 ± 3 Ma and 100 ± 9 Ma (Murray, 1994; Draper,
2002a, b).

Galilee Basin

• Late Permian to Middle Triassic terrestrial fluvial and lacustrine sediment and coal deposition (e.g.,
Colinlea Sandstone, Bandanna Formation, Betts Creek beds, Rewan Group, Clematis Group,
Moolayember Formation; Evans, 1980; Scott et al., 1995).
• Deposition ended when the basin was affected by the Hunter-Bowen Orogeny causing uplift and
tilting in the west and deformation in the east (Evans, 1980; Van Heeswijck, 2004). The upper unit–
–the Moolayember Formation––has an angular unconformity with the overlying Jurassic–
Cretaceous Eromanga Basin as a result of this deformation (Van Heeswijck, 2004).

Bowen and Gunnedah basins

• Following the early deformation of the Hunter-Bowen Orogeny––the Aldebaran Event of Korsch et
al. (2009b)––deposition in the Bowen and Gunnedah basins comprised fluvial to deltaic to shallow
marine sedimentation, prior to a number of transgression and progradation phases (including the
appearance of tuffs, e.g., Draper, in Donchak et al., 2013) culminating in widespread coal
deposition in the Late Permian (e.g., Rangal Coal Measures; Shaw, 2002; Brakel et al., 2009;
Draper, in Donchak et al., 2013). These sequences which in the Bowen Basin comprise the upper
part of the Back Creek Group, correspond to Supersequences C, D, E and F of Brakel et al. (2009)
and Totterdell et al. (2009).
• Subsequent sedimentation in the Early Triassic saw a switch to the dominantly fluvial, more
oxidising conditions of the Rewan Group and the Gunnedah Basin equivalent, the Digby Formation
(Green et al., 1997a; Shaw, 2002; Draper, in Donchak et al., 2013), corresponding to
Supersequence G of Brakel et al. (2009) and Totterdell et al. (2009). The large quantities of
volcanolithic material suggest detritus was largely sourced from the New England Orogen to the
east (Green et al., 1997a; Shaw, 2002; Brakel et al., 2009).
• Fluvial and lacustrine sedimentation continued until the end of the Middle Triassic (Clematis
Group, Moolayember Formation). The abundance of quartz sandstone in the Clematis Group
reflects a change in sedimentary source from the eastern arc to the uplifted craton in the west

Geodynamic Synthesis of the Phanerozoic of eastern Australia 105


(Fielding et al., 1990; Brakel et al., 2009). Volcanic sediments are present in the sequence and
most likely originated from a volcanic arc source located in the east (Green et al., 1997a).
• Deposition in the Bowen and Gunnedah basins through this time period reflects a foreland basin
setting (e.g., Korsch et al., 2009a, b), with a significant number of contractional deformation events
recognised––from the late Permian Baralaba Event, followed by the Bellata, Brumby, Clematis,
Showgrounds, and Middle–Late Triassic Goondiwindi events (Korsch et al., 2009b), all related to
the Hunter-Bowen Orogeny. Deposition ended with the Goondiwindi Event, prior to deposition of
the overlying Surat Basin.

Southern Thomson

• In the Middle Triassic, magmatism produced highly fractionated I-type felsic intrusions (Midway
Granite) and comagmatic quartz-feldspar porphyry dykes in the Bourke region. The Midway
Granite has been dated at 235 ± 1.4 Ma and is associated with tin and other base metal
mineralisation (e.g., skarn-type Doradilla prospect; Burton et al., 2007).

106 Geodynamic Synthesis of the Phanerozoic of eastern Australia


3 Regional overview of the tectonic development of
eastern Australia in the Phanerozoic

DC Champion, modified from DC Champion and N Kositcin (2009)

3.1 Introduction
The geology and tectonic development of eastern Australia, particularly the Phanerozoic component––
the Tasman Orogen (Scheibner and Veevers, 2000; Veevers, 2000, 2004; Cawood, 2005; Glen,
2005)––has been the focus of numerous studies, with a voluminous literature, including numerous
orogen-based or more regional reviews (e.g., Murray, 1986; 1997b; Murray et al., 1987; Coney, 1992;
Seymour and Calver, 1995; Bain and Draper, 1997; Gray et al., 1997, 2003; Gray, 1997; Gray and
Foster, 1997; 2004; Scheibner and Basden, 1998; VandenBerg et al., 2000; Veevers, 2000; Li and
Powell, 2001; Crawford et al., 2003a; Glen, 2005, 2013; Cawood, 2005; Fergusson, 2010; Greenfield
et al., 2011; Cayley, 2011; Gibson et al., 2011, 2015; Henderson et al., 2011; Offler and Murray, 2011;
Glen et al., 2013; Moresi et al., 2014). The focus of this research has led to a plethora of tectonic
models with perhaps the majority of differences focussed on the Lachlan Orogen (e.g., see Gray,
1997; Gray and Foster, 2004; Figure 3.1 to Figure 3.4). Importantly, despite the differences, there is a
general consensus that since the late Neoproterozoic, eastern Australia has, broadly, been in three
fundamental tectonic states:

• Rodinian-breakup (rifting) and ensuing passive margin––in its simplest form, basically formation of
the Pacific Ocean; also corresponds broadly to development of the Australian component of the
Gondwana margin (e.g., Li and Powell, 2001; Direen and Crawford, 2003; Cawood, 2005).
• Alternating extension and compression in an overall convergent arc environment––the Tasman
Orogen–-commencing in the early Cambrian and continuing through to the Mesozoic, resulting in
accretionary growth––as evidenced in the Lachlan, Thomson, Mossman and New England
orogens. Orogenic events include the Cambrian Delamerian through to the Permian-Triassic
Hunter-Bowen orogenies. The Tasman orogen was effectively terminated at ca. 230 Ma with the
stepping out of the plate margin to the east following the Hunter-Bowen Orogeny and ensuing
extension leading to subsequent breakup of Gondwana–Pangaea.
• Rifting and passive margin (± hotspot activity), related to rifting of crustal fragments, and opening
of ocean basins, especially related to Gondwana breakup.

Although broadly simple, in detail the tectonic evolution of eastern Australia is clearly complex. Not
only is there evidence for diachronous events (Gray et al., 2003), and possible arc switches (e.g.,
Crawford and Berry, 1992; Murray, 2007; Offler and Murray, 2011; Gibson et al., 2015), it is also clear
that the current make-up of eastern Australian provinces (especially Paleozoic to early Mesozoic) may
represent an amalgamation of terranes that were not necessarily juxtaposed, that is, there are both
allochthonous and autochthonous terranes (e.g., Cayley, 2011). Controversy and uncertainty involves
the original positions of potentially allochthonous blocks such as western Tasmania and the Selwyn
Block (Cayley et al., 2002; Cayley, 2011; Gibson et al., 2011), the role of strike-slip or other lateral
movement (Willman et al., 2002; Cayley, 2011), including orocline formation (e.g., Rosenbaum et al.,
2013; Moresi et al., 2014), the presence of domains such as the Melbourne Zone which is missing
evidence for major deformation events, as well as the nature of possible oceanic arc remnants

Geodynamic Synthesis of the Phanerozoic of eastern Australia 107


(Macquarie Arc; Gamilaroi-Calliope, Jamieson?) and whether they represent island arcs or not (e.g.,
Offler and Murray, 2011; Glen, 2013; Quinn et al., 2014). Another area of controversy concerns the
actual positions of, and number of, arcs (if any), as best exemplified in the Ordovician and Silurian of
the Lachlan Orogen (e.g., Wyborn, 1992; Gray, 1997; Soesoo et al., 1997; Collins and Hobbs, 2001;
Willman et al., 2002; Fergusson, 2003; Spaggiari et al., 2004; Figure 3.3, Figure 3.4), but also in the
New England Orogen (Murray, 2007; Offler and Murray, 2011; Glen, 2013). More mundane but
equally important controversies concern the actual location of arcs and discriminating between arc-
forearc and backarc environments. Perhaps the best example of this is the ongoing debate regarding
the interpretation of the sedimentary rocks in the Hodgkinson Province (e.g., Henderson, 1980;
Bultitude et al., 1993; 1997; Henderson et al., 2013).

Much debate regarding tectonic reconstructions for eastern Australia centres around the Lachlan
Orogen, which although probably an accretionary margin (although cf. Glen, 2013), has, as
summarised by Gray (1997), numerous features including its width, the presence of interpreted
oceanic arc remnants (Macquarie Arc) within the Ordovician turbidites, and variable deformation, that
are not easy to explain by simple models (e.g., see Collins and Vernon, 1994; Moresi et al., 2014).
Further uncertainty concerns the Thomson Orogen, particularly the nature and age of the basement.
The orogen is largely undercover, and as such poorly understood. Recent seismic results across the
southern margin of this orogen (Glen et al., 2013), suggest that this boundary is not simple: the MOHO
thickens to the north and the boundary may represent collision between the Lachlan and Thomson
orogens (although Burton (2010) presents an alternative view). Glen et al. (2007c) suggested that any
collision probably predated the Late Devonian, as sedimentary rocks of this age are found in both the
Thomson and Lachlan orogens. Recent geochronology (e.g., Draper, 2006; Bultitude and Cross, in
Fergusson and Henderson et al., 2013; Glen et al., 2013) have pushed ages of granites and volcanics
back in time to the Early Ordovician, i.e., slightly younger than the age of the Delamerian Orogeny,
raising the possibility of Delamerian or older crust at least within parts of the orogen (e.g., Ferguson et
al., 2007c). The possibility of such a scenario raises questions about the Thomson Orogen as a whole,
and how it relates to northern Australia, especially the ‘Tasman Line’ and the geological interpretation
of the latter. It is also evident that there are significant differences between the southern and northern
parts of the Tasman Orogen, and so models for the Lachlan and Thomson orogens also have to be
able to explain the geology of the north Queensland region. In the latter region it would appear that
successive continental margins have largely developed on top of, or close to, each other throughout
the Paleozoic, and where the New England Orogen is only some few hundred kilometres from Paleo-
to Mesoproterozoic crust.

In the following discussions possible geodynamic environments are discussed by time periods, from
the late Neoproterozoic (prior to the Delamerian Orogeny) through to the Mesozoic and the terminal
orogeny––the Hunter-Bowen Orogeny–-in the New England Orogen. Key points for each time period
(summarised from Sections 2.1 to 2.4) are discussed along with the range of geodynamic models that
have been suggested in the literature. We also present one possible tectonic interpretation for each
period; major problems and difficulties with these models and areas of uncertainty are also discussed.

108 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.1 Tectonic evolution model of eastern Australia. Figure modified from Gray and Foster (2004).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 109


Figure 3.2 Interpreted tectonic environments for eastern Australia in the Paleozoic, illustrating the cyclic
alternation of extension and shortening. Figure based on and modified from Collins and Richards (2008).

110 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.3 Schematic tectonic reconstructions for the Lachlan Orogen for the Ordovician to Devonian period.
Tectonic model of VandenBerg et al. (2000) for the Western Lachlan Orogen in Victoria (Whitelaw Terrane). The
reconstruction incorporates the Selwyn Block and the Baragwanath Transform, based on the models presented in
VandenBerg et al. (2000), Cayley et al. (2002), and Willman et al. (2002). Figure based on (and modified from)
from VandenBerg et al. (2000).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 111


Figure 3.4 Schematic tectonic reconstructions for the Lachlan Orogen for the Ordovician to Devonian period.
Tectonic model of Gray (1997) for the Western, Central and Eastern Lachlan Orogen, featuring the multiple
subduction model presented by Gray and co-workers (e.g., Gray, 1997; Soesoo et al., 1997; Foster and Gray,
2004; Spaggiari et al., 2003). Figure based on (and modified from) Gray (1997).

112 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.5. The Tasmanide accretionary model of Moresi et al. (2014). The figure illustrates the indentation of eastern Australia (as it was in the early Paleozoic) with a
microcontinent block (the VanDieland block of Cayley, 2011). Indentation resulted in orocline formation and significant lateral disruption, rotating parts of the orogen and
dismembering and displacing other parts (e.g., the Macquarie Arc). Figure adapted from Moresi et al. (2014). The current position of the interpreted Lachlan–Thomson orogen
boundary (Glen et al., 2013) is shown on all figures for reference, to emphasise the amount of lateral north to south transport of the Tasman Orogen from the beginning of
indentation in the Ordovician to re-establishment of a ‘simple’ arc-subduction geometry in the Early Devonian.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 113


3.2 Tectonic summary of eastern Australia by time period

3.2.1 Late Neoproterozoic to early Cambrian (ca. 600 Ma to ca. 515 Ma):
Rodinian breakup
The Late Neoproterozoic (ca. 600 Ma) to mid-Cambrian geological history of southeastern Australia
records episodic glacial and marine sedimentation, thought to be related to global glaciation (e.g.,
Hoffman and Schrag, 2002), as well as a cycle of continental rifting and ocean opening. The latter was
related to the breakup of Rodinia, and ensuing formation of passive margins and initiation of the
Pacific Ocean (e.g., Li and Powell, 2001; Cawood, 2005; Li et al., 2008; Figure 3.6, Figure 3.7). This
passive margin setting continued in eastern Australia until it was effectively ended by subduction
(starting at least by ca. 515 Ma in the southern Delamerian; Foden et al., 2006). Parts of the southern
Tasman Orogen record arc-continent collision around this time, ca. 510–505 Ma (Crawford and Berry,
1992), though it is uncertain whether this continued into mainland Australia (e.g., Cayley, 2011;
Gibson et al., 2015). Glen (2005) called this interval the Delamerian Cycle, and suggested that it
lasted more than 300 Ma, commencing ca. 830–780 Ma (see also Li et al., 2008). Most of this period
falls outside the time range of this report and only the late Neoproterozoic onwards is covered here in
any detail (see Drexel and Preiss, 1995; Calver and Walter, 2000; Crawford et al., 2003a; Glen, 2005;
and references therein for more information).

Marine sedimentary successions of this age occur throughout eastern and central Australia (e.g., Li
and Powell, 2001; Figure 3.6). In southern Australia, these consist of Neoproterozoic successions
such as the ca. 700 Ma Sturtian and ca. 600–580 Marinoan glacial successions of South Australia
(e.g., Walter et al., 2000) and similar rocks in Tasmania (Calver and Walter, 2000; Calver et al., 2004)
that are suggested to be related to ‘snowball earth’ and subsequent deglaciation events (e.g., Hoffman
and Schrag, 2002). Also present are widely distributed marine sediments which contain evidence for
continental breakup related to Rodinian rifting (e.g., Cawood, 2005; Crawford et al., 2003a). This
evidence is best preserved (Figure 3.6, Figure 3.7, Figure 3.8) in rocks ca. 600 Ma in age and younger
(to 500 Ma) in the Delamerian Orogen: western Tasmania and King Island (e.g., Calver and Walter,
2000; Calver et al., 2004; Meffre et al., 2004); South Australia (e.g., Drexel and Preiss, 1995; Foden et
al., 2001); western Victoria (VandenBerg et al., 2000; Crawford et al., 2003b); and in the Koonenberry
region, western New South Wales (e.g., Crawford et al., 1997; Direen and Crawford, 2003; Greenfield
et al., 2010, 2011), and possibly the Georgina Basin (Greene, 2010). More equivocal evidence (due to
lack of definitive geochronology) is also found in the Thomson Orogen––in the Greenvale region
(Fergusson et al., 2007a), Charters Towers region (Hutton et al., 1997; Fergusson et al., 2001; 2007c;
Cross et al., 2015a), in the Anakie Province (Withnall et al., 1995) and possibly in the Barnard
Province (Bultitude et al., 1997; Fergusson and Henderson et al., 2013). As summarised by Crawford
et al. (1997; 2003a,b) and Fergusson et al. (2007a,c), rocks of this age contain alkaline and/or
tholeiitic assemblages consistent with rift tectonics and a passive margin and mantle-plume
magmatism. Crawford et al. (2003a) suggested rifting was oriented largely northwest-southeast to
explain the distribution of rift volcanism at this time (Figure 3.7, Figure 3.8).

Palaeogeographic reconstructions of Rodinia often suggest that break-up occurred early, ca. 800–750
Ma, related to a mantle plume, also responsible for initiation of the Centralian Superbasin (e.g., Walter
et al., 1995; Li and Powell, 2001), and that rifting was well offshore of Australia by 600 Ma, (e.g., Li
and Powell, 2001; Li et al., 2008). The apparent abundance of 600–570 Ma rift-related magmatism in
eastern Australia would appear to indicate, as suggested by Crawford et al. (2003a), Fergusson et al.
(2007c), and Glen (2013), for example, that either actual break-up, or a second phase of rifting,
occurred at ca. 600 Ma. Fergusson et al. (2007c) suggested that initial break-up (at 800–600 Ma)

114 Geodynamic Synthesis of the Phanerozoic of eastern Australia


possibly occurred further outboard of eastern Australia to explain the lack of evidence for earlier rifting.
Greenfield et al. (2011) also invoked such rifting at ca. 580 Ma, and show (their Figure 10) the rifting of
a Proterozoic block away from the Koonenberry region at that time (Figure 3.7). This configuration
raises a number of potential difficulties, including: how far north did this rift extend; what is its
relationship to the Thomson Orogen and implications for the basement to the latter; and where are the
present locations of these calved off blocks?

Detrital zircon ages obtained by Fergusson et al. (2007c) in the Thomson Orogen and southern north
Queensland Orogen show that the late Neoproterozoic (ca. 600 Ma) rocks (e.g., Cape River
Metamorphics, lower Argentine Metamorphics (Charters Towers region) and Bathampton
Metamorphics (Anakie Province) generally contain abundant ca. 1200–1000 Ma (Grenville-age)
detrital zircons and only minor 1870–1550 Ma zircons, indicating at best only minor input from
(present-day nearby) cratonic regions such as Mount Isa and Georgetown. Fergusson et al. (2007c)
suggested the ca. 1200 Ma zircons were possibly derived from an extension of the Late
Mesoproterozoic orogenic belt represented by the Musgrave Inlier (central Australia) 1500 km to the
west. Maidment et al. (2007) suggested that similar zircon populations in the Amadeus Basin, central
Australia, reflected uplift and erosion of the Musgrave Province during the Petermann Orogeny. If
Fergusson et al. (2007c) are correct then at least part of the Thomson Orogen may well be underlain
by this older crust. As pointed out by Fergusson et al. (2007c) there is geophysical evidence that
appears to support this (e.g., Kennett et al., 2004). There is also some (more equivocal) support for
this from Sm-Nd isotopic data (Champion, 2013). Gibson et al. (2015) adopted a similar interpretation
and suggested an alternative Rodinian margin that runs across the centre of the Thomson Orogen, not
unlike that in Fergusson et al. (2007c). In such a configuration the conjugate margin can be extended
to join up with pre-Delamerian rocks of the northern Thomson, i.e., the Anakie, Georgetown and
Greenvale provinces (Figure 3.2). We have partly adopted this scenario. The available geochronology
suggests that the northeastern outcropping parts of the Thomson Orogen, e.g., Anakie Metamorphic
Group does indeed correlate with Neoproterozoic-Cambrian units in the Delamerian orogen
(Fergusson et al., 2001). The question then is how widespread are rocks of this age in the Thomson
Orogen? As discussed in Section 2.4.2.2, the best evidence for older (pre-Delamerian) basement in
parts of the non-outcropping Thomson Orogen are in the northern part where metasediments are
overlain by, apparently subhorizontal, Early Ordovician volcanic rocks, and intruded by Early
Ordovician granites (Figure 3.6). The implications are that in order for these granites to have formed in
the Early Ordovician then there must have been deformation (and crustal thickening) prior to this, i.e.,
in the Delamerian Orogeny. Accordingly we show the outcropping regions to the northeast and the
undercover areas in the north overlain by Ordovician volcanics, as being underlain by older crust
(Figure 3.6, Figure 3.7). How far south or west this basement extends is difficult to determine and
more dating is required. Note that this does not negate oceanic crust as basement in (parts of) this
region––only that if it is, then it is older than that commonly inferred for the Lachlan Orogen and
perhaps the southern part of the Thomson Orogen.

In a similar vein, it is not clear where this older region was actually located relative to both the
Koonenberry region and Proterozoic north Queensland, nor, indeed, the relative positions of each
province to one another. The present day geometry of Cambrian to Silurian elements of the Anakie,
Charters Towers, Greenvale and Barnard provinces, suggests some rotation of these blocks (relative
to one another), at the very least. One speculative possibility (Figure 3.7, Figure 3.8) is that these
blocks lay further to the northeast. Tectonic reconstructions are hindered by the lack of definitive
geochronology. Moreover, there is no apparent magmatic record in this region of any subsequent
Cambrian arc, in contrast to the geological record further south. The earliest calcalkaline magmatism
in northern Queensland is at ca. 480–470 Ma (in the Charters Towers and Greenvale provinces,
Figure 3.13), making the location of the northern blocks, prior to the early Ordovician, very uncertain.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 115


Blocks of this age, and possibly older, also occur within both the Lachlan and New England orogens,
most notably within the Selwyn Block (VandenBerg et al., 2000). The Selwyn Block hypothesis
proposes the presence of older continental basement beneath the Melbourne Zone, which has been
linked to western Tasmania (Cayley et al., 2002; Cayley, 2011; Figure 3.6, Figure 3.7). Although some
authors have suggested the Lachlan Orogen is underlain by oceanic crust, only (e.g., Gray, 1997;
Gray and Foster, 1997; Figure 3.3, Figure 3.4), the presence of the Selwyn Block appears to be
confirmed by the recent Victorian seismic acquisition (Korsch et al., 2008; Cayley et al., 2011). The
latter seismic data show the Melbourne Zone to be underlain by something distinct to the interpreted
oceanic crust basement in the geological zones to the west. The Selwyn Block contains Cambrian
calcalkaline volcanics (e.g., VandenBerg et al., 2000; Spaggiari et al., 2003), which have many
similarities to, and have been correlated with, the Mount Read Volcanics (Crawford et al., 2003a,b).
Crawford et al. (2003a) suggested that the Selwyn Block may have been part of Australia, prior to
being split off by rifting at 600 Ma (Figure 3.7), though the actual origin and location of Tasmania and
the Selwyn Block––the VanDieland microcontinent of Cayley (2011)––is uncertain (e.g., see Cayley,
2011; Gibson et al., 2011; Figure 3.8). Glen (2013) has suggested that the Hay-Booligal block in NSW,
north of the Selwyn Block, may also represent part of the Australian continent calved off by rifting at
600 Ma, though evidence either way is minimal. He further suggested that part of the New England
basement may have also evolved similarly, and there is some isotopic evidence for older basement
rocks and lithosphere in the southern New England Orogen (e.g., Powell and O’Reilly, 2007;
Champion, 2013). The Anakie Province may also have formed in an analogous manner; Fergusson et
al. (2001) have suggested that this province may have been related to splitting or rifting of a younger
microcontinent from the Gondwanan margin. Glen (2013) presented a similar model for the Anakie
Province though pointed out that rifting may have been Neoproterozoic, Cambrian or Ordovician in
age. The available data for the Thomson Orogen make this difficult to prove or disprove.

Late Neoproterozoic and early (to late) Cambrian mafic and ultramafic rocks also occur within the
Lachlan and New England orogens (Figure 3.6, Figure 3.7). These rocks are of some importance as
they provide possible constraints on arc-related environments that must have been present out board
of the Rodinian (rift and) passive margin and which eventually interacted with this margin during the
later parts of the Delamerian cycle. These rocks comprise mafic and ultramafic Cambrian (and older?)
igneous rocks, preserved as remnants along major faults, in the Victorian part of the Lachlan (e.g.,
VandenBerg et al., 2000, Spaggiari et al., 2004). These include an ultramafic and a tholeiitic-boninitic
association (e.g., Crawford and Keays, 1987; Crawford et al., 1984, 2003b; VandenBerg et al., 2000),
that are typically interpreted as having formed in a suprasubduction zone environment (e.g., Crawford
and Keays, 1987; Crawford et al., 1984, 2003a,b; Spaggiari et al., 2003, 2004). In the Stawell Zone,
the mafic volcanics (Magdala Volcanics) have a backarc signature and are underlain by continental-
derived turbiditic sediments (Crawford et al., 2003b; Squire et al., 2006; Figure 3.7). These authors
interpreted the Stawell succession to represent a distal backarc environment, related to a west-dipping
subduction zone to the east. This is also consistent with the observation that at least some of these
mafic-ultramafic successions, for example, in the Bendigo and Tabberabbera zones, appear to form
the basement to those zones (e.g., Spaggiari et al., 2003, 2004; Korsch et al., 2008), that is, floored by
oceanic crust as suggested by Gray, Foster and co-workers (e.g., Gray and Foster, 2004).

116 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.6 General distribution of inferred Neoproterozoic to mid Cambrian (ca. 600 Ma to 515 Ma), pre-
Delamerian rocks in eastern Australia. Ages and locations of early Ordovician felsic magmatic rocks in the
undercover Thomson Orogen (from drill core) are also shown. These are shown here to indicate the inferred
minimum extent of Late Neoproterozoic–Early Cambrian (or older) crust basement in the Thomson Orogen. Refer
to text for discussion and references. Rifted sequences, possibly of this age may also occur in parts of the
Georgina basin (Greene, 2010).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 117


Figure 3.7 Interpreted tectonic environment of eastern Australia for the late Neoproterozoic to early Cambrian
period––ca. 600 to Ma 515 Ma. Refer to text for discussion of tectonic interpretation. Location of the Melbourne
Zone and Tasmania in this time period is uncertain (see Figure 3.8).

118 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.8. A speculative tectonic reconstruction for the Tasman Orogen in late Neoproterozoic to early Cambrian
time. The inferred extant blocks of the Thomson Orogen are shown reassembled as a microcontinent block, lying
somewhere to the northeast of Australia at this time. The timing of rifting along the present-day eastern margin of
the Thomson Orogen is poorly constrained and may have been earlier or later. The Selwyn Block/VanDieland
microcontinent is situated south of Australia. Note that this figure assumes little relative movement between
northern and southern Australia since the Late Neoproterozoic.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 119


These are overlain, conformably in places, by Cambrian, deep marine, often pelagic sedimentation
(VandenBerg et al., 2000; Spaggiari et al., 2003). Examples, such as those in the Bendigo Zone and
further east, were apparently not affected by the Delamerian Orogeny (Spaggiari et al., 2003, 2004).
Mafic-ultramafic successions in the west, including those in the western Stawell Zone, were deformed
during the Delamerian Orogeny (e.g., Miller et al., 2005). Gibson et al. (2015) have suggested that
some of these may actually represent underlying mantle uplifted in a hyperextended margin
environment, prior to commencement of the Delamerian Orogeny.

The New England Orogen also contains similar mafic-ultramafic remnants. These are Neoproterozoic
to Cambrian tectonic blocks, largely of oceanic fragments, including island arc-related remnants
(Figure 3.6). They occur in the southern NEO, as accreted blocks along the Peel-Manning Fault
System (e.g., Offler and Shaw, 2006), and in the northern NEO, as represented by the ca. 565 Ma
Princhester and related ophiolites (Bruce et al., 2000, Murray and Blake, 2005). In both the Lachlan
and New England orogens, these rock associations are commonly interpreted to provide records of
subduction and other oceanic environments outboard of continental Australia in the Neoproterozoic
and Cambrian (Figure 3.7). The initiation of early Paleozoic subduction in the southern New England
Orogen is recorded by the formation of suprasubduction zone ophiolites (~530 Ma; Aitchison and
Ireland, 1995; Fanning et al., 2002; Sano et al., 2004), as well as blocks of (Late Neoproterozoic–)
early Cambrian eclogite and intrusive rocks (~530 Ma ages; Aitchison et al., 1992a; Sano et al., 2004;
Watanabe et al., 1998; Fanning et al., 2002), and rocks representative of Cambrian magmatic island
arc development (Cawood and Leitch, 1985). Whereas contacts between exposed intraoceanic
elements are faulted, their character and distribution suggest development in an east-facing arc with
the Cambrian ophiolite separating, and underlying, the western arc-flanking basin from the eastern
accretionary prism (Cawood and Leitch 1985; Holcombe et al., 1997a, b; Jenkins et al., 2002). This
record of Cambrian subduction is slightly earlier than boninitic magmatism recorded in Tasmania
(514 ± 5 Ma; Black et al., 1997) and in Victoria (519–514 Ma; VandenBerg et al., 2000).

Rocks at Port Macquarie also contain a fragmentary late Neoproterozoic to early Paleozoic history that
includes subduction and associated metamorphism under high P- low T conditions at ~536 Ma
(Watanabe et al., 1998; Och et al., 2007). These ages suggest subduction had begun by at least 530
Ma (e.g., Li and Powell, 2001), and possibly earlier (e.g., Gray and Foster, 2004). Glen (2013) has
recently reinterpreted such rocks and presented a unified model. He suggested that the early
Cambrian arc (represented by ca. 530 Ma remnants in the NEO) is the same arc that was
subsequently present along the Delamerian margin of eastern Australia, i.e., responsible for arc
volcanism in the Koonenberry region (Greenfield et al., 2010, 2011), western Victoria and further
south. The arc subsequently migrated eastwards, evidence for which is represented by the various
remnants found in the Lachlan Orogen (discussed above) and also in the NEO (Glen, 2013). This
model is discussed further in the next section.

Unrestored, and (a possible) restored tectonic interpretations for the Late Neoproterozoic are shown in
Figure 3.7 and Figure 3.8. As shown in the figures and discussed previously most uncertainties
concern the location of the Selwyn Block (VanDieland of Cayley, 2011), and the location and
reconstruction of the Thomson orogen blocks and northern extensions. The latter is hindered not just
by lack of geochronology and exposure but also uncertainty regarding the relative positions of blocks
within the orogen.

120 Geodynamic Synthesis of the Phanerozoic of eastern Australia


3.2.2 Early to latest Cambrian (ca. 515 Ma to 490 Ma)
The breakup of Rodinia and associated extension of southeastern Australia between the late
Neoproterozoic (ca. 600 Ma) and the early Cambrian was halted with the onset of subduction and
accompanying contractional orogenesis––called the Delamerian Orogeny (e.g., Foden et al., 2006).
The best evidence for this orogeny is in southern Australia, western Victoria, Tasmania, the Selwyn
block and the Koonenberry region (and in the then contiguous Antarctica and associated Ross
Orogeny). This is in distinct contrast to the northern Thomson Orogen and the north Queensland
region where evidence is at best equivocal.

In South Australia, western Victoria and Tasmania, the Delamerian Orogeny (Tyennan Orogeny in
Tasmania) commenced by at least ca. 515 Ma (e.g., Foden et al., 2006; Seymour and Calver, 1995;
Figure 3.9, Figure 3.10), and possibly earlier at ca. 545 Ma (e.g., Turner et al., 2009). In South
Australia, the Delamerian Orogeny was long-lived, from ca. 515 (and earlier?) to 490 Ma (e.g., Drexel
and Preiss, 1995; Foden et al., 2006). VandenBerg et al. (2000), amongst others, suggested that there
may have been two significant stages, ca. 515 and ca. 490 Ma, and indicated that, in Victoria, the first
phase is evident in the Glenelg Zone, the second only in the Grampians-Stavely Zone. Miller et al.
(2005) showed that the Delamerian Orogeny also affects the western part of the Stawell Zone, based
on metamorphic ages of ca. 500–490 Ma. Western Tasmania also records at least two discrete
deformational events, separated by a significant extension event, and VandenBerg et al. (2000)
showed that, in many respects, western Victoria and western Tasmania share similar geological
histories.

The Delamerian Orogeny in both western Tasmania, and perhaps western Victoria, has been
suggested to have been triggered by arc-continent collision around 515–510 Ma (Berry and Crawford,
1988; VandenBerg et al., 2000; Crawford et al., 2003a), by an east-dipping subduction zone east of
Tasmania (e.g., Crawford and Berry, 1992; Crawford et al., 2003a; Gibson et al., 2011), which flipped
to west-dipping after collision. In both areas, collision is interpreted to have been accompanied by the
accretion of Cambrian forearc boninitic crust––the Tasmanian mafic-ultramafic complex in Tasmania
(Crawford and Berry, 1992; Crawford et al., 2003a), and (parts of) the Dimboola Igneous Complex in
western Victoria (VandenBerg et al., 2000; Crawford et al., 2003b; Figure 3.9). Miller et al. (2005),
Foden et al. (2006) and Cayley (2011), amongst others, have suggested, however, that western
Victoria is more consistent with west-dipping subduction and that emplacement of the Dimboola
Igneous Complex was via back-thrusting and not by continent-arc collision. In these models the
Delamerian deformation is related to either subduction (e.g., Miller et al., 2005; Foden et al., 2006)
and/or docking of the microcontinent VanDieland with the Antarctic mainland, further south (Cayley,
2011).

Regardless, most tectonic reconstructions agree that, by the late Delamerian Orogeny, subduction
was west-dipping (e.g., Gray and Foster, 2004; Foden et al., 2006; Cayley, 2011; Gibson et al., 2011,
2015). The presence of ca. 510–500 Ma calcalkaline volcanics in western Tasmania, western Victoria
and the Selwyn Block (Figure 3.9, Figure 3.11), all appear consistent with a west-dipping arc. As noted
by Squire et al. (2006), this scenario also explains the observed temporal sequence (forearc to
backarc) evident in many of the mafic-ultramafic complexes (seafloor remnants) preserved in both the
Delamerian and Lachlan orogens (Crawford et al., 2003b).

High temperature, low pressure, metamorphism and metamorphic complexes were developed in both
Tasmania and western Victoria, and syntectonic I- and S-type granites were emplaced in the Glenelg
River Metamorphic Complex in Western Victoria (VandenBerg et al., 2000; Crawford et al., 2003b).
Both regions are suggested to have undergone subsequent post-collisional extension (possibly in a
backarc environment), leading to emplacement of calcalkaline volcanics––the Mount Read Volcanics,

Geodynamic Synthesis of the Phanerozoic of eastern Australia 121


correlatives, and intrusives, in Tasmania (Crawford and Berry, 1992; Crawford et al., 2003a), and the
Mount Stavely Volcanic Complex in Victoria (Crawford et al., 1996, 2003b; VandenBerg et al., 2000).
This was followed by renewed deformation at ca. 500 Ma to 490 Ma.

The Selwyn Block, Victoria, contains Cambrian calcalkaline volcanics (Jamieson-Licola, e.g.,
VandenBerg et al., 2000; Spaggiari et al., 2003; Figure 3.9, Figure 3.11), which have similarities to,
and have been correlated with, the Mount Read Volcanics in Tasmania (Crawford et al., 2003a, b).
Crawford et al. (2003b) suggested they may represent ‘along strike continuations’ of the Mount Read
Volcanics. Cayley (2011) suggested a similar scenario. In the oceanic crust basement model, the
calcalkaline rocks are interpreted as island arc volcanics, e.g., Jamieson Island Arc (Gray and Foster;
2004; Spaggiari et al., 2003; Figure 3.4).

Other evidence for island arc terranes of this age are present in the New England Orogen (e.g., Offler
and Shaw, 2006), as well as in New Zealand, e.g., the ca. 515 Ma Takaka Island Arc (e.g., Munker
and Crawford, 2000). The southern New England Orogen in this period comprises Cambrian tectonic
blocks, largely of oceanic fragments, including island arc-related remnants, such as suprasubduction
zone ophiolites (~530 Ma; Aitchison and Ireland, 1995; Fanning et al., 2002; Sano et al., 2004;
Figure 3.11) and Cambrian magmatic arc rocks (Cawood and Leitch, 1985), which provide a record of
continuing subduction environments offshore of continental Australia in the pre-Delamerian and
Delamerian, that is, (latest Neoproterozoic-) Cambrian. These remnants have been interpreted to
suggest an east-facing arc (e.g., Cawood and Leitch 1985). Rocks at Port Macquarie also contain late
Neoproterozoic to early Paleozoic remnants that indicate subduction and associated metamorphism
under high P low T conditions at ~536 Ma (Watanabe et al., 1998; Och et al., 2007). Cambrian
magmatic arc development in the New England Orogen is also recorded by Cambrian volcaniclastic
rocks, apparently derived from a low-K intra-oceanic island arc (Cawood and Leitch, 1985). These
rocks––Murrawong Creek and Pipeclay Creek Formations––occur in the Gamilaroi Terrane
immediately west of the Peel-Manning Fault. These oceanic remnants in the Lachlan and New
England Orogens provide important tectonic constraints. They represent fragments of oceanic and
island arc crust that were accreted during the Delamerian Orogeny and later, and are consistent with
eastern Australia not only facing the Paleo-pacific Ocean since the late Neoproterozoic and earliest
Paleozoic but also consistent with an overall arc environment for much (all?) of this time (e.g.,
Crawford et al., 2003a; Gray and Fergusson, 2004; Collins and Richards, 2008; Li et al., 2008;
Figure 3.1). They also indicate that parts of the now contiguous orogens were probably significantly
separated in the Early Paleozoic, as suggested by many authors (e.g., VandenBerg et al., 2000;
Cayley et al., 2002; Gray and Foster, 2004; Figure 3.11).

Evidence for the Delamerian Orogeny is also found in New South Wales and Queensland. Although
broadly similar ages to South Australia and Victoria have been recognised, deformation appears to
have been shorter-lived farther to the north (Foden et al., 2006; Black, 2007; Fergusson et al., 2007a,
b). The Delamerian Orogeny deformed and metamorphosed rocks of the Anakie Province, Charters
Towers region (see below), Koonenberry Belt and central Thomson Orogen. Based on comparative
evidence from the Anakie Province, Draper (2006) suggested that deformation of subsurface
metasedimentary rocks in the eastern Thomson Orogen was also likely to have occurred during the
Delamerian Orogeny. The contractional event was predominantly east-west in the Thomson Orogen,
and northwest-southeast in the Koonenberry Belt (Gilmore et al., 2007; Greenfield et al., 2010, 2011).

122 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.9 General distribution of Cambrian (ca. 520 Ma to ca. 490 Ma) rocks in eastern Australia. Refer to text for
discussion.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 123


The Delamerian Orogeny is poorly recorded in the Warburton Basin (e.g., Murray and Kirkegaard,
1978) though uplift and erosion of the basin was apparently coincident with the orogeny (recognised
as the Mootwingee Movement; Gravestock and Gatehouse, 1995). Distal effects of the orogeny also
appear to be recorded in the Georgina Basin (e.g., Greene, 2010, and references therein).

Delamerian deformation was accompanied by ca. 510 Ma calcalkaline magmatism in the Koonenberry
Belt (Crawford et al., 1997; Gilmore et al., 2007; Greenfield et al., 2010, 2011), and both are
suggested to relate to the presence of a continental arc. Similar-aged felsic magmatism is also
recorded in the Warburton Basin (Gatehouse, 1986; Draper, 2006) where it has also been attributed to
an arc (Gatehouse, 1986), though may also reflect backarc magmatism given its felsic nature.

Notably, magmatic rocks of this age (or nature) have not been recorded within the Thomson Orogen
or northern Queensland, with the apparent exception of the ca. 508 Ma Bucklands Hill Diorite in the
Charters Towers region (Fergusson and Henderson et al., 2013). This unit, an older phase of the
largely post-Delamerian Macrossan Igneous Association, is tholeiitic with a relatively primitive isotopic
signature (Fergusson and Henderson et al., 2013). It is equivocal whether it is arc-related or not, and
so there is little definitive evidence for the existence of an arc environment at this time in north
Queensland. Although this may simply reflect a lack of preservation (e.g., detrital zircons of this age
are present, e.g., Cross et al., 2015), it is noted that relatively widespread evidence exists in this
region for an arc-environment post-Delamerian in this region (e.g., Henderson, 1986; Withnall et al.,
1991; Stolz, 1994; Henderson et al., 2011). Similarly, effects of the Delamerian Orogeny itself are
poorly represented in north Queensland, though this may partly reflect the geochronological
uncertainty of many of the units. The best evidence appears to be within the Greenvale Province (ca.
520–510 Ma; Nishiya et al., 2003) and in the Charters Towers region (ca. 495 Ma; Fergusson et al.,
2007a). Potential Delamerian deformational may also occur in the Georgetown and Coen regions, but
geochronological data are missing. A number of deformations that could be interpreted as Delamerian
have, however, been shown, at least partly, to represent post-Delamerian extension (Fergusson et al.,
2007a,b).

The possible presence of an arc in the Warburton-Koonenberry region at this time (Figure 3.11), well
west, and (presently) inboard, of the Anakie Province, is also problematical and suggests that the
Anakie Province, and probably the Charters Towers and Greenvale provinces, were not in their
present relative positions. Withnall et al. (1995), conversely, suggested that this configuration may
have either reflected a very wide Delamerian Orogen or subsequent (post-Delamerian) extension and
rifting of the Anakie Province eastwards, away from the western Thomson (somewhat analogous to
the model of Glen, 2013). The former is problematical on width alone, and the latter does not appear
to be consistent with the recent discovery of apparently flat-lying Ordovician Volcanic rocks overlying
deformed and metamorphosed metasedimentary rocks in the central Thomson Orogen (Draper,
2006). One possible solution to this is that this rifting may have been occurring during the Delamerian
Orogeny. The difficulties are further exacerbated if the volcanic rocks in the Warburton Basin are
considered arc-related as has been suggested (Gatehouse, 1986), as the reconstructions of
Greenfield et al. (2011) suggest there must have been at least some oceanic crust to the east of the
Koonenberry belt at this time, post-Rodinian rifting. Fergusson and Henderson (their Figure 3.121, in
Fergusson and Henderson et al., 2013) adopted an intermediate view by continuing the Koonenberry
arc, via a transform, to an arc east of the Anakie and Charters Towers provinces. The ca. 508–505 Ma
Kalkarindji continental flood basalt province (Glass and Phillips, 2006), in northwestern Australia, may
have played some (indirect) part in the Warburton Basin magmatism. The nearest definitive Kalkarindji
volcanism are the Colless Volcanics in the Georgina Basin (Glass and Phillips, 2006; Figure 3.9).
Although presumably not directly related to the Kalkarindji event, the Warburton Basin volcanics may
result from extension related to the arc to the south-east and to the flood basalt province to the north-
west. It could be speculated that this configuration may have been responsible for initiating a triple

124 Geodynamic Synthesis of the Phanerozoic of eastern Australia


junction and rift that some tectonic reconstructions invoke for the formation of the post-Delamerian
Larapintine seaway (e.g., Gray and Foster, 2004; Maidment et al., 2007; Figure 3.1, Figure 3.15).
Such incipient rifting may also have moved, and perhaps rotated, parts of the now eastern Thomson
Orogen, such as the Anakie Province, to the east.

The geological history of the Koonenberry region also has implications for tectonic reconstructions to
the south. As indicated by Greenfield et al. (2011) Cambrian rocks, locally overlie, and are structurally
concordant with, ca. 580 Ma Neoproterozoic rocks, implying no structural disturbance in the period
between the two units, suggesting no early collisional event occurred in this region. The geological
evidence from the Koonenberry region (Greenfield et al., 2011) favours consistent west-dipping
subduction, and suggests that a similar situation probably also occurred along strike in western
Victoria, i.e, western Victoria also records west-dipping subduction throughout the Delamerian cycle
(e.g., Gray and Foster, 2004; Foden et al., 2006; Figure 3.1, Figure 3.11) with no prior opposite
subduction and subsequent collision and subduction flip as advocated by Crawford and Berry (1992),
Munker and Crawford (2000), and Gibson et al. (2011). Regardless of the correct scenario for western
Victoria it would appear that western Tasmania (and the rest of the Selwyn Block?) had a slightly
different tectonic history than the Koonenberry region, which in turn appears to differ from further north
in the eastern Thomson and north Queensland region where evidence for a Delamerian arc is at best
equivocal.

Unanswered questions concern the position of Tasmania (VanDieland of Cayley, 2011; Figure 3.11),
i.e. along strike of western Victoria and part of Antarctica (e.g., Gibson et al., 2011), or offshore and
colliding with Australia-Antarctica during the Delamerian Orogeny (e.g., Cayley, 2011; Munker and
Crawford, 2000). Also unanswered are the number of arcs present during this time, or more
specifically, the relationship between preserved ca. 535 Ma and 500 Ma fragments in the New
England Orogen versus those in the Delamerian and Lachlan orogens (Figure 3.11). As noted in the
previous section Glen (2013) has recently suggested that many of the preserved fragments of arc-
related material, including supra-subduction zone relicts, in the Delamerian, Lachlan and New England
orogens may be remnants of the one advancing and retreating arc, i.e., starting in the east (New
England at ca. 535–530 Ma, advancing to the Delamerian margin (ca. 515–510 Ma) and subsequently
retreating, with successively younger Cambrian relicts preserved in the Lachlan and New England
orogens. Although this is a possibility, we have suggested the presence of multiple subduction zones
here. Finally, the tectonic settings for the eastern and northern Thomson Orogen and north
Queensland are not clear, nor are the positions of the eastern and northern Thomson Orogen relative
to the Australian mainland as it then was. Our preferred model for the Delamerian cycle (Figure 3.11)
is somewhat akin to that of Cayley (2011) for the southern part of Australia and our own for the
northern part.

Following the Delamerian Orogeny, the arc appears to have shifted well to the east, at least in
southern Australia, as represented by the interpreted Ordovician Macquarie Arc (Crawford et al.,
2007a; assuming that this is indeed an arc). This easterly shift, in part, reflects continuing roll-back as
evidenced by the forearc volcanism-backarc volcanism-deep marine sediment sequences of Cambrian
age preserved in fault-bounded blocks in the Lachlan Orogen, such as the Heathcote Greenstone Belt
(Crawford et al., 2003b).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 125


Figure 3.10 Generalised (approximate) distribution of deformation within the Delamerian Orogeny (ca. 520–490
Ma) as defined by outcrop and drill hole information (Thomson Orogen), except for the Selwyn Block. Extent of
deformation in the latter has been extrapolated from minor outcrops, e.g., Waratah Bay (e.g., VandenBerg et al.,
2000). For all areas actual extent of deformation may be greater and more continuous. The intensity of
deformation is variable within the areas indicated. See text for data sources.

126 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.11 Interpreted tectonic model for eastern Australia for the early to latest Cambrian (ca. 520–490 Ma).
Refer to text for detailed discussion. Location of the Melbourne Zone and Tasmania in this time period is
uncertain.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 127


Figure 3.12 A speculative tectonic reconstruction for the Tasman Orogen for the early to latest Cambrian. The
inferred extant blocks of the Thomson orogen are shown reassembled as a microcontinent block, lying
somewhere to the northeast of Australia at this time. The presence of convergence along the present-day eastern
margin of the Thomson Orogen is poorly constrained, with little unequivocal evidence. The Selwyn
Block/VanDieland microcontinent is situated south of Australia. Note that this figure assumes little relative
movement between northern and southern Australia since the latest Cambrian.

128 Geodynamic Synthesis of the Phanerozoic of eastern Australia


3.2.3 Latest Cambrian to early Silurian (ca. 490 Ma to 430 Ma)
Eastern Australia through this period (post-Delamerian Orogeny to Benambran Orogeny) is dominated
by two contrasting rock packages:

• deep water quartz-rich turbidites of cratonic provenance and associated pelagic sediments,
commonly interpreted as a backarc and/or passive margin environment.
• calcalkaline magmatism and volcaniclastics and marine sediments with common carbonates,
commonly interpreted as having formed in oceanic arcs, and/or backarc environments.

These contrasting rock packages are best exemplified in the Lachlan Orogen. From the Late Cambrian
to the end of the Ordovician–early Silurian, most of the Lachlan Orogen (New South Wales, Victoria and
Tasmania) was the site of deep marine sedimentation (Figure 3.13). These sediments consist of quartz-
rich turbiditic successions (Western, Central and Eastern Lachlan) and late Middle to Late Ordovician
black shale-dominated sediments (Central and Eastern Lachlan). In a number of regions, for example,
the Bendigo and Tabberabbera zones (e.g., VandenBerg et al., 2000; Fergusson and VandenBerg,
2003; Spaggiari et al., 2003), this sedimentation appears to be conformable upon mafic and ultramafic
rocks interpreted as oceanic crust (see previous section). Remnants of mafic volcanics, chert,
serpentinites and ultramafic rocks interpreted as oceanic crust or ocean islands also occur within New
South Wales (e.g., Warren et al., 1995; Prendergast and Offler, 2012) and are also interpreted to
probably represent basement to the turbidite successions. In most areas of the Lachlan Orogen, this
deep water sedimentation ended with the Benambran Orogeny. Sedimentation, however, continued in
both the Melbourne Zone and northeastern Tasmania, and both these regions show no evidence for the
Benambran Orogeny (e.g., Fergusson and VandenBerg, 2003; Seymour and Calver, 1995; Figure 3.14).

Contemporaneous with quartz-rich turbiditic sedimentation was the Early Ordovician to earliest Silurian
Macquarie Arc––the Macquarie Volcanic Belt of Quinn et al. (2014). The belt is commonly interpreted
to have formed in an intra-oceanic arc setting (e.g., Crawford et al., 2007a), though Wyborn (1992),
and more recently Quinn et al. (2014), for example, have suggested an alternate, non-arc, setting. The
remnants of the arc, which include calcalkaline and shoshonitic volcanic rocks, intrusions and
volcaniclastic and carbonate-rich successions, are preserved as four elongate remnants, almost totally
located in New South Wales though extending into northern Victoria (Fergusson and VandenBerg,
2003; Figure 3.13). Detailed work by Glen, Crawford and co-workers (e.g., Crawford et al., 2007a and
companion papers) suggests that the Macquarie Arc was built up over four successive phases of
growth, within two distinct (east and west) provinces, which may not have been together until accretion
(Percival and Glen, 2007). Other possible exotic oceanic rocks, interpreted to have accreted to
Australia during the Benambran Orogeny, are the deep marine sediments and underlying mafic
volcanics of the Narooma Terrane (Glen et al., 2004; Figure 3.13). Miller and Gray (1997), and
Prendergast and Offler (2012), however, suggested these rocks do not represent an exotic terrane,
but rather form part of an accretionary wedge–the Narooma accretionary complex–that was related to
subduction in the Ordovician.

Most workers indicate that one or more submarine fans were the sites for the Ordovician deep marine
sedimentation (e.g., Fergusson and VandenBerg, 2003; Gray and Foster, 2004; Glen, 2005; Glen et
al., 2007b). This may have reflected uplift of the Delamerian Orogen in the Early Ordovician (e.g.,
VandenBerg et al., 2000; Fergusson and VandenBerg, 2003), although there is uncertainty regarding
the possible relative positions of the respective parts of the Lachlan Orogen at this time. The major
change in sedimentation recorded by the switch to black shale-dominated pelagic sediments in the
late Middle Ordovician and their localisation largely to the Central and Eastern Lachlan, is in part
contemporaneous with early Benambran deformation (ca. 455 Ma) recorded in the Western Lachlan
(VandenBerg et al., 2000; Gray et al., 2003; Miller et al., 2005).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 129


Two end-member tectonic models––oceanic island arc versus marginal basin rift––have been invoked
for the Macquarie Arc, each with significant implications for tectonic reconstructions. Both models
agree that the volcanics are arc-related, either directly within an oceanic island arc environment (e.g.,
Crawford et al., 2007b,c) or indirectly by reworking of previously arc-metasomatised mantle (e.g.,
Wyborn, 1992). In the arc model, the turbiditic sediments of the Lachlan Orogen have been suggested
to have little or no evidence for volcanic detritus that may have come from the arc. This led numerous
workers (e.g., Gray and Foster, 2004; Meffre et al., 2007), to suggest that the contemporaneous
Macquarie Arc was disconnected from this deep marine sedimentation, perhaps by hundreds of
kilometres (Meffre et al., 2007). In this model, the tectonic environment for the quartz-rich sediments is
commonly interpreted as a passive margin environment, well behind the Macquarie Arc (e.g.,
Fergusson and VandenBerg, 2003; Gray and Foster, 2004; Glen, 2005; Glen et al., 2007b; Figure 3.1
to Figure 3.5). The quartz-rich sediments and the Macquarie Arc are thought to have been juxtaposed
when the arc accreted to eastern Australia, in the early Silurian as part of the Benambran Orogeny
(Glen et al., 2007b; Meffre et al., 2007).

In the non-arc models of Wyborn (1992) and Quinn et al. (2014) there is no requirement for significant
relative movement of the quartz-rich and volcanic sediments. As shown by Quinn et al. (2014),
however, based on detailed mapping and biostratigraphy, it appears that conformable interfingering
relationships exist between sediments and volcanic layers in the Kiandra area. This is further
supported by the observation that the appearance of volcanism in the Macquarie Arc, throughout the
Ordovician, is contemporaneous with switches in the surrounding turbiditic sequences from deposition
of sands to deposition of black shales and cherts (Quinn et al., 2014). This synchronicity indicates a
link between the two sequences, and led Quinn et al. (2014) to the conclusion that the rocks of the
Macquarie Arc were not significantly separated spatially from the turbiditic sediments when they
formed. This in turn led to the suggestion that the volcanics were generated in an episodic rift
environment within a marginal basin, in distinct contrast to the island arc models. Additional evidence
for a closer spatial association between the two sequences is provided by detrital zircon and other
data. Ages and Hf isotopic signatures of detrital zircons in sediments associated with phase 1 of the
Macquarie Arc (ca. 480 Ma) suggest a provenance similar to that of the Ordovician sediments (Glen et
al., 2011). There is also evidence for young detrital zircons (ca. 476 Ma) within the Ordovician
sediments of the Girilambone Group in the northern Lachlan Orogen (Gilmore et al., 2012). These
sediments also contain Early Ordovician age basaltic rocks of MORB and OIB composition (Burton et
al., 2012, 2013) as well as VHMS mineralisation which has Pb isotopic model ages of ca. 490–470 Ma
consistent with a back-arc environment (Huston et al., 2015). Further support is provided by the Fifield
suite (e.g., Barron et al., 2004) which occurs as a >300 km belt within the eastern part of the
Girilambone Group. These rocks have an age similar to the last phase of (shoshonitic) magmatism in
the adjacent Macquarie Arc (Fraser et al., 2014), and as pointed out by Barron et al. (2004) the
feldspathic components of the Fifield suite have a shoshonitic affinity. This suggests that this part of
the Girilambone Group was underlain by the Macquarie Arc mantle.

Interestingly, Fergusson et al. (2013; their Figure 9) have presented a modified version of the arc
model which may explain some of the observations for a closer spatial association. In their model the
Macquarie Arc has an east-west strike orientation (similar to Cayley, 2011 and Gibson et al., 2011),
with the Ordovician turbidite fans to both the north and south.

130 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.13 General distribution of latest Cambrian to early Silurian (ca. 490 Ma to ca. 430 Ma) rocks in eastern
Australia. Ages of felsic magmatic rocks in the undercover Thomson Orogen (from drill core) are also shown.
Refer to text for discussion and references.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 131


Although largely dismembered, the north Queensland region (e.g., Henderson, 1987), like the Lachlan
Orogen, consists of both quartz-rich sediments and calcalkaline volcanism (Figure 3.13). The region
consists of Ordovician deep water (turbiditic), dominantly quartz-rich sediments, preserved within fault-
bounded remnants east of, and probably derived from, the outcropping Mesoproterozoic basement in
the Georgetown and Coen regions (e.g., Withnall and Lang, 1993; Garrad and Bultitude, 1999). The
sediment successions in north Queensland also contain interlayered tholeiitic magmatism (Withnall
and Lang, 1993; Bultitude et al., 1997; Withnall et al., 1997a), consistent with an extensional
environment. Contemporaneous calcalkaline magmatism is preserved as Early to Middle Ordovician
volcanic- or volcaniclastic-dominated successions (e.g., Seventy Mile Range Group, Balcooma
Metavolcanic Group: Henderson, 1986; Withnall et al., 1991; Stolz, 1995; Fergusson et al., 2007a)
and Early and Late Ordovician volcanic and carbonate-dominated sequences, for example, in the
Broken River Province (Withnall and Lang, 1993; Figure 3.13). Many authors have suggested backarc,
continental-margin arc or island-arc affinities for the sediments and calcalkaline successions (e.g.,
Withnall et al., 1991, 1997b; Henderson, 1986; Stolz, 1994), suggesting an environment not dissimilar
to Lachlan Orogen rocks of the same age (e.g., Gray and Foster, 2004; Glen, 2005). Unlike the
Lachlan Orogen, however, units in north Queensland locally contain both quartz-rich marine
sediments and calcalkaline volcanics (e.g., the Judea Formation; Withnall and Lang, 1993), as well as
volcanic clasts in conglomerate which appear to correlate with known calcalkaline volcanic units in the
region (Garrad and Bultitude, 1999). These features suggest proximity between arc-related volcanism
and cratonic-derived sedimentation, and support suggestions that the quartz-rich sediments were
deposited within a back-arc environment. A similar scenario is recorded within the Anakie Province
(eastern Thomson Orogen) which contains Late Ordovician marine sediments, including carbonates,
and associated mafic to intermediate volcanic rocks (Fork Lagoons Beds, Withnall et al., 1985;
Figure 3.13). The sediments appear to be derived from cratonic and volcanic provenances (Fergusson
et al., 2007c) and the volcanic rocks have geochemistry consistent with either arc or backarc
environments (Withnall et al., 1995). The Warraweena Volcanics in the southern Thomson Orogen, in
New South Wales (Figure 3.13), contain calcalkaline volcanics, interpreted by Burton et al. (2008) to
be arc-related. According to Watkins (2007) and Burton et al. (2008), these appear to be chemically
similar to (nearby) Macquarie Arc magmatism. Notably, contemporaneous within-plate magmatism is
also recorded, which Watkins (2007) placed in a backarc environment. The age of the Warraweena
Volcanics, however, is uncertain. The unit could be as old as Neoproterozoic (e.g., Glen et al., 2013)
or as young as Silurian (e.g., Fraser et al., 2014). Early to Middle Ordovician felsic (rhyolitic) volcanism
and associated granites, occur within the central Thomson Orogen. Chemical affinities are unknown.
Draper (2006) suggested the magmatism may relate to extension and crustal thinning in the
Ordovician.

West of the Thomson Orogen, this time period corresponds with the onset of more widespread marine
basin sedimentation following the Delamerian Orogeny and intracratonic rifting (e.g., in the Warburton
Basin; Gatehouse and Cooper, 1986). Progressive marine incursion caused an expansion of basins
across central Gondwana as a shallow epieric sea (Larapintine Sea; Webby, 1978; Figure 3.13),
believed to link the Warburton, Amadeus and Canning Basins (Webby, 1978; Li and Powell, 2001;
Gravestock and Gatehouse, 1995; Maidment et al., 2007). This seaway may have also connected
basinal environments in the Koonenberry region, as suggested by Webby (1978), although this is
unproven. Sedimentation continued until interrupted by the Benambran/Alice Springs (1) Orogeny
which caused uplift and associated regression of the Larapintine Sea (Webby, 1978; Gravestock and
Gatehouse, 1995). Shallow marine platform sedimentation of this age also occurs within the Georgina
Basin (e.g., Greene, 2010; Draper, in Fergusson and Henderson et al., 2013).

Contemporaneous shallow marine and terrestrial sedimentation also occurred over Delamerian rocks
of western Tasmania (Seymour and Calver, 1995), but not western Victoria.

132 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.14 Generalised (approximate) distribution of deformation (cross-hatching) within the Benambran
Orogeny (ca. 455 Ma to ca. 430 Ma) as defined by outcrop, drill core and extrapolation; actual extent of
deformation may be greater and more continuous. The intensity of deformation is variable within the areas
indicated.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 133


The history of the New England Orogen in the Late Ordovician and Silurian is fragmentary, but
probably involved at least some periods of convergent margin activity (Cawood and Leitch, 1985). In
the southern NEO, Late Ordovician rocks are recorded both west and east of the Peel-Manning Fault
(Figure 3.13). To the west of the fault, arc-derived sediments in fault blocks, with blocks of limestone
(Cawood, 1976), lie unconformably on Lower Ordovician and Cambrian strata (Cawood and Leitch,
1985). Siliciclastic sedimentary rocks and fossiliferous limestones (Hall, 1975) occur within the
imbricate zone of the fault system, and include the Trelawney beds and Haedon Formation (Gamilaroi
Terrane). East of the fault, Upper Ordovician limestone is recorded below the Silverwood Group (Wass
and Dennis, 1977). Late Ordovician coral limestone, probably representing partly accreted seamounts,
was deposited in an ocean basin environment and is now incorporated into the Woolomin Terrane
(435–428 Ma; Hall, 1978). Middle to Upper Ordovician rocks of the Watonga Formation at Port
Macquarie have been interpreted to have accumulated on an oceanic plate during its passage from
spreading ridge to trench (Och et al., 2007).

Widely distributed post-Delamerian intrusive magmatism occurred across northern- and central-
eastern Australia in this cycle (Figure 3.13). In north Queensland, ca. 490 Ma to ca. 455 Ma (e.g.,
Hutton et al., 1997), dominantly felsic I-type, and mafic, (mantle-derived) magmatism comprises the
Macrossan Province (Bain and Draper, 1997). These magmatic rocks are best represented in the
Charters Towers region. Intrusive magmatism of this age has also recently been confirmed for the
Thomson Orogen (e.g., Draper, 2006; Cross et al., 2015a; Kositcin et al., 2015a), although the actual
extent is uncertain (Murray, 1994). The Ordovician magmatism in north Queensland is temporally and,
at least partly spatially, associated with extensional deformation and low-P high-T metamorphism
documented by Fergusson et al. (2007a, b) for the volcanic and sedimentary successions in the
Greenvale and Charters Towers regions. Fergusson et al. (2007a, b) suggested this extension was
related to backarc development. Such a scenario for the southern Charters Towers region, suggests
that the Macrossan Province magmatism in the northern part of the region may represent the actual
magmatic arc (e.g., Henderson, 1980; Figure 3.15). The only recorded older intrusive magmatism in
southeastern Australia (apart from the Macquarie Arc) consist of post-tectonic A- and I-type
magmatism in the southern Delamerian Orogen, e.g., in the Glenelg Zone, Victoria, and South
Australia (Foden et al., 2006).

Younger, syn- (to post-) Benambran (ca. 435 Ma and younger) intrusive magmatism occurs
throughout eastern Australia, and represents the first manifestation of the voluminous Silurian to
Devonian magmatism present within the Lachlan, Thomson and north Queensland orogens (e.g.,
Chappell et al., 1988; Murray, 1994; Bain and Draper, 1997; Figure 3.13). This includes the
dominantly I-type magmatism in north Queensland (Georgetown and possibly the Charters Towers
region; Withnall et al., 1997a; Hutton et al., 1997; Fergusson et al., 2007a), Thomson Orogen (ca. 430
Ma; Draper, 2006), and largely S-type magmatism of early Silurian age in the Central and Eastern
Lachlan Orogen (e.g., Collins and Hobbs, 2001; Ickert and Williams, 2011; Bodorkos et al., 2013,
2015; Chisolm et al., 2014b, c; Fraser et al., 2014; Figure 3.13).

Early and Middle Ordovician high-pressure blueschist metamorphism is recorded in Cambrian to


Lower Ordovician rocks at Port Macquarie (ca. 469 Ma Fukui, 1991; Watanabe et al., 1993; Fukui et
al., 1995; Offler, 1999) and along the Peel-Manning Fault (482–467 Ma; Fukui et al., 1995). The latter
is contemporaneous with interpreted arc-related plutonism, such as the ca. 480 Ma Attunga gabbro
(Fanning et al., 2002). Offler and Shaw (2006) also suggest Late Ordovician arc magmatism (e.g., ca.
445 Ma for hornblende gabbro of calc-alkaline affinity). Similar ages of ca. 436 Ma for tonalite of the
Pola Fogal Suite (Kimbrough et al., 1993) supports suggestions of an early Silurian island arc (Korsch
et al., 1990).

134 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Sedimentation and volcanism (but not intrusive magmatism) in eastern Australia largely ended with
the early Silurian Benambran Orogeny (Figure 3.14). In the Lachlan, the Benambran Orogeny (e.g.,
VandenBerg et al., 2000; Glen et al., 2007b) appears to have occurred in two main pulses, ca. 440
and 430 Ma (e.g., Glen et al., 2007b). Deformation associated with the Benambran Orogeny
commenced in the Western Lachlan, e.g., Stawell and Bendigo zones, ca. 455–440 Ma (VandenBerg
et al., 2000; Gray et al., 2003; Gray and Foster, 2004) and ca. 440 Ma (and slightly older) in other
parts of the Lachlan (e.g., Collins and Hobbs, 2001; Gray et al., 2003; Glen et al., 2007b). The
resulting Benambran Orogeny affected most of the Lachlan, with the exceptions of the Melbourne
Zone and all of Tasmania (VandenBerg et al., 2000; Seymour and Calver, 1995, 1998). The
deformation was accompanied by large amounts of shortening, crustal thickening, uplift, and regional
metamorphism (e.g., Gray, 1997; Fergusson, 2003). It also marked the end of recorded volcanism in
the Macquarie Arc (Crawford et al., 2007a). Significant higher grade metamorphism (e.g., Gray, 1997;
Gray et al., 2003) as well as syn-tectonic S-type magmatism (e.g., Collins and Hobbs, 2001),
accompanied the orogeny; both are concentrated in two (non-parallel) belts, in the Central (the
Wagga-Omeo metamorphic belt), and Eastern Lachlan.

Many of the north Queensland rocks were deformed in the early Silurian by a shortening event
coupled with metamorphism – called the Benambran Orogeny by Fergusson et al. (2007a). Evidence
for this deformation is found in the Georgetown region, dated at ca. 430 Ma, synchronous with I-type
magmatism (Withnall et al., 1997a; Fergusson et al., 2007a), although it may, in part, be younger (ca.
400 Ma; Withnall et al., 1997a). A similar deformation is recorded in the Charters Towers region,
possibly at ca. 440 Ma (Fergusson et al., 2007b). Contractional deformation, related to the Benambran
Orogeny(?), is present within the Hodgkinson and Broken River Provinces, but appears to be earlier –
probably Late Ordovician. Fergusson et al. (2007a) suggested that island-arc terranes within the
Camel Creek Subprovince (Broken River Province) were accreted at this time, which they correlated
with the Benambran Orogeny (Figure 3.13, Figure 3.15). Garrad and Bultitude (1999) have suggested
there may be a time break between Ordovician and Silurian rocks in the Hodgkinson Province that
corresponds to uplift (Benambran Orogeny) in the Georgetown region to the west.

The early Silurian contractional Benambran Orogeny is recorded in the Warburton Basin, Koonenberry
Belt, Georgina Basin (e.g., Greene, 2010; Maidment et al., 2007, part of the Alice Springs Orogeny)
and possibly the eastern Thomson Orogen (Gatehouse, 1986; Withnall et al., 1995; Gilmore et al.,
2007). Based on metamorphic ages, it is also presumed that the orogeny was experienced by rocks of
the central Thomson basement (e.g., Draper, 2006), although the areal extent of the deformation
cannot be resolved. The Benambran Orogeny in the New England Orogen apparently coincided with
the hiatus between Middle Ordovician and Lower Devonian strata at the base of the Tamworth Group
(Cawood and Leitch, 1985).

Overall, eastern Australia during the Benambran cycle appears to record both a relatively simple
passive-margin to deep marine environment and an oceanic arc environment. These are often
depicted as a back-arc/marginal basin behind an oceanic arc and west-dipping slab (e.g., Glen et al.,
1998; Li and Powell, 2001; Cayley et al., 2002; Fergusson, 2003; Gray et al., 2003; Gray and Foster,
2004; Glen, 2005; Figure 3.1, Figure 3.3 to Figure 3.5). Most tectonic models for eastern Australia
broadly concur especially for the period prior to the onset of the Benambran Orogeny. The main
exceptions to this are interpretations that suggest a non-arc environment for the Macquarie Arc (e.g.,
Wyborn, 1992; Quinn et al., 2014).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 135


Figure 3.15 Interpreted tectonic model for eastern Australia during the latest Cambrian to early Silurian (ca. 490
Ma to ca. 430 Ma). Refer to text for detailed discussion. Location of the Melbourne Zone and Tasmania in this
time period is uncertain, as are the relative positions of the West, Central and Eastern Lachlan.

136 Geodynamic Synthesis of the Phanerozoic of eastern Australia


One of the main reasons for the plethora of tectonic models for the southern Tasman Orogen (e.g.,
Coney, 1992; Wyborn, 1992; Gray, 1997; Soesoo et al., 1997; VandenBerg et al., 2000; Collins and
Hobbs, 2001; Cayley et al., 2002; Willman et al., 2002; Cas et al., 2003; Fergusson, 2003; Gray et al.,
2003; Spaggiari et al., 2003, 2004; Glen, 2005, 2013; Glen et al., 2007b; Cayley, 2011; Gibson et al.,
2011, 2015; Moresi et al., 2014), reflect the complex present-day arrangement of geological blocks
within the Lachlan Orogen, e.g., across-orogen variations in magmatism, metamorphism (especially
regions of higher grade), deformation (especially changes in structural vergence), the presence of
blueschists, the fragmented sections of the Macquarie Arc, bounded on both sides the Ordovician
turbidites, as well as the actual width of the orogen. Most workers agree that the present-day
arrangement was largely in place by the Early Devonian, and perhaps earlier. Certainly within the
Central and Eastern Lachlan, post-Benambran Orogeny extension developed within and across the
Ordovician turbidite successions and the Macquarie Arc remnants in New South Wales (Meakin and
Morgan, 1999; Lyons et al., 2000; Glen et al., 2007b).

As a result, tectonic models have to explain the evolution from passive-margin and an oceanic arc
environment in the early half of the Ordovician, to the present-day arrangement of blocks, in place by
the Early to Middle Devonian. Suggested tectonic models can be broadly broken into in-situ style
models dominantly invoking a complex arrangement of multiple subduction zones (plus or minus
subsequent shuffling) to modified models invoking the one subduction zone with complex subsequent
rearrangement of blocks. Many models for the Lachlan Orogen invoke a number of ephemeral
subduction zones (e.g., Gray, 1997; Soesoo et al., 1997, Collins and Hobbs, 2001; Fergusson, 2003;
Spaggiari et al., 2003, 2004; Aitchinson and Buckman, 2012; Figure 3.4), to explain the across-orogen
variations. These contrast with the tectonically ‘simpler’ single subduction models of VandenBerg et al.
(2000), Cayley et al. (2002) and Willman et al. (2002; Figure 3.3), Glen et al. (2011), Ferguson et al.
(2013), amongst others. These include the model of Quinn et al. (2014), who although interpret the
Macquarie Arc as a marginal basin rift, place it behind an arc to the east. A more complex variant of
the single subduction zone model has been recently proposed by Cayley and co-workers (e.g.,
Cayley, 2011; and more fully in Moresi et al., 2014). Their model (Figure 3.5) proposes a period of
oroclinal readjustment (and significant rotation of blocks) in the southern Tasman Orogen, following
collision of the Selwyn Block/VanDieland microcontinent with southeastern Australia. This and other
models are discussed further in the following section.

In summary, it can be seen that the Benambran Orogeny forms the beginnings of a period of (tectonic
and/or subsequent structural) complexity that resulted in the present-day complex arrangement of
geological terranes in those regions. This is especially so for eastern New South Wales and Victoria.
Interpreted accretion in the Early Silurian includes the Macquarie Arc terrane and elements of the
Adaminaby Superterrane (Glen, 2005; Glen et al., 2007b) and the Benambra Terrane (Willman et al.,
2002), as well as possibly the Narooma Terrane (Glen, 2005). This period of complexity is probably
also the case in north Queensland, and perhaps in the southern and eastern Thomson Orogen
(Figure 3.15). In the north Queensland region, it has been suggested that calcalkaline rocks in the
Broken River Province represent island arc remnants accreted during the Benambran Orogeny
(Fergusson et al., 2007a). Also, the interpreted backarc volcanic rocks in the Charters Towers region
and Greenvale Subprovince have quite different present-day orientations–-east-west versus NNE-
SSW, respectively (e.g., Bain and Draper, 1997; Figure 3.13, Figure 3.15). This suggests either later
relative movement between the regions, due to deformation (e.g., Bell, 1980; Fergusson et al., 2007a)
and/or perhaps the volcanism formed independently on different crustal fragments. Possible island arc
successions also occur in the eastern Thomson Orogen (e.g., Withnall et al., 1995). These rocks
include fault-bounded serpentinites (Withnall et al., 1995), consistent with accretion.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 137


Figure 3.16 A speculative tectonic reconstruction for the Tasman Orogen in latest Cambrian to early Silurian time.
The inferred extant blocks of the Thomson Orogen are shown reassembled as a possible microcontinent block,
lying somewhere to the northeast of Australia at this time. The Selwyn Block/VanDieland microcontinent is now
immediately south of Australia, prior to collision (e.g., Cayley, 2011; Moresi et al., 2014). Speculative extension is
shown forming the Larapintine seaway, possibly related to a triple junction. It is possible that the subduction
zones shown for southern and northern Australia are part of the one system. Note that this figure assumes little
relative movement between northern and southern Australia since the late Cambrian.

138 Geodynamic Synthesis of the Phanerozoic of eastern Australia


The possible occurrence of arc-related rocks in the southern Thomson Orogen also suggests
accretion, especially if they represent an oceanic arc, as suggested by Watkins (2007). It is, possible
however, that they (assuming they are Ordovician) represent an in-situ arc, and some models (e.g.,
Gray and Foster, 2004), do invoke an east-west arc at this time (Figure 3.1).

We have shown a speculative subduction-based model (Figure 3.15, Figure 3.16) which invokes a
subduction zone and arc in northern and southern Australia (possibly part of the one system), and an
additional offshore south (now represented by accreted fragments in the New England Orogen). We
have also shown a speculative ephemeral rift related to the formation of the Larapintine Sea (e.g.,
Webby, 1878), which may or may not form part of a triple junction and ephemeral arc in the southern
Thomson Orogen region (as also suggested by Gray and Foster, 2004). This arc may join with those
suggested for the southern and northern Queensland.

3.2.4 Middle Silurian to Late Devonian (ca. 430 Ma to ca. 380 Ma)
This time period (Post-Benambran to Tabberabberan Orogeny) is marked by widespread extensional
episodes with accompanying basin formation and widely distributed extrusive and intrusive magmatism.
The extension is commonly thought to have been related to significant arc rollback after the Benambran
Orogeny (e.g., Glen et al., 2004; Spaggiari et al., 2004), possibly accompanying orocline formation
(Moresi et al., 2014; Figure 3.5). Two orogenies––the Bindian and Tabberabberan––appear to break the
extension into two episodes, though the Bindian Orogeny is not well defined. Both orogenies are
recorded in most regions, with the only significant exceptions being the Melbourne Zone and Tasmania
(i.e., the Selwyn Block of Cayley et al., 2002). Significant accretion and amalgamation is also inferred to
have occurred during the Tabberabberan Orogeny. This is most evident in the Lachlan Orogen, where,
for example, differing geological histories are apparently observed between the Western Lachlan
(Whitelaw Terrane) and the Central and Eastern Lachlan (Benambra Terrane), suggesting these regions
(terranes of VandenBerg et al., 2000) were separate for most of this time period (e.g., Gray and Foster,
1997, 2004; VandenBerg et al., 2000; Willman et al., 2002; Spaggiari et al., 2004; Figure 3.1, Figure 3.3,
Figure 3.4). The recent continental accretion model of Cayley and co-workers (Cayley, 2011; Moresi et
al., 2014) suggests an alternative model, namely that these terranes are lateral equivalents modified by
significant rotation and translation and accompanying extension, processes, after indentation by the
Selwyn Block/VanDieland microcontinent (Figure 3.5).The probable accretion of the Calliope-Gamilaroi
island arc terranes to the Gondwana margin in the early Late Devonian within the New England Orogen
are also suggested to have occurred during this orogeny, providing an additional complexity.

Within the Central and Eastern Lachlan, post-Benambran extension developed within and across the
Ordovician turbidite successions and the Macquarie Arc remnants in New South Wales (Meakin and
Morgan, 1999; Lyons et al., 2000; Glen et al., 2007b), and within Ordovician turbidite successions in
Victoria (e.g., VandenBerg et al., 2000). This resulted in widespread, largely deep to shallow marine
sedimentation, including carbonates, in various basins (Pogson and Watkins, 1998; VandenBerg et
al., 2000; VandenBerg, 2003; Colquhoun et al., 2005; Glen, 2005; Lyons et al., 2000; Meakin and
Morgan, 1999; Fergusson, 2010; Figure 3.17). Accompanying magmatism included S- and I-type
granites and is often bimodal (Lyons et al., 2000; Collins and Hobbs, 2001, Chappell and White; 1992;
Rossiter, 2003; Black et al., 2005; Gray et al., 2003). These basins were inverted during the poorly-
defined, diachronous(?), latest Silurian–Early Devonian Bindian Orogeny (ca. 420–400 Ma; e.g.,
Perkins et al., 1994; VandenBerg et al., 2000; Gray et al., 2003). Renewed extension and
development of rift basins continued into the Early Devonian in the Central and Eastern Lachlan
following the Bindian Orogeny (e.g., Willman et al., 2002; Fergusson, 2010). This resulted in deep to
shallow marine sedimentation (including carbonates) and widespread bimodal or felsic volcanism in
new and existing basins in the Central and Eastern Lachlan in both Victoria and New South Wales

Geodynamic Synthesis of the Phanerozoic of eastern Australia 139


(Meakin and Morgan, 1999; Lyons et al., 2000; VandenBerg et al., 2000, Willman et al., 2002;
Colquhoun et al., 2005; Glen, 2005; Fergusson, 2010).

Sedimentation in the Western Lachlan was apparently confined to the Melbourne Zone of Victoria
(VandenBerg et al., 2000; Figure 3.17), where largely deep marine sedimentation is recorded. Similar
extensive deep marine sedimentation occurred in northeastern Tasmania (Seymour and Calver,
1995). The Bindian Orogeny appears to be absent from northeastern Tasmania and the Melbourne
Zone of Victoria (Seymour and Calver, 1995; VandenBerg et al., 2000). As a result, sedimentation in
both regions is largely continuous throughout this cycle, continuing from Benambran times (e.g.,
VandenBerg et al., 2000; Fergusson et al., 2003; Seymour and Calver, 1995). In the Melbourne Zone,
sedimentation shallows upward to terrestrial sedimentation at the top, and contains evidence for a
change in sediment transport direction with the appearance of lithic and volcaniclastic detritus derived
from the east (VandenBerg et al., 2000; VandenBerg, 2003). Willman et al. (2002) and VandenBerg
(2003) suggested the new source provenance reflects the arrival of the Benambra Terrane, that is, the
Benambra and Whitelaw terranes were separate prior to this––a conclusion agreed upon by many
workers, regardless of tectonic model (e.g., Gray, 1997; Gray and Foster, 1997, 2004; Fergusson,
2003, Spaggiari et al., 2004; Moresi et al., 2014).

Widespread felsic-dominated magmatism occurs across the Western, Central and Eastern Lachlan within
the Tabberabberan cycle (Lyons et al., 2000; Collins and Hobbs, 2001, Chappell and White; 1992;
Rossiter, 2003; Black et al., 2005; Gray et al., 2003; Ickert and Williams, 2011; Bodorkos et al, 2013, 2015;
Chisholm et al., 2014a,b; Fraser et al., 2014). Ages largely fall between ca. 430 and 390 Ma but continued
into the Kanimblan cycle (e.g., Chappell and White, 1992; Gray et al., 2003; Black et al., 2005; Ickert and
Williams, 2011). The oldest granites in New South Wales and Victoria are dominantly S-types in the Central
and Eastern Lachlan (e.g., Collins and Hobbs, 2001; Willman et al., 2002, Ickert and Williams, 2011;
Bodorkos et al., 2013, 2015), including a continuation of dominantly S-type magmatism that commenced
during the Benambran Orogeny (e.g., Collins and Hobbs, 2001; Figure 3.14, Figure 3.17). Early Silurian
magmatism appears to be absent from the Western Lachlan (Whitelaw Terrane) in Victoria (VandenBerg et
al., 2000; Willman et al., 2002). Magmatic ages of ca. 435–415 Ma are widespread across the Central and
Eastern Lachlan Orogen and continue northward into the Thomson Orogen (Figure 3.17).

Granite ages appear to show a variety of diachronous trends (Figure 3.17), though are only definitive in
Tasmania where most granites have been dated. In the Eastern Lachlan of New South Wales and Victoria,
ages appear to decrease eastwards, from 435–415 Ma, to 415–400 Ma, to ca. 380–360 Ma (Lewis et al.,
1994; VandenBerg et al., 2000) in the east, possibly reflecting arc rollback. It is noted, however, that older
granites (ca. 400 Ma) are present in the east also (Bodorkos et al., 2015), and that the younger magmatism
(415–360 Ma) occurs within the Eastern and Central Lachlan (and also the adjoining Thomson Orogen). In
Victoria, granites mostly decrease in age towards the Melbourne Zone (VandenBerg et al., 2000; Gray et
al., 2003; Rossiter, 2003). Granites east and west of the Central Victorian Magmatic Province are largely
ca. 420–380 Ma in age (VandenBerg et al., 2000; Gray et al., 2003). The youngest rocks occur within the
post-Tabberabberan Middle Devonian to Early Carboniferous (ca. 385–350 Ma) Central Victorian
Magmatic Province (VandenBerg et al., 2000; Rossiter, 2003). A similar diachronous trend is evident in
Tasmania, where granite ages record a pronounced westward younging from ca. 400–375 Ma, pre-, syn-,
and post-tectonic granites in the northeast to post-tectonic granites, ca. 370–350 Ma, in western Tasmania
(Black et al., 2005; Figure 3.17).

Silurian to Mid Devonian in the southern Delamerian Orogen include terrestrial to marine
sedimentation in western Victoria, largely in the Grampians-Stavely Zone (VandenBerg et al., 2000),
and deep-water, clastic-dominated sedimentation in Western Tasmania (Seymour and Calver, 1995,
1998; Figure 3.17). Sediments in Western Victoria were apparently deformed at ca. 420–410 Ma
(Bindian Orogeny?) and are overlain by post-deformation Early Devonian volcanic rocks (VandenBerg

140 Geodynamic Synthesis of the Phanerozoic of eastern Australia


et al., 2000) with associated plutonism. The Bindian Orogeny appears to be absent in western
Tasmania, although hiatuses which may correspond to this orogeny are recorded (Seymour and
Calver, 1995, 1998) and there is lode gold mineralisation of this age (Bierlein et al., 2005).

The Silurian to Mid Devonian in north Queensland is characterised by extensive, pre- and post-
Bindian, sedimentation in the Hodgkinson and Broken River Provinces (along the eastern and
southeastern margins of the Proterozoic Etheridge Province), post-Bindian sedimentation in the
Charters Towers region, and minor pre-Bindian sedimentation in the Georgetown region (Figure 3.17).
Sedimentation in these regions includes marine siliciclastic sediments and carbonates, and, locally
abundant, pre-Bindian tholeiitic mafic volcanism (Arnold and Fawckner, 1980). Sediment provenance
is dominantly cratonic (e.g., Bultitude et al., 1997) but does include volcaniclastic material, some of
which is older (for example, Garrad and Bultitude (1999) record dacitic clast ages of 465 Ma in the
Hodgkinson Province). The geodynamic setting for this sedimentation is controversial. Most models
invoke either a backarc or a forearc setting (see summaries in Arnold and Fawckner (1980) and
Garrad and Bultitude (1999), and more recently in Henderson et al., 2013). Arnold (in Arnold and
Fawckner, 1980), Henderson et al. (1980), Henderson (1987) and Korsch et al. (2012), amongst
others, suggested that the Hodgkinson and Broken River Province sedimentation was part of a forearc
basin and accretionary wedge. Resolution of the geodynamic setting is critical to understanding the
tectonics of the widespread Pama Province magmatism. The latter forms an extensive quasi-
continuous belt around the Hodgkinson and Broken River Provinces, from Charters Towers in the
south, north to Cape York. Pama Province magmatism in the region is diachronous. It ranges from ca.
430–420 Ma to 406 Ma (syn- and post-Benambran) in the Georgetown region (e.g., Withnall et al.,
1997a; Murgulov et al., 2007, 2009), to ca. 425–405 Ma (and younger) in the Charters Towers region
(Hutton et al., 1997), to ca. 410–395 Ma in the Coen region (Black et al., 1992). As pointed out by
Champion and Bultitude (2003), these age differences are also matched by changes in geochemical
signature. Notably, early Pama magmatism (in the Georgetown region) does have geochemical
signatures more consistent with arc magmatism. It is also feasible that the region started with a forearc
setting which, with rollback, evolved into a backarc setting. The presence of ca. 380–375 Ma S-type
granites in the eastern Hodgkinson (Kositcin et al., 2015a) could be indicative of such a process, i.e.,
melting of the old forearc (though the timing of these intrusives may be more related to the
Tabberabberan deformation). Regardless of which model is correct for the Hodgkinson and Broken
River provinces, both require the presence of west-dipping subduction. This model, however, has
difficulties explaining the tholeiitic volcanism, especially in the Hodgkinson Province. The latter is more
consistent with rift or backarc models (e.g., Fawckner in Arnold and Fawckner, 1980; Bultitude et al.,
1997), though the magmatism may simply reflect either accreted oceanic crust fragments in the
accretionary wedge model or perhaps extension in the forearc.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 141


Figure 3.17 General distribution of Middle Silurian to Late Devonian (ca. 430 to 380 Ma), Tabberabberan cycle
rocks in eastern Australia. Refer to text for discussion.

142 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Sedimentation and associated extrusive and intrusive magmatism in the Thomson Orogen and
Koonenberry region also occurred within two episodes, also probably in response to extension
following the Benambran and Bindian orogenies (e.g., Olgers, 1972; Neef and Bottrill, 1991; Murray,
1994, 1997a; McKillop et al., 2007; Withnall et al., 1995; Gilmore et al., 2007; Greenfield et al., 2010).
Post-Benambran, Late Silurian to Early Devonian marine and terrestrial sediments, of cratonic and/or
volcanic provenance, are also recorded in the Koonenberry region (Neef and Bottrill, 1991), and in the
Thomson Orogen (Olgers, 1972; Murray, 1994; 1997a; Withnall et al., 1995; Figure 3.17). These are
associated with mafic and felsic volcanic rocks in the Koonenberry region (Neef and Bottrill, 1991;
Greenfield et al., 2010) and the Anakie Province (Withnall et al., 1995), and felsic intrusives in the
Thomson Orogen basement (e.g., Murray, 1994; Fraser et al., 2014; Cross et al., 2015a) and
Koonenberry region (Gilmore et al., 2007; Greenfield et al., 2010; Figure 3.17). Both Murray (1994)
and Thalhammer et al. (1998) have suggested continental settings. Widespread felsic intrusive
magmatic rocks of this age (ca. 425–405 Ma), belonging to the Pama Igneous Association (Bain and
Draper, 1997) occur within the Charters Towers region (Hutton et al., 1997; Figure 3.17).

Renewed extension, following the Bindian Orogeny, produced the Early to Late Devonian, terrestrial to
shallow marine Adavale Basin in the central Thomson Orogen (McKillop et al., 2007), terrestrial to
shallow marine sedimentation in the Burdekin Basin (Charters Towers region; Hutton et al., 1997), on-
marine sandstones in the Georgina Basin (Haines et al., 2001), and Early to Middle Devonian quartz-
rich sedimentation in the Koonenberry region (Neef, 2004). The Adavale Basin is considered to have
formed in a continental setting, possibly as an intracontinental volcanic rift (Murray, 1994; McKillop et
al., 2007). Felsic intrusive magmatism accompanied extension in the Thomson Orogen (e.g., Murray,
1994; Evans et al., 1990; Hutton et al., 1997). McKillop et al. (2007) suggest that extension may have
been the result of far field events, such as subduction further to the east (in the New England Orogen).
In contrast, sedimentation in the Georgina and Amadeus and other basins, is thought to have been
synorogenic, related to the Alice Springs Orogeny (Haines et al., 2001).

The Tabberabberan cycle in the New England Orogen is characterised by a convergent margin phase,
represented by the Late Silurian to Middle Devonian Gamilaroi-Calliope island arc and an enigmatic
collisional phase (Tabberabberan Orogeny), associated with the possible accretion of these island arc
terranes in the Middle–Late Devonian (e.g., Offler and Murray, 2011). Elements of a Late Silurian to
Middle Devonian arc occur in fault-bounded blocks along the western flank of the Peel-Manning Fault
in the southern New England Orogen, and in the vicinity of the Yarrol Belt in the northern part of the
orogen (Figure 3.17). The arc successions consisting of volcaniclastic, extrusive and intrusive rocks of
the Gamilaroi Terrane and Calliope Arc, are interpreted to have formed as a single intraoceanic island
arc (low-K calc-alkaline signature) (van Noord, 1999; Offler and Gamble, 2002; Murray and Blake,
2005), though Morand (1993a) suggested a primitive continental arc environment was possible.
Volcanic rocks in the Gamilaroi Terrane lie below the Middle Devonian–Carboniferous strata of the
Tamworth Belt. The Calliope Arc succession, host to the Mount Morgan gold-copper deposit (~380
Ma), is overlain by strata of the Yarrol Belt and is intruded by the Mount Morgan Trondhjemite (381 ± 5
Ma; Golding et al., 1994), which has an arc or rifted-arc geochemistry (Murray, 2003). Cawood and
Flood (1989) and Offler and Gamble (2002) have suggested that the arc developed above a west-
dipping subduction zone, although Aitchison and Flood (1995) argued for east-dipping subduction, as
did Murray (2007). The Silverwood Group and Willowie Creek beds have been interpreted as a
southern continuation of the Calliope Arc (Day et al., 1978). Island arc magmatism is also recorded
along the Peel-Manning Fault (early Silurian ca. 425 Ma hornblende cumulates and diorites; Sano et
al., 2004) and in the Marlborough Block (Middle Devonian). In the latter, calcalkaline basalts, dolerites
and gabbros and fault-bounded blocks in the Marlborough ophiolite (380 ± 19 Ma) have trace element
data suggestive of an intra-oceanic island arc (Bruce and Niu, 2000). The arc, probably built on
Neoproterozoic oceanic crust, lay only a short distance offshore and was accreted to Gondwana

Geodynamic Synthesis of the Phanerozoic of eastern Australia 143


(Bruce and Niu, 2000; Murray, 2007). Late Silurian to Devonian ocean basin sedimentation is also
recorded in the New England Orogen, in the Woolomin Terrane (428–380 Ma chert and basalt;
Aitchison et al., 1992), and in the Coastal Block (e.g., ca. 418 Ma to ca. 360 Ma Doonside Formation;
Fergusson et al., 1993; Figure 3.17). Offler and Murray (2011) have suggested the presence of two
arcs during this period: a west-dipping subduction zone facing mainland Australia and an east-dipping
subduction zone and oceanic arc offshore. The presence of the west-dipping subduction would be
consistent with suggested settings for the Mossman Orogen (e.g., Henderson et al., 2013).

The Bindian Orogeny (ca. 420–400 Ma), although largely poorly defined and variably developed,
appears to be present within the Eastern and Central Lachlan, Delamerian, Thomson and Mossman
orogens, but not the New England Orogen (Figure 3.18). The orogeny is most evident in the Central
and Eastern Lachlan in Victoria and New South Wales, where it has been suggested to be
transpressive and to have resulted in significant strike-slip movement between the western and central
Lachlan (VandenBerg et al., 2000; Willman et al., 2002; Glen, 2005), with possibly up to 600 km of
dextral movement (e.g., Willman et al., 2002; Figure 3.3). Other authors have questioned this, and it is
possible that deformation of this age may relate more to continuing subduction-accretion effects if the
multiple subduction zone models of Gray (1997), Gray and Foster (1997), Soesoo et al. (1997),
Spaggiari et al. (2003, 2004; Figure 3.4), or the indenter model of Moresi et al. (2014), are correct.
This does not necessarily negate strike-slip effects in the Central and Eastern Lachlan. Bindian
deformation in far eastern Victoria appears to relate more to east-west contraction (e.g., Willman et al.,
2002). Despite being absent in the Western Lachlan (Willman et al., 2002), Bindian-aged deformation
may be present in parts of the Delamerian Orogen, for example, the Tabberabberan cycle sediments
in Western Victoria were deformed at ca. 420–410 Ma (VandenBerg et al., 2000). The Bindian
Orogeny appears to be absent in western Tasmania, although hiatuses which may correspond to this
orogeny are recorded (Seymour and Calver, 1995, 1998) and lode gold mineralisation at Beaconsfield
is reportedly also of this age (Bierlein et al., 2005).

Local folding and metamorphism, probably with associated felsic magmatism, took place in the
Koonenberry region and the Thomson Orogen during the Bindian Orogeny (e.g., Thalhammer et al.,
1998). The orogeny there may have been diachronous. Deformation in the Koonenberry region is
recorded as Late Silurian to Early Devonian (Gilmore et al., 2007), while it appears to be late Early
Devonian in the Thomson Orogen, constrained by ca. 408 and 402 Ma volcanic rocks in the post-
Bindian Adavale Basin (McKillop et al., 2007). Like the Lachlan Orogen, ca. 420–400 Ma magmatism
appears to be widely distributed (Greenfield et al., 2010; Fraser et al., 2014; Cross et al., 2015a).

Bindian-aged deformation, ca. 410–400 Ma, is also recorded in the Georgetown, Coen and Charters
Towers regions of north Queensland, where it coincides with part of the extensive Pama Province
magmatism in those regions (Figure 3.18). Bultitude et al. (1997) record a change in sedimentation in
the Hodgkinson Province in the Late Lochkovian (ca. 412 Ma), and Withnall et al. (1997b) record a
hiatus in sedimentation at this time in the Graveyard Creek Subprovince. Both suggested these
changes were related to hinterland uplift (due to Bindian orogeny?). Sedimentation also appears to
recommence at this time in the Charters Towers region.

The Tabberabberan Orogeny affected large parts of eastern Australia (Figure 3.19). Within the
Lachlan Orogen, the ca. 390–380 Ma dominantly east-west contractional Tabberabberan Orogeny
(e.g., Gray and Foster, 1997, 2004; Spaggiari et al., 2003, 2004), effectively cratonised the whole
orogen, and has been interpreted to have been responsible for amalgamation of the terranes and
zones of the Lachlan (Gray and Foster, 1997, Soesoo et al., 1997; VandenBerg et al., 2000; Willman
et al., 2002; Spaggiari et al., 2003, 2004; Seymour and Calver, 1995; Black et al., 2005).

144 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.18 The generalised (approximate) distribution of deformation (cross-hatching) during the Bindian
Orogeny deformation (ca. 420 to 400 Ma) as defined by outcrop and drill hole information––actual extent of
deformation may be greater and more continuous. The intensity of deformation is variable within the areas
indicated. See text for data sources. Refer to Figure 3.17 for additional geological explanations.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 145


Figure 3.19 The generalised (approximate) distribution of deformation (cross-hatching) during the Tabberabberan
Orogeny (ca. 390 to 380 Ma) as defined by outcrop and drill hole information––actual extent of deformation may
be greater and more continuous. The intensity of deformation is variable within the areas indicated. See text for
data sources. Refer to Figure 3.17 for additional geological explanations.

146 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.20 Interpreted tectonic model for the Middle Silurian to Late Devonian period (ca. 440 to 380 Ma) of
eastern Australia. Refer to text for detailed discussion.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 147


Figure 3.21. Speculative tectonic reconstruction of the Tasman Orogen for the Middle Devonian, just prior to the
inferred collision of the Gamilaroi-Calliope arc and mainland Australia. During this time period the Lachlan,
Thomson and Mossman orogens were assembled into the present day configuration. The processes controlling
the reorganisation, and the degree of reorganisation, are uncertain and are thus not shown. Two, opposite facing
offshore subduction zones are shown (following Offler and Murray, 2011). The necessity for both is largely
dependent on the interpretation of the Gamilaroi-Calliope arc as either oceanic or continental. Note that this figure
assumes little relative movement between northern and southern Australia since the Late Cambrian.

148 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Although dominantly east-west, a component of north-south contraction is also recorded as part of this
orogeny within the Lachlan Orogen (e.g., VandenBerg et al., 2000). This may reflect oblique terrane
amalgamation (e.g., Spaggiari et al., 2003) or final amalgamation of the Selwyn Block with Australia
(Cayley, 2011; Moresi et al., 2014)

Deformations around this age in north Queensland are best represented in the Broken River and
Hodgkinson provinces and also in the Charters Towers region, where they include time breaks and
slight angular unconformities (Withnall and Lang, 1993; Bultitude et al., 1997; Figure 3.19). Henderson
(1987) suggested that this event resulted in the cessation of deep-marine sedimentation in the Camel
Creek Subprovince, and produced the Middle Devonian angular unconformity observed in the
Graveyard Creek Subprovince. Garrad and Bultitude (1999) recorded east to north-east thrusting and
north-northwest-trending shear zones in the Hodgkinson Province, which they suggested were of Late
Devonian age, possibly related to basin inversion. Given the strong commonalities between the
Broken River and Hodgkinson Provinces it is possible deep-water marine sedimentation ceased
simultaneously in both. Granites in the eastern Hodgkinson Province have been recently dated as ca.
380–375 Ma (Kositcin et al., 2015a); their generation may also relate to, and date, the Tabberabberan
deformation in that region (e.g., Zucchetto et al., 1999). The Tabberabberan Orogeny in the
Koonenberry region resulted in east-northeast–west-southwest contractional deformation at ca. 395
Ma (Mills and David, 2004; Neef, 2004). Deformation of this age in the Thomson Orogen (often called
the Alice Springs Orogeny (2)) is either poorly developed or difficult to distinguish from other events
(Withnall et al., 1995; Hutton et al., 1997). In the Adavale Basin, the orogeny appears to have
produced an unconformity between terrestrial and overlying shallow marine sedimentary rocks, and a
possible change to restricted basin conditions in the late Middle Devonian (McKillop et al., 2007).

The Tabberabberan Orogeny is also recorded in the New England Orogen, although there is no
general agreement as to its effects and timing. In the southern New England Orogen, Flood and
Aitchison (1992) suggested that the Calliope-Gamilaroi intraoceanic arc accreted to the Australian
plate in the Late Devonian, based on the first appearance of distinctive, westerly-derived quartzite
clasts in the uppermost Devonian Keepit Conglomerate of the Tamworth Group, considered to have
been derived from the Lachlan Orogen. This is controversial, however, and according to Glen (2005),
no clear evidence of contractional deformation related to arc accretion has been identified. In the
northern NEO, supporters of the intraoceanic model for the Calliope Arc argue that an unconformity
close to the Middle–Late Devonian boundary reflects accretion of the arc to the Gondwana margin
(e.g., Murray et al., 2003). In contrast, others (e.g., Leitch et al., 1992; Morand, 1993b; Bryan et al.,
2003) have argued that the unconformity is low angle, deformation was minor and that there was no
break at the base of the overlying Yarrol Belt. If accretion of an island arc did occur (e.g., Offler and
Murray, 2011), then the ages of tonalitic rocks in the northern NEO (ca. 380–370 Ma; see Purdy, in
Donchak et al., 2013) suggest it may have been slightly later (Middle–Late Devonian) than often
suggested, e.g., ca. 388 Ma in Tasmania (Black et al., 2005). The NEO ages are similar to those of
the syn-tectonic S-type granites in the Hodgkinson Province of north Queensland.

Overall, it would appear that the ca. 430 to 380 Ma period records a relatively simple overall backarc
environment behind a subduction zone to the east most likely within the New England Orogen (e.g.,
Gray, 1997; Glen et al., 1998; Cayley et al., 2002; Gray and Foster, 2004; Glen, 2005; 2013;
Fergusson, 2010; Offler and Murray, 2011; Donchak et al., 2013; Figure 3.20). As for the previous
time period, however, it is evident that there is unresolved tectonic complexity required to explain the
present day geometry of the Tasman Orogen––a geometry most agree was largely in place before or
by the end of the Tabberabberan Orogeny. This is particularly the case for the Lachlan Orogen where
a range of somewhat disparate models have been proposed. As discussed earlier, suggested models
for the Lachlan Orogen can be broadly broken into in-situ style models invoking multiple subduction
zones (plus or minus subsequent shuffling; e.g., Gray, 1997; Soesoo et al., 1997, Collins and Hobbs,

Geodynamic Synthesis of the Phanerozoic of eastern Australia 149


2001; Fergusson, 2003; Spaggiari et al., 2003, 2004; Aitchinson and Buckman, 2012; Figure 3.4), to
models invoking the one subduction zone with complex subsequent rearrangement of blocks (e.g.,
VandenBerg et al., 2000; Cayley et al., 2002; Willman et al., 2002; Glen, 2013; Fergusson et al., 2013,
amongst others; Figure 3.3). The latter include the model of Quinn et al. (2014), who interpret the
Macquarie Arc as a marginal basin rift, behind an arc to the east. A more complex variant of the single
subduction zone model was recently proposed by Cayley and co-workers (e.g., Cayley, (2011), and
more fully in Moresi et al., 2014). Their model (Figure 3.5) proposes a period of oroclinal readjustment
(and rotation of blocks) in the southern Tasman Orogen, following collision by the Selwyn
Block/VanDieland microcontinent. Each of these models better explain certain, but not all, aspects of
the Tabberabberan cycle geology. All suffer from apparent difficulties: e.g., the dominantly felsic
nature of much of the Lachlan magmatism (e.g., Chappell and White, 1992), is not consistent with
multiple subduction zones (though cf. Collins and Hobbs, 2001). Similarly, there is not a lot of
evidence for large amounts of strike slip especially along the inferred Baragwanath Transform, and
there appears to be only modest offsets observed along a number of significant other faults (e.g.,
Fergusson, 2010). The more recent model of Cayley (2011) and Moresi et al. (2014; Figure 3.5),
provides a tantalising way forward, potentially explaining the present day geometry and the evidence
for both contemporaneous extension and compression in different parts of the orogen, as suggested
by the range of ages for the poorly understood Bindian Orogeny. Their model, however, also suffers
from apparent problems. Perhaps the most obvious difficulty, based on the modelling of Moresi et al.
(2014), is the large amount of extension and ocean formation, required by the model for the region
now occupied by the Darling Basin (see Figure 2 of Moresi et al. 2014; also Early Silurian in
Figure 3.5). This is not supported by seismic data, for example, which suggests that the Darling Basin
(at least now) overlies Lachlan and Delamerian basement (R. Korsch, pers. comm., 2015). It is noted
that results of the numerical modelling by Moresi et al. (2014) may ameliorate this problem through a
combination of different parameters, e.g., differing yield strengths and lithosphere age. The Cayley
and co-workers model also requires a significant amount of north-south shortening (note the current
position of the Thomson Orogen/Lachlan Orogen boundary in Figure 3.5), with flow on effects for the
Thomson and north Queensland portions of the Tasman Orogen. This may be testable via
paleomagnetic data. The recent model of Quinn et al. (2014), not dissimilar to that of Wyborn (1992),
provides another potential solution to the Lachlan Orogen conundrum, largely by removing the
necessity to have to explain the juxtaposition of oceanic arc fragments in between the Ordovician
turbidites. It does not explain other difficulties however, e.g., the changes in structural vergence across
the Victorian part of the Lachlan Orogen (e.g., Gray and Foster, 1997, 2004).

The large areal distribution of magmatism in eastern Australia during the Tabberabberan cycle
(Figure 3.17) is also problematical and has led many authors to speculate on tectonic scenarios,
especially to explain the (present day) anomalously wide zone of magmatism in the Lachlan Orogen. It
is noted, however, that magmatism of this age in the Thomson Orogen and Koonenberry region
(Figure 3.17), although not as well understood and with much less geochronological control, also
appears to occupy a wide east-west extent, similar to the Lachlan Orogen. Recent dating of a granite
between Mount Isa and Georgetown (ca. 380 Ma; Geoscience Australia unpublished data) shows that
Devonian magmatism is also present in that area, suggesting that the wide east-west extent of this
magmatism is a feature of eastern Australia, and that backarc to behind-arc extension of previously
thickened crust (and not multiple subduction zones or other inferred tectonic processes) may perhaps
be sufficient to explain the width of magmatism in the Lachlan and elsewhere at this time. Like the
Lachlan, differing tectonic models have been advocated to explain the widespread Pama Igneous
Association magmatism in north Queensland and Charters Towers (Figure 3.17). As discussed earlier,
the resolution of this is linked to geodynamic models for the Hodgkinson and Broken River Provinces,
which have been interpreted as either forearc or backarc (Figure 3.20). The diachronous nature of the
Pama Igneous Association magmatism and the corresponding changes in geochemical signature

150 Geodynamic Synthesis of the Phanerozoic of eastern Australia


(e.g., Champion and Bultitude, 2003), may perhaps be best interpreted as a switch from an early
forearc environment (ca. 430 Ma) to a backarc environment (ca. 420–400 Ma and younger?), similar
to the model of Collins and Richards (2008) for the Lachlan Orogen.

The uncertainty is not confined to the southern or northern parts of the Tasman Orogen. There is also
argument about the nature of the arc in the NEO at this time (island versus continental arc), its
orientation (east- or west-facing) or indeed whether there may indeed been two arcs (Offler and
Murray, 2011; Figure 3.20, Figure 3.21).

Finally, interpretations for the drivers of the Bindian and Tabberabberan orogenies are also varied.
Again, most work has concentrated on the Lachlan Orogen and conclusions largely reflect the different
tectonic models already discussed. Gray and co-workers (e.g., Gray, 1997; Gray and Foster, 1997,
2004; Soesoo et al., 1997; Spaggiari et al., 2003, 2004) suggested a collisional event that was (at
least partly) related to the closure of a marginal basin (effectively the Melbourne Zone) and an end to
double divergent subduction (e.g., Gray and Foster, 1997; Soesoo et al., 1997; Figure 3.2, Figure 3.4,
Figure 3.20). Conversely, Willman et al. (2002) and Cayley et al. (2002) suggested the Bindian
Orogeny was responsible for southward transport of their Benambra Terrane, which ended in the
Tabberabberan Orogeny when the Whitelaw and Benambra terranes were amalgamated (Figure 3.3).
In the indenter model of Cayley and co-workers (e.g., Cayley, 2011), most of what is called the
Bindian Orogeny reflects the variable contemporaneous shortening and extension evident across the
orogen throughout the Silurian and Early Devonian.

Importantly, most models indicate that the Western and Central-Eastern Lachlan were not in their
current positions for much of the Tabberabberan cycle and that they were amalgamated during the
Tabberabberan Orogeny (e.g., Gray, 1997; Gray and Foster, 1997, 2004; VandenBerg et al., 2000;
Willman et al., 2002; Fergusson, 2003, Spaggiari et al., 2004). A similar argument seems valid for
Tasmania. Detrital zircon data from sediments of this age in northeastern Tasmania indicate no
apparent sourcing of material from western Tasmania, which led Black et al. (2004) to suggest that
northeastern and western Tasmania were also separate at this time, such that in Tasmania the
Tabberabberan deformation (ca. 388 Ma; Black et al., 2005) may relate to docking of northeastern
Tasmania to western Tasmania (Black et al., 2004; Figure 3.20, Figure 3.21). It would appear,
therefore, that at least part of the Tabberabberan Orogeny relates to amalgamation of the Lachlan
Orogen. It is also probable, however, that this deformation may also relate to docking of the Calliope-
Gamilaroi arc in the New England Orogen at this time, although as discussed earlier, the effects and
timing of any amalgamation are not clear cut or agreed (e.g., Flood and Aitchison, 1992; Leitch et al.,
1992; Morand, 1993a, b; Bryan et al., 2003; Murray et al., 2003; Glen, 2005). What does appear
evident, however, is that the end result of the Tabberabberan Orogeny was a crustal configuration for
eastern Australia not far removed from that we see today (Figure 3.21). In our speculative tectonic
model (Figure 3.21) we have not tried to depict the possible processes that may have been
responsible for the complex reorganisation inferred to many to have occurred in the Lachlan and other
orogens during the Silurian to late Mid Devonian. This largely reflects both the great variety of
proposed tectonic models but also the uncertainty over just how much reorganisation has actually
occurred, from very significant (e.g., Cayley, 2011; Moresi et al., 2014; Figure 3.5), to perhaps very
little (e.g., Quinn et al., 2014). In our model, the Tabberabberan Orogeny is largely driven by collision
of the Gamilaroi-Calliope arc.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 151


3.2.5 Late Devonian to early Carboniferous (ca. 380 to ca. 350 Ma)
The interpreted accretion of the Gamilaroi-Calliope island arc to the Gondwana margin in the
Tabberabberan Orogeny resulted in the initiation of the New England Orogen as an Andean-style
continental margin, with a westerly dipping subduction zone (Cawood, 1982; Murray et al., 1987), that
continued until the Kanimblan Orogeny (ca. 380–350 Ma). During this period, the now largely cratonic
Lachlan, Delamerian, Thomson and Mossman Orogens were marked by widespread extension, rifting
and accompanying basin formation, and locally significant extrusive and intrusive magmatism (e.g.,
VandenBerg et al., 2000; Lyons et al., 2000; Glen, 2005). Like earlier orogenic cycles, this extension is
thought to reflect behind-arc processes, including a backarc basin in the north, related to renewed
rollback in the subduction zone and arc within the New England Orogen (e.g., Glen, 2005; Collins and
Richards, 2008). This extension ended with the Kanimblan Orogeny (ca. 350 Ma), although the arc
itself––the Connors-Auburn Arc––continued throughout the Kanimblan cycle and into the next cycle,
from the Late Devonian to the late part of the Carboniferous (ca. 370–305 Ma; e.g., Roberts et al.,
1995, Murray, 2003). The subsequent history of the New England Orogen, therefore, records an event
history in many ways distinct from the remainder of the Tasmanides (e.g., Glen, 2005).

The Andean-style continental margin in the New England Orogen resulted in the development of a
magmatic arc in the west (Connors-Auburn Arc), flanked by a forearc basin and an accretionary
wedge in the east (Cawood, 1982; Murray et al., 1987; Figure 3.22). In the northern NEO, the
continental magmatic arc is represented by granitic and mafic to silicic rocks of the Connors-Auburn
Volcanic Arc (Murray, 1986). Although magmatism probably began earlier, the preserved record
shows that igneous activity in both regions commenced ca. 350 Ma––late in the Kanimblan Cycle––
with the main pulse of granite formation between ca. 324–313 Ma in the Auburn Subprovince and ca.
316–305 Ma in the Connors Subprovince (Murray, 2003). In the southern NEO, the magmatic arc is
now either buried or has been removed. Its presence is inferred from large Late Devonian olistromal
blocks of andesitic volcanic rocks in the inboard part of the forearc basin (Brown, 1987) and by Late
Devonian to late Carboniferous sandstones largely of arc-derived volcaniclastic composition (Cawood,
1983; Korsch, 1984). The Tamworth Belt in the southern NEO and the Yarrol Belt in the northern
NEO, as well as intervening blocks, are usually interpreted to be parts of a continuous forearc basin
(Cawood and Leitch; 1985; Korsch et al., 1990; Murray et al., 2003; Figure 3.22). The Late Devonian
to Late Carboniferous successions consist predominantly of continental to shallow marine clastic
sediments, including limestone, with a provenance predominantly from the volcanic arc to the west
(Cawood, 1983; Korsch, 1984; Yarrol Project Team, 1997). The Upper Devonian rocks of the Yarrol
Province are considered to represent the transitional change from an intraoceanic setting to a
continental margin setting in the Carboniferous (Murray and Blake, 2005). Volcanism and deposition
continued until the Late Carboniferous (Roberts et al., 2004). The Woolomin, Central and Coffs
Harbour blocks in the southern NEO and the Coastal, Yarraman, North D’Aguilar, South D’Aguilar and
Beenleigh blocks in the northern NEO are interpreted as a once continuous accretionary wedge
(Figure 3.22) that grew ocean-wards by accreting trench-fill volcaniclastic turbidites (derived from a
magmatic arc) and minor amounts of oceanic crust (e.g., Korsch et al., 1990). A backarc basin occurs
in the north (Drummond Basin, e.g., Henderson and Blake, in Ferguson and Henderson et al., 2013),
with the Late Devonian of the Lachlan Orogen being the backarc equivalent in the south.

The earliest Kanimblan cycle rocks in the Lachlan Orogen consist of Middle to Late Devonian
sediments and A-type and bimodal extrusive and intrusive magmatism (e.g., Meakin and Morgan,
1999; Lyons et al., 2000; Lewis et al., 1994; Wormald et al., 2004), probably related to initiation of
extension and associated rifting after the Tabberabberan Orogeny. This sedimentation and
magmatism was followed by Lachlan-wide (New South Wales and Victoria) Late Devonian to Early
Carboniferous, clastic, mostly continental, sedimentation, including red beds, of the ‘Lambie facies’,
reflecting continuing, more widespread extension (Lewis et al., 1994; Warren et al., 1995; Meakin and

152 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Morgan, 1999; Lyons et al., 2000; VandenBerg et al., 2000; Cas et al., 2003; Glen, 2004; Figure 3.22).
Middle Devonian to earliest Carboniferous intrusive I-, S- and A-type magmatism, ca. 380 to 350 Ma
(Chappell and White, 1992; Gray et al., 2003; Wormald et al., 2004; Black et al., 2005), occurred
throughout this period.

The granites in the Lachlan Orogen appear to show a variety of diachronous trends (Figure 3.22). In
the Eastern Lachlan of New South Wales and Victoria, ages appear to decrease eastwards, from ca.
380 to ca. 360 Ma (Lewis et al., 1994; VandenBerg et al., 2000; Glen, 2005), most probably reflecting
subduction zone rollback. In Victoria, granites appear to mostly decrease in age towards the
Melbourne Zone, from both east and west sides (VandenBerg et al., 2000; Gray et al., 2003; Rossiter,
2003). The youngest rocks occur within the post-Tabberabberan Middle Devonian to Early
Carboniferous (ca. 385–350 Ma) Central Victorian Magmatic Province (VandenBerg et al., 2000;
Rossiter, 2003). A similar diachronous trend is evident in Tasmania, where granite ages record a
pronounced westward younging in granite age from ca. 400–375 Ma, pre-, syn-, and post-tectonic
granites in the northeast to ca. 370–350 Ma post-tectonic granites in western Tasmania (Black et al.,
2005; Figure 3.17, Figure 3.22). There is little evidence for arc-related magmatism during this period in
the Lachlan, with most evidence clearly indicating that the arc was located further east in the New
England Orogen (e.g., Meakin and Morgan, 1999; Glen, 2005; Collins and Richard, 2008).

The Late Devonian to early Carboniferous in north Queensland consists of largely non-volcanic
(cratonic provenance), terrestrial and lesser marine sedimentation across all regions, best preserved
in the lower successions of the Bundock, Clarke River and Burdekin basins of the Broken River and
Charters Towers regions. Minor andesitic volcanism is recorded in the Georgetown region (Withnall et
al., 1997a), and minor volcaniclastic input is recorded in several of the regions.

During the Late Devonian to early Carboniferous, both terrestrial and marine sedimentation, often with
accompanying volcanism, occurred in the Thomson Orogen and in the Koonenberry region, largely in
response to intracratonic extension following the Tabberabberan Orogeny, but also in response to
backarc extension behind the arc in the New England Orogen (e.g., Neef and Bottrill, 1991; Murray,
1994; Withnall et al., 1995; Henderson et al., 1998; Draper et al., 2004; McKillop et al., 2007; Gilmore
et al., 2007). Backarc extension resulted in rifting and initiation of the Drummond Basin in the latest
Devonian (Henderson et al., 1998; Henderson and Blake, in Fergusson and Henderson et al., 2013).
The Drummond Basin contains a thick succession of continental and lesser marine sediments and
volcanic rocks (Olgers, 1972; Hutton et al., 1998a; Henderson et al., 1998), with the lowermost units
being latest Devonian to early Carboniferous syn-rift related volcanics rocks and associated marine to
terrestrial volcaniclastic sediments (Olgers, 1972; Henderson et al., 1998). In the Thomson Orogen,
felsic and lesser intermediate and mafic magmatism accompanied extension episodically throughout
this time period. This includes intrusive and related extrusive magmatism in the Anakie Province, prior
to Drummond Basin formation, regionally extensive, Late Devonian to early Carboniferous, felsic
magmatism in the Thomson Orogen, including within the Drummond Basin (e.g., Olgers, 1972;
Murray, 1994), and Late Devonian to early Carboniferous granites in the Thomson Orogen basement
and Warburton Basin strata (e.g., Murray, 1994; Kositcin et al., 2015a). The Late Devonian to early
Carboniferous backarc silicic magmatism in the Drummond Basin at this time may be related to
episodes of silicic magmatism in the New England Orogen (e.g., Bryan et al., 2004).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 153


Figure 3.22 General distribution of Late Devonian to early Carboniferous rocks (ca. 380 Ma to ca. 350 Ma) in
eastern Australia. Refer to text for discussion. The New England Orogen continued in this configuration (arc,
forearc and accretionary wedge) until the Late Carboniferous.

154 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.23 Generalised (approximate) distribution of deformation (cross-hatching) within the Kanimblan Orogeny
(ca. 360 to 350 Ma) as defined by outcrop and drill hole information – actual extent of deformation may be greater
and more continuous. The intensity of deformation is variable within the areas indicated. See text for data
sources. Refer to Figure 3.22 for additional geological explanations.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 155


Figure 3.24 Interpreted tectonic model for the Kanimblan cycle (ca. 380 to 350 Ma) of eastern Australia. Refer to
text for detailed discussion.

156 Geodynamic Synthesis of the Phanerozoic of eastern Australia


The extension was ended by the early Carboniferous (ca. 360–340 Ma) contractional east-west
Kanimblan Orogeny (e.g., Gray, 1997; Meakin and Morgan, 1999; VandenBerg et al., 2000; Gray et
al., 2003; Glen, 2005). This orogeny folded and inverted Kanimblan cycle and older rocks. It occurred
across the Lachlan Orogen, into the Delamerian Orogen (e.g., Gilmore et al., 2007), but is best
expressed in the Eastern Lachlan (Gray, 1997; Gray et al., 2003; Glen, 2005). As outlined by Willman
et al. (2002), the post-Tabberabberan sedimentation and volcanic rocks in the Lachlan Orogen overlie
major faults and interpreted suture zones belonging to the Tabberabberan Orogeny, with little
evidence for significant later reactivation. Kanimblan deformation in north Queensland was minor in
nearly all areas, with the exception of the Hodgkinson Province, where significant east-west
shortening and further basin inversion occurred (Garrad and Bultitude, 1999). This deformation
immediately pre-dates the commencement of the voluminous and widespread extrusive and intrusive
magmatism of the Kennedy Province.

The early to ?middle Carboniferous Kanimblan Orogeny, or phase 3 of the Alice Springs Orogeny as it
is known in the Thomson Orogen, (also referred to as the Quilpie Orogeny in the Devonian basins of
Queensland; Finlayson et al., 1990; Finlayson, 1993), produced a major episode of faulting and
deformation in the Koonenberry Belt, slight contraction in the Drummond Basin and regional-scale
folding and subsequent erosion in the Adavale Basin (e.g., Olgers, 1972; Neef, 2004; Gilmore et al.,
2007). Kanimblan deformation in the Drummond Basin is recorded by unconformities between the
Drummond and overlying sequences, e.g., Galilee Basin, Bowen Basin, Bulgonunna Volcanics
(Henderson et al., 1998; Henderson and Blake, in Fergusson and Henderson et al., 2013). This
deformation event is suspected to have driven regional-scale, southward thrusting of the Thomson
over the Lachlan Orogen (Glen et al., 2013). Veevers (2013) has suggested the Kanimblan Orogeny
occurred slightly later––ca. 330 Ma. We, however, place the Kanimblan prior to the voluminous silicic
magmatism of the Kennedy Igneous Association in northern Queensland which commenced ca.345
Ma and continued until ca. 260–250 Ma (e.g., Champion and Bultitude, 2013a).

3.2.6 Early Carboniferous to late Permian (ca. 350 Ma to ca. 265 Ma)
This time period is largely defined based on the tectonic evolution of the New England Orogen. It
includes the continuation of Connors-Auburn Arc in the New England Orogen until its apparent end in
the late Carboniferous (ca. 305 Ma; Cawood, 1982; Murray et al., 1987), and, the ensuing extensional,
backarc phase from the late Carboniferous to late Permian (ca. 305 Ma to ca. 270–265 Ma; Holcombe
et al., 1997a, b). This interval is characterised by extension, accompanied by widespread magmatism
and sedimentation, including initiation of the Sydney-Gunnedah-Bowen Basin System (Figure 3.25).
Extension was terminated by deformation and formation of the Texas and Coffs Harbour Oroclines
(Figure 3.26).

The Andean-style continental margin in the New England Orogen appears to have been minimally
affected by the Kanimblan Orogeny and continued into the post-Kanimblan with little change to the
magmatic arc (Connors-Auburn Arc), forearc basin or accretionary wedge (Cawood, 1982; Murray et
al., 1987; Figure 3.22, Figure 3.25). The orogeny did, however, result in changes to the backarc,
terminating sedimentation in the Drummond Basin (Olgers, 1972), for example. The main recorded
pulses of granite formation in the Connors-Auburn Arc occurred during this time period, between ca.
324–313 Ma in the Auburn Arch and ca. 316–305 Ma in the Connors Arch (Murray, 2003). Similar
aged felsic intrusive magmatism occurs in the northeastern Lachlan Orogen (Figure 3.25), presumably
within a backarc type position. Although there is no record of the actual magmatic arc in the southern
New England Orogen (either buried or removed), its presence is inferred from voluminous arc-derived
volcaniclastics in the forearc (Cawood, 1983; Korsch, 1984).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 157


Figure 3.25 General distribution of early Carboniferous to late Permian rocks (ca. 350–265 Ma) in eastern
Australia. Refer to text for discussion. Note: The configuration shown for the New England Orogen largely reflects
the latter (late-Carboniferous and younger) part of the time period. Prior to this, the New England Orogen
consisted of a continuation of the magmatic arc, forearc and accretionary wedge as shown in Figure 3.24.

158 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.26 Generalised (approximate) distribution of deformation within the Permian orocline forming event in the
New England Orogen. Actual extent of deformation may be greater and more continuous than shown. The
intensity of deformation is variable within the areas indicated. Both two (Texas and Coffs Harbour oroclines in the
north––blue solid line) and four (Manning and Nambucca oroclines in the south––blue dotted line) oroclines have
been proposed for the southern New England Orogen. See text for data sources.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 159


Figure 3.27 Interpreted tectonic model for the early Carboniferous to late Permian period (ca. 350–265 Ma) of
eastern Australia. Refer to text for detailed discussion.

160 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Volcanism and deposition, predominantly of continental to shallow marine clastic sediments, continued
within the forearc basin (e.g., Tamworth Belt, Yarrol Belt; Cawood, 1983; Korsch, 1984; Cawood and
Leitch; 1985; Korsch et al., 1990; Yarrol Project Team, 1997; Murray et al., 2003; Figure 3.22,
Figure 3.25) until the late Carboniferous (Roberts et al., 2004). Similarly, ocean-ward accretion of
trench-fill volcaniclastic turbidites (derived from a magmatic arc) and minor amounts of oceanic crust
(basalt, chert, mudstone), continued within the accretionary wedge east of the arc (e.g., Korsch et al.,
1990; Figure 3.22, Figure 3.25).

The New England Orogen underwent a transition from active accretion in the early Carboniferous to
widespread extension through the late Carboniferous to the late Permian (Leitch, 1988; Holcombe et
al., 1997a), interpreted to reflect the eastward retreat of the subducting slab, and migration of the
volcanic arc offshore (Holcombe et al., 1997a; Roberts et al., 2004). Mechanisms invoked for arc
migration include slab breakoff (Caprarelli and Leitch, 2002) and rollback (Jenkins et al., 2002). By the
Early Permian, much of the New England Orogen was in an extensional continental backarc setting
(Holcombe et al., 1997a, b; Figure 3.25). Deposition of bimodal volcanics, volcaniclastics and
siliciclastic sedimentary rocks occurred in numerous extensional basins (Leitch, 1988; Roberts et al.,
1996; Korsch et al., 1998, 2009a, b; Figure 3.27), overlying older forearc and accretionary wedge
successions. Extension is also recorded by the emplacement of granites into the former accretionary
wedge (Glen, 2005).

Cessation of subduction (~305 Ma; synchronous with cessation of deposition in the Tamworth Belt;
Roberts et al., 1995, 1996) and an eastward shift in magmatism at ca. 300 Ma, probably via rollback,
resulted in generation of the S-type Hillgrove Suite in an extensional backarc environment in the
southern New England Orogen (~302 Ma; Collins et al., 1993; Kent, 1994; Jenkins et al., 2002). S-
type plutons in the northern New England Orogen (North D’Aguilar Block) are also possibly related to
the earliest phase of extension (e.g., 309 ± 4 Ma Gallangowan Granite; 310 ± 4 Ma Yabba Creek
Granite; Holcombe et al., 1997a).

Early Permian subsidence in the western regions of the backarc gave rise to the Sydney-Gunnedah-
Bowen Basin System. The event that initiated the formation of these basins was extensional (at ~305
Ma; Denison Event; Korsch et al., 2009a, b), and the continental crust was stretched to form the
significant Early Permian East Australian Rift System, as defined by Korsch et al. (1998; 2009a;
Figure 3.27). Deposition within the Sydney-Gunnedah-Bowen Basin System was largely controlled by
events within the orogen to the east, particularly the switch (later in the Permian) of the basin from
back-arc to a foreland basin system (Roberts et al., 1996; Korsch et al., 1998; see below).

An inferred Early Permian backarc setting for the New England Orogen and Sydney-Gunnedah-
Bowen Basin requires that an arc environment existed outboard at this time (Figure 3.25, Figure 3.27).
The Gympie Terrane contains early Permian submarine and subaerial volcanic rocks which have been
suggested to have a chemical signature indicative of a juvenile island-arc terrain, isolated from the
influence of continental crust (Sivell and Waterhouse, 1988; Sivell and McCulloch, 1997, 2001;
Figure 3.27). These rocks are intimately associated with primitive oceanic backarc basalts (Sivell and
McCulloch, 1997) suggesting that a well-developed backarc separated the primitive Gympie arc from
the New England Orogen, that is, the intraoceanic Gympie island arc was located east of the
continental margin of Gondwana in the early Permian (Figure 3.27). Detrital zircon data from Gympie
(Korsch et al., 2009c) suggest that the Gympie arc was attached back to eastern Australia at the end
of the Permian or the start of the Triassic.

Also in the early Permian, but probably after extension, oroclinal bending of the forearc and
accretionary wedge successions (~285–265 Ma) produced the Texas and Coffs Harbour oroclines
(Korsch and Harrington, 1987; Murray et al., 1987; Figure 3.26) and possibly also the Manning and

Geodynamic Synthesis of the Phanerozoic of eastern Australia 161


Nambucca oroclines (e.g., Li and Rosenbaum, 2014). Drivers for the oroclinal folding are varied,
including transform faulting, transtension and transpression (e.g., Korsch and Harrington, 1987;
Murray et al., 1987; Fergusson and Leitch, 1993; Korsch et al., 1990; Offler and Foster, 2008),
although, as summarised by Offler and Foster (2008), most models invoke dextral movement
(although cf. Li and Rosenbaum, 2014). The timing of oroclinal bending has also been the subject of
much debate (e.g., see Murray et al., 1987; Fergusson and Leitch 1993; Korsch and Harrington, 1987;
Offler and Foster, 2008); we place oroclinal development after cessation of early Permian backarc
extension but prior to the main phase of the Hunter-Bowen Orogeny. Recent work by Offler and Foster
(2008) suggests orocline deformation commenced at ca. 276 Ma. Li and Rosenbaum (2014) proposed
multiple stages of bending, based on Mediterranean analogies, commencing with initial curvature
forming in the early half of the Permian (related to unequal rates of arc roll-back), followed by
tightening of the orocline structures by late Permian deformation.

The Kanimblan Orogeny was the terminal event in the Lachlan Orogen (e.g., Pogson and Watkins,
1998), and subsequent geology there relates largely to the New England Orogen to the east. From the
Middle Carboniferous to Permian, Eastern Australia was dominated by tectonic extension and rifting,
and formation of intracratonic basins. Within Victoria and New South Wales, the only significant
magmatic event was the north-northwest belt of Carboniferous (ca. 340–310 Ma) intrusive magmatism
in the northeastern Lachlan (e.g., Bryan et al., 1966; Pogson and Watkins, 1998; Meakin and Morgan,
1999; Bodorkos et al., 2010; Figure 3.25). This magmatism consists of the I-type, mostly felsic,
granites, such as those of the Bathurst Batholith and the Gulgong Suite (Bathurst basement terrane of
Chappell et al., 1988). They are of similar age and geochemistry to volcanic and intrusive rocks in the
New England Orogen (e.g., Chappell et al., 1988; Pogson and Watkins, 1998), and may relate to
continental arc formation, although Meakin and Morgan (1999) suggest emplacement in an
extensional environment, presumably behind the continental arc of the New England Orogen. The
eastern part of the Lachlan Orogen is overlain by the latest Carboniferous to Triassic Sydney-
Gunnedah-Bowen Basin System, which also, at least, initially developed as a backarc rift behind the
New England Orogen (e.g., Korsch et al., 2009a). The basin rocks overlie the Carboniferous granites
of the Bathurst-Gulgong area and further south. Of similar age to the Sydney-Gunnedah-Bowen Basin
System are sediments of the Parmeener Supergroup in the Tasmania Basin in Tasmania, which
developed over the suture (Tamar Fracture) between western and northeastern Tasmania (Seymour
and Calver, 1995, 1998; Figure 3.25). These consist of a lower succession of glacial and marine
sediments and an upper succession (late Permian and younger) of nonmarine sediments, including
coal measures (Seymour and Calver, 1995, 1998). Remnants of possibly more widespread Permian
glacial and marine sedimentation are also recorded (outcrop and sub-surface, e.g., beneath the
Murray Basin) in Victoria and southern New South Wales (O’Brien et al., 2003).

This period in north Queensland and in the northern Thomson Orogen (Charters Towers region and
northern Drummond Basin) is characterised by the commencement of the widespread and voluminous
extrusive and intrusive magmatism of the Kennedy Igneous Association (Champion and Bultitude,
2013a), plus associated, mostly minor sedimentation (Figure 3.25). As documented by numerous
authors (e.g., Richards et al., 1966), this magmatism is crudely diachronous, commencing earlier in
the Georgetown, Broken River and Charters Towers regions (ca. 345–335 Ma), and younging to mid-
Permian in the Hodgkinson Province (See Champion and Bultitude, 2013a). There are also
accompanying changes in geochemistry. Magmatism in the Carboniferous is almost exclusively I-type,
along with some mantle-derived magmatism. In the early Permian, magmatism switched to A- and I-
type in the Georgetown and western Hodgkinson regions, and to S- and I-type in the central and
eastern Hodgkinson (Champion and Chappell, 1992; Champion and Bultitude, 2003, 2013a, b). The
tectonic regime for this magmatism is not well understood. Despite the strong crustal input into the
magmatism, it is generally thought to be broadly arc-related (e.g., Champion and Bultitude, 2003,

162 Geodynamic Synthesis of the Phanerozoic of eastern Australia


2013a, though Murray et al. (1987), suggested a transform environment), probably in an extensional
backarc or behind-arc position relative to the continental arc in the New England Orogen, as shown in
Figure 3.27. This is consistent with the younging of magmatism to the east matching the inferred
migration of the Connors-Auburn Arc eastwards at this time (Holcombe et al., 1997a; Jenkins et al.,
2002; Roberts et al., 2004). This backarc position is accentuated if the oroclinal folding of the New
England Orogen is restored, pushing the northern part of the orogen east of northern Queensland
(e.g., Korsch et al., 2009c).

Deformation in this time period in north Queensland appears to have occurred at least twice. There is
a widespread but minor north-south contraction, thought to be late Carboniferous in age, and
commonly equated with phase 3 of the Alice Springs Orogeny, found in the Broken River, Hodgkinson
and Georgetown regions (Withnall et al., 1997a, b; Bultitude et al., 1997). In addition, there is a more
significant deformation in the early Permian, possibly related to an early phase of the Hunter-Bowen
Orogeny. This penetrative east-west shortening deformation is best documented in the Hodgkinson
Province (e.g., Davis, 1994; Davis et al., 1996; Garrad and Bultitude, 1999; Figure 3.26).

From the late Carboniferous to early Permian, tectonic extension and rifting initiated extensive
intracratonic basin formation in eastern Australia (e.g., Cooper and Galilee basins; Bain and Draper,
1997; Figure 3.27). Continental margin extension in the Early Permian led to formation of the Bowen-
Gunnedah-Sydney Basin System in a backarc setting (Korsch et al., 1998; Figure 3.27), consistent
with the presence of latest Carboniferous to early Permian bimodal and calcalkaline volcanics at the
base of the Bowen Basin succession (e.g., Green et al., 1997b). Sedimentation in the Bowen Basin
was contiguous with the Gunnedah Basin and the former unconformably overlies the Drummond
Basin (e.g., Green et al., 1997a). By the early Permian, the Bowen Basin, together with the Gunnedah
and Sydney Basins, formed the ‘East Australian Rift System’ (Korsch et al., 1998; Figure 3.27).
Although extension was located on the continental margin, basins such as the Galilee and Cooper
formed on the craton further west during the late Carboniferous to Permian at this time (Draper and
McKellar, 2002). These contain widespread terrestrial and glacial sediments, including coal measures
(e.g., Scott et al., 1995; Draper, 2002a, b; Gray and McKellar, 2002). A connection between the
Cooper and Galilee basins meant the two basins experienced related sediment deposition (Scott et
al., 1995). It has been suggested that deformation by the Alice Springs Orogeny in the Carboniferous
caused convective downwelling and regional downwarp of the Drummond Basin, resulting in the
formation of troughs and depressions in the Galilee Basin (Jackson et al., 1981; Middleton and Hunt,
1989). Our relatively simple tectonic interpretation for this period is shown in Figure 3.27. In our model
we have not accounted for oroclinal shortening, as did Murray et al. (1987) and Korsch et al. (2009c),
who both extended New England northwards to be east of north Queensland.

3.2.7 Late Permian to late Triassic (ca. 265 Ma to 230 Ma)


This time period in eastern Australia is largely defined by the New England Orogen and associated
backarc basins. The cycle commenced in the late Permian (~265–262 Ma) when a magmatic arc was
apparently re-established along the Paleo-Pacific continental margin of Australia and the previous
backarc switched from an extensional to a contractional regime (Korsch and Totterdell, 1995;
Figure 3.30). This is thought to have led to the formation of a retroforeland fold-thrust belt west of the
magmatic arc and the development of a major retroforeland basin phase in the Bowen-Gunnedah-
Sydney Basin System that continued until at least the Middle Triassic (Korsch and Totterdell, 1995).
This period is, therefore, characterised by renewed arc magmatism, the foreland basin stage of
development of the Bowen-Gunnedah-Sydney Basin System and the Hunter-Bowen Orogeny
(Figure 3.28, Figure 3.29, Figure 3.30).

Geodynamic Synthesis of the Phanerozoic of eastern Australia 163


The onset of the Hunter-Bowen Orogeny and the resultant change from extensional to contractional
tectonism in the late Permian, is marked by a well-developed unconformity in the backarc Bowen Basin
and intracratonic Galilee and Cooper basins (e.g., Stephens et al., 1996; Korsch et al., 1998). The switch
to contraction resulted in the transition of the Bowen-Gunnedah-Sydney basins from backarc extensional
basins to foreland basins in a backarc setting (Roberts et al., 1996; Korsch et al., 2009b).

Subsidence and sedimentation is suggested to have been driven by thrust loading related to
westward-propagating thrust sheets from the New England Orogen (Korsch et al., 2009b). The
Bowen-Gunnedah-Sydney Basin System retained this foreland basin setting until the Middle or Late
Triassic (e.g., Harrington and Korsch, 1985; Korsch and Totterdell, 1995; Korsch et al., 2009b).
Foreland basin sedimentation is also recorded in the Gogango Thrust Zone, Yarrol Terrane, and in the
Gympie Terrane (Figure 3.28). Related forearc and accretionary wedge sedimentation in the New
England Orogen is though to be located further east (now offshore, e.g., Korsch, pers. comm., 2008.)

Terrestrial and marine sedimentation continued throughout the Permian and up to the late Triassic in
both foreland (e.g., Bowen-Gunnedah-Sydney basins), and intracratonic basins (e.g., Cooper and
Galilee basins). Fluvial and lacustrine systems were associated with extensive peat swamps (coal
measures) in the Cooper and Galilee basins (Gray and McKellar, 2002; Scott et al., 1995; Cowley,
2007). In the late Permian, coastal swamps also formed in the subsiding Bowen Basin, leading to an
accumulation of extensive coal deposits (Shaw, 2002). Contemporaneous sedimentation is also
recorded in the Parmeener Supergroup of the Tasmania Basin in Tasmania (Seymour and Calver,
1995, 1998). These contain an upper succession (late Permian and younger) of non-marine
sediments, which also include coal measures (Seymour and Calver, 1995, 1998).

Permian glacial and marine sedimentation recorded (outcrop and sub-surface, e.g., beneath the
Murray Basin) in Victoria and southern New South Wales do not appear to have continued into this
time period (O’Brien et al., 2003). Local, possibly originally more extensive, outcropping and
concealed (by younger cover), late Permian largely terrestrial sediments and coal measures occur in
the Coen region and in the Hodgkinson Province in northern Queensland (McConachie et al., 1997;
Garrad and Bultitude, 1999).

The later part of the Permian saw a re-establishment of a magmatic arc along the paleo-Pacific
continental margin of Australia, as recorded by voluminous Late Permian to Early (and Middle)
Triassic calcalkaline, intrusive and extrusive, magmatism throughout the New England Orogen (ca.
270–230 Ma; Gust et al., 1993; Bryant et al., 1997; Holcombe et al., 1997b; Van Noord, 1999; Murray,
2003; Donchak et al., 2013; Figure 3.28, Figure 3.30). Magmatism was dominantly andesitic to felsic
in composition, consistent with a magmatic arc interpretation (Gust et al., 1993; Bryant et al., 1997;
Holcombe et al., 1997b; Murray, 2003; Donchak et al., 2013), although Murray (2003) notes that the
granites are compositionally very similar to the earlier (late Carboniferous to early Permian), backarc,
magmatism in the New England Orogen. Magmatism appears to wane by the late Triassic (Gust et al.,
1993), although age control is poor.

Widespread late Permian intrusive and extrusive magmatism (and minor sediments) is also recorded
in north Queensland, where it represents the youngest component of the voluminous Carboniferous-
Permian Kennedy Province magmatism (Bain and Draper, 1997; Champion and Bultitude, 2013a;
Figure 3.28). Late Permian Kennedy Province magmatism is largely confined to the eastern
Hodgkinson Province (dominantly S-type; Bultitude and Champion, 1992; Bultitude et al., 1997;
Garrad and Bultitude, 1999; Champion and Bultitude, 2013a,b) and the Charters Towers region (I-type
and mantle-derived; Hutton et al., 1997), but also includes A-, I- and mantle-derived extrusive and
intrusive magmatism in the Coen, Georgetown and western Hodgkinson regions (Withnall et al.,
1997a, Blewett et al., 1997).

164 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.28 General distribution of late Permian to late Triassic (ca. 265–230 Ma) rocks in eastern Australia. Refer
to text for discussion.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 165


Figure 3.29 Generalised (approximate) distribution of deformation within the Hunter-Bowen Orogeny as defined
by outcrop–the actual extent of deformation may be greater and more continuous than shown. The intensity of
deformation is variable within the areas indicated. Deformation in the New England Orogen during this period was
episodic, in apparent contrast in the region to the west. See text for data sources.

166 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Figure 3.30 Interpreted tectonic model for the late Permian to late Triassic period (ca. 265–230 Ma) in eastern
Australia. Refer to text for detailed discussion.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 167


This magmatism is generally thought to be broadly arc-related (e.g., Champion and Bultitude, 2003,
2013a), probably in an extensional backarc or behind arc position relative to the continental arc in the
New England Orogen (Figure 3.30). Unlike the New England Orogen, magmatism in north
Queensland ceased by the end of the Permian and no Early Triassic magmatism is recorded. West of
the New England Orogen, restricted Triassic magmatic activity is recorded in the Cooper Basin
(Murray, 1994; Draper, 2002a, b) and in the Bourke-Louth region (Burton et al., 2007; Figure 3.28).
These magmatic events suggest that tectonism on the eastern continental margin may have
influenced the craton further west, for example, Burton et al. (2007) suggest that the Midway Granite,
and coeval intrusives east of Bourke, indicate a more spatially widespread Middle Triassic magmatic
pulse than is currently recognised.

The Hunter-Bowen Orogeny covers a period of about 35 Myr from ca. 265 Ma to ca. 230 Ma, (Murray,
1997b; Holcombe et al., 1997b; Roberts et al., 2006; Korsch et al., 2009b; Figure 3.29) and
encompasses all of the Hunter-Bowen cycle. The initiation of this event in the New England Orogen is
marked by a major unconformity in the middle part of the Permian (Korsch et al., 1998). In the New
England Orogen, deformation is characterised by retrothrusting driven by subduction further to the
east. This major west-directed thrusting led to the formation of a retroforeland fold-thrust belt (Korsch
et al., 1990; 1997; Fergusson, 1991; Holcombe et al., 1997b; Korsch, 2004). Subsidence of the Bowen
and Gunnedah basins during the foreland basin phase was driven by thrust loading related to these
westward-propagating thrust sheets (Korsch et al., 2009b). The foreland basin phase of sedimentation
associated with the Hunter-Bowen Orogeny was punctuated by a series of discrete contractional
events (Korsch et al., 2009b) in the Permian and Triassic.

Recent detrital zircon age data suggest that the island arc component of the Gympie Terrane came
into contact with the mainland New England Orogen, that is, continent-island arc collision, prior to
~250 Ma (Permian-Triassic boundary; Korsch et al., 2009c). Approximately similar timing is recorded
in north Queensland, where late Permian east-west deformation, equated with the Hunter-Bowen
Orogeny, is best developed in the Hodgkinson and Barnard provinces (Garrad and Bultitude, 1999;
Figure 3.29). Contractional events of the Hunter-Bowen Orogeny also appear to have thrust the
Tamworth Belt westwards over the eastern edge of the Sydney-Gunnedah Basin and Lachlan Craton
(Korsch et al., 1997; Roberts et al., 2004) around this time in the Late Permian and earliest Triassic.

In the late Middle to Late Triassic, regional contraction resulted in uplift and erosion, and cessation of
deposition in the Galilee, Cooper and Bowen basins (Apak et al., 1997; Bain and Draper, 1997; Green
et al., 1997b; Korsch et al., 1998), as well as east-west deformation and uplift in the Drummond Basin
(Olgers, 1972; Fenton and Jackson, 1989; Murray, 1990; Johnson and Henderson, 1991).

Our inferred tectonic interpretation for the late Permian to late Triassic period and Hunter-Bowen
Orogeny is shown in Figure 3.30. The Hunter-Bowen Orogeny marks the timing of effective
cratonisation of eastern Australia. Within Australia, the period after ca. 230 Ma was characterised by a
switch in geodynamics back to an extensional, probably backarc environment (e.g., Holcombe et al.,
1997b). This resulted in a change in plutonism (A-type granites), felsic and/or bimodal volcanism
(Stephens et al., 1993; Holcombe et al., 1997b) and development of extensional basins with coal-
bearing successions (Holcombe et al., 1997b; Shaw, 2002).

The Tasman Orogen was also effectively terminated at this time (ca. 230 Ma) with the stepping out of
the plate margin to the east following the Hunter-Bowen Orogeny and ensuing extension leading to
subsequent breakup of Gondwana–Pangaea (e.g., Veevers, 2004, 2013; Cawood, 2005; Glen, 2005).

168 Geodynamic Synthesis of the Phanerozoic of eastern Australia


4 References

Adams, C.J., Wormald, R. and Henderson, R.A., 2013. Detrital zircons from the Hodgkinson
Formation: constraints on its maximum depositional age and provenance. In P.A. Jell (ed.) Geology
of Queensland. Geological Survey of Queensland, pp. 239-241.
Aitchison, J. C. 1988. Early Carboniferous (Tournaisian) Radiolaria from the Neranleigh-Fernvale beds
near Brisbane. Queensland Government Mining Journal 89, 240-241.
Aitchinson, J.C. and Buckman, S., 2012. Accordion vs. quantum tectonics: Insights into continental
growth processes from the Paleozoic of eastern Gondwana. Gondwana Research 22, 674–680.
Aitchison, J. C. and Flood, P.G. 1992. Early Permian transform margin development of the southern
New England orogen, eastern Australia (eastern Gondwana). Tectonics 11, 1385-1391.
Aitchison, J. C. and Flood, P. G. 1995. Gamilaroi Terrane: a Devonian rifted intra-oceanic island-arc
assemblage, NSW, Australia. In: Smellie, J. L. (ed.) Volcanism associated with extension at
consuming plate margins. Geological Society of Australia Special Publication 81, 155-168.
Aitchison, J. C. and Ireland, T. R. 1995. Age profile of ophiolitic rocks across the late Palaeozoic New
England Orogen, New South Wales: implications for tectonic models. Australian Journal of Earth
Sciences 42, 11-23.
Aitchison, J. C., Blake, M. C., Jr., Flood, P. G. and Jayko, A. S. 1994. Palaeozoic ophiolitic
assemblages within the southern New England Orogen of eastern Australia: implications for growth
of the Gondwana margin. Tectonics 13, 1135-1149.
Aitchison, J. C., Ireland, T.R., Blake Jr, M.C. and Flood, P.G. 1992a. 530 Ma zircon age for ophiolite
from the New England orogen: Oldest rocks known from eastern Australia. Geology 20, 125-128.
Aitchison, J. C., Flood, P. G. and Spiller, F. C. P. 1992b. Tectonic setting and paleoenvironment of
terranes in the southern New England Orogen, eastern Australia as constrained by radiolarian
biostratigraphy. Palaeogeogr. Palaeoclimatol. Palaeoecol. 94, 31-54.
Ali, A., 2010. The tectono-metamorphic evolution of the Balcooma Metamorphic Group, north-eastern
Australia: a multidisciplinary approach. Journal of Metamorphic Geology 28, 397-422.
Apak, S.N., Stuart, W.J., Lemon, N.M. and Wood, G. 1997. Structural evolution of the Permian -
Triassic Cooper Basin, Australia: relation to hydrocarbon trap styles. American Association of
Petroleum Geologists, Bulletin 81, 533-555.
Arnold, G.O. and Fawckner, J.F. 1980. The Broken River and Hodgkinson Provinces. In: Henderson,
R.A. and Stephenson, P.J. (eds). The Geology and Geophysics of Northeastern Australia.
Geological Society of Australia, Queensland Division, Townsville. pp. 175-189.
Bain, J.H.C. and Draper, J.J (eds), 1997. North Queensland Geology. Bulletin of the Australian
Geological Survey Organisation 240, and Queensland Department of Mines and Energy Geology
9.
Barnicoat, A.C. 2007. Mineral Systems and Exploration Science: Linking fundamental controls on ore
deposition with the exploration process. In: Andrew, C.J., et al. (eds) Digging Deeper: Proceedings
th rd
of the Ninth Biannual SGA Meeting, Dublin, Ireland 20 -23 August 2007, p. 1407-1410.
Barnicoat, A.C. 2008. The Mineral Systems approach of the pmd*CRC. In: Korsch, R.J. and Barnicoat,
A.C. (eds) New Perspectives: The foundations and future of Australian exploration. Abstracts for
the June 2008 pmd*CRC Conference. Geoscience Australia Record 2008/09, p. 1-6.
Barron, L. M., Ohnenstetter, M., Barron, B. J., Suppel, D.W. and Falloon, T., 2004. Geology of the
Fifield Alaskan-type complexes, New South Wales, Australia. Geological Survey of New South
Wales Report GS2004/323.
Bell, T.H. 1980. The deformation history of northeastern Queensland - a new framework. In:
Henderson, R.A. and Stephenson, P.J. (eds). The Geology and Geophysics of Northeastern
Australia. Geological Society of Australia, Queensland Division, Townsville. pp. 307-313.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 169


Belperio, A.P., Preiss, W.V., Fairclough, M.C., Gatehouse, C.G., Gum, J., Hough, J. and Burtt, A.,
1998. Tectonic and metallogenic framework of the Cambrian Stansbury Basin–Kanmantoo Trough,
South Australia. AGSO Journal of geology and Geophysics 17, 183–200.
Berry, R. F. and Crawford, A. R. 1988. The tectonic significance of Cambrian allochthonous mafic-
ultramafic complexes in Tasmania. Australian Journal of Earth Sciences 35, 523-533.
Berry, R.F., Huston, D.L., Stolz, A.J., Hill, A.P., Beams, S.D., Kuronen, U. and Taube, A. 1992.
Stratigraphy, structure and volcanic-hosted mineralisation of the Mount Windsor Subprovince,
North Queensland, Australia. Economic Geology 87, 739-763.
Berry, R.F., Chmielowski, R.M., Steele, D.A. and Meffre, S. 2007. Chemical U-Th-Pb monazite dating
of the Cambrian Tyennan Orogeny, Tasmania. Australian Journal of Earth Sciences 54, 757-771.
Bierlein, F.P., Foster, D.A., Gray, D.R. and Davidson, G.J. 2005. Timing of orogenic gold
mineralisation in northeastern Tasmania; implications for the tectonic and metallogenetic evolution
of Palaeozoic SE Australia. Australian Journal of Earth Sciences 39, 890-903.
Birch, W.D. (ed.) 2003. Geology of Victoria. Geological Society of Australia Special Publication 23, pp.
842.
Black, L.P., 1994. Appendix 2: U–Pb zircon geochronology, Australian Geological Survey Organisation
Record 1994/21, pp. 163–8.
Black, L.P. 2005, SHRIMP U-Pb zircon ages obtained during 2004/05 for the NSW Geological Survey:
Unpublished report for NSW Department of Primary Industries.
Black, L.P. 2006, SHRIMP U-Pb zircon ages obtained during 2005/06 for NSW Geological Survey
projects. Unpublished report for NSW Department of Primary Industries.
Black, L.P. 2007, SHRIMP U-Pb zircon ages obtained during 2006/07 for NSW Geological Survey
projects. Unpublished report for NSW Department of Primary Industries.
Black, L.P., Bultitude, R.J., Sun, S.S., Knutson, J. and Blewett, R.S., 1992. Emplacement ages of
granitic rocks in the Coen Inlier (Cape York): implications for local geological evolution and regional
correlation. BMR Journal of Australian Geology and Geophysics 13, 191–200.
Black, L. P., Seymour, D. B., Corbett, K. D., Cox, S. E., Streit, J. E., Bottrill, R. S., Calver, C. R.,
Everard, J. L., Green, G. R., McClenaghan, M. P., Pemberton, J., Taheri, J. and Turner, N. J. 1997.
Dating Tasmania’s oldest geological events. Australian Geological Survey Organisation Record
1997/15.
Black, L.P., Calver, C.R., Seymour, D.B. and Reed, A., 2004. SHRIMP U-Pb detrital zircon ages from
Proterozoic and early Palaeozoic sandstones and their bearing on the early geological evolution of
Tasmania. Australian Journal of Earth Sciences 51, 885-900.
Black, L.P., McClenaghan, M.P., Korsch, R.J., Everard, J.L and Foudoulis, C. 2005. Significance of
Devonian-Carboniferous igneous activity in Tasmania as derived from U-Pb SHRIMP dating of
zircon. Australian Journal of Earth Sciences 52, 807-829.
Blewett, R.S., Denaro, T.J., Knutson, J., Wellman, P., Mackenzie, D.E., Cruikshank, B.I., Wilford, J.R.,
Von Gnielinski, F.E., Pain, C.F., Sun, S.S and Bultitude, R.J. 1997. Coen Region. In: Bain, J.H.C.
and Draper, J.J (eds). North Queensland Geology. Bulletin of the Australian Geological Survey
Organisation 240, and Queensland Department of Mines and Energy Geology 9, p. 117-138.
Bodorkos, S., Simpson, C.J., Thomas, O.D., Trigg, S.J., Blevin, P.L., Campbell, L.M., Deyssing, L.J.
and Fitzherbert, J.A., 2010. New SHRIMP U-Pb zircon ages from the eastern Lachlan Orogen,
New South Wales, July 2008–June 2009. Geoscience Australia Record 2010/31.
Bodorkos, S., Blevin, P.L., Simpson, C.J., Gilmore, P.J., Glen, R.A., Greenfield, J.E., Hegarty, R. and
Quinn, C.D. 2013. New SHRIMP U-Pb zircon ages from the Lachlan, Thomson and Delamerian
orogens, New South Wales: July 2009–June 2010. Geoscience Australia Record 2013/29,
Geological Survey of New South Wales Report GS2013/0427. Geoscience Australia: Canberra.
Bodorkos, S., Blevin, P.L., Eastlake, M.A., Downes, P.M., Campbell, L.M., Gilmore, P.J., Hughes,
K.S., Parker, P.J. and Trigg, S.J. 2015. New SHRIMP U-Pb zircon ages from the central and
eastern Lachlan Orogen, New South Wales: July 2013–June 2014. Record 2015/02, Geoscience
Australia, Canberra; Report GS2015/0002. Geological Survey of New South Wales, Maitland.
http://dx.doi.org/10.11636/Record.2015.002.
Boucher, R. K. 2001. Warburton Basin GIS data atlas (CD ROM). PIRSA Volume, CD ROM.

170 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Brakel, A.T., Totterdell, J.M., Wells, A.T. and Nicoll, M.G., 2009, Sequence stratigraphy and fill history
of the Bowen Basin, Queensland. Australian Journal of Earth Sciences 56, 401–32.
Brown, D.D., Carr, P.A. and Purdy, D.J., 2012. Database of basement drill holes in the Thomson
Orogen and Roma Shelf regions, Queensland. Geological Survey of Queensland Record 2012/06.
Brown, R.E. 1987. Newly defined stratigraphic units from the western New England Fold Belt, Manilla
1:250 000 sheet area. Quarterly notes of the Geological Survey of New South Wales 69, 1-9.
Bruce, M. C. and Niu, Y. 2000. Evidence for Palaeozoic magmatism recorded in the late
Neoproterozoic Marlborough ophiolite, New England Fold Belt, central Queensland. Australian
Journal of Earth Sciences 47, 1065-1076.
Bruce, M.C., Niu, Y., Harbort, T.A. and Holcombe, R.J., 2000. Petrological, geochemical and
geochronological evidence for a Neoproterozoic ocean basin recorded in the Marlborough terrane
of the northern New England Fold Belt. Australian Journal of Earth Sciences 47, 1053-1064.
Bryan, J.H., McElroy, C.T. and Rose, G., 1966. Sydney 1:250 000 Geological Sheet SI/56-5, third
edition. Geological Survey of New South Wales, Sydney.
Bryan, S. B., Holcombe, R. J. and Fielding, C. R. 2003. Reply. Yarrol terrane of the northern New
England Fold Belt: forearc or backarc? Australian Journal of Earth Sciences 50, 278-293.
Bryan, S. B., Holcombe, R. J. and Fielding, C. R., 2001. Yarrol terrane of the northern New England
Fold Belt: forearc or backarc? Australian Journal of Earth Sciences 48, 293-316.
Bryan, S. E., Allen, C. M., Holcombe, R. J. and Fielding, C. R. 2004. U-Pb zircon geochronology of
Late Devonian to Early Carboniferous extension-related silicic volcanism in the northern New
England Fold Belt. Australian Journal of Earth Sciences 51, 645-664.
Bryant, C.J., Arculus, R.J. and Chappell, B.W. 1997. Clarence River Supersuite; 250 Ma Cordilleran
tonalitic I-type intrusions in eastern Australia. Journal of Petrology 38, 975-1001.
Bultitude, R.J. and Champion, D.C. 1992. Granites of the eastern Hodgkinson Province - their field
and petrographic characteristics. Queensland Resource Industries Record 1992/6.
Bultitude, R.J., Donchak, P.J.T., Domagala, J. and Fordham, B.G., 1993. The pre-Mesozoic
stratigraphy and structure of the western Hodgkinson Province and environs. Department of
Minerals and Energy, Queensland Geological Record 1993/29.
Bultitude, R.J., Garrard, P.D., Donchak, P.J.T., Domagala, J., Champion, D.C., Rees, I.D., Mackenzie,
D.E., Wellman, P., Knutson, J., Fanning, C.M., Fordham, B.G., Grimes, K.G., Oversby, B.S.,
Rienks, I.P., Stephenson, P.J., Chappell, B.W., Pain, C.F., Wilford, J.F., Rigby, J.F and Woodbury,
M.J. 1997. Cairns Region. In: Bain, J.H.C. and Draper, J.J (eds). North Queensland Geology.
Bulletin of the Australian Geological Survey Organisation 240, and Queensland Department of
Mines and Energy Geology 9, p. 225-198.
Burton, G.R., 2010. New structural model to explain geophysical features in northwestern New South
Wales: implications for the tectonic framework of the Tasmanides. Australian Journal of Earth
Sciences 57, 23–49.
Burton, G.R., Trigg, S.J. and Black, L.P. 2007. A Middle Triassic age for felsic intrusions and
associated mineralisation in the Doradilla prospect area, New South Wales. NSW Quarterly Notes,
125: 1-11.
Burton, G.R., Dadd, K.A. and Vickery, N.M., 2008. Volcanic arc-type rocks beneath cover 35km to the
northeast of Bourke. Quarterly Notes of the Geological Survey of New South Wales, 127, 1-23.
Burton, G.R., Trigg, S.J. & Campbell L. M., 2012. Byrock 1:100 000 geological sheet 8136.
Explanatory Notes. Geological Survey of New South Wales, Maitland.
Burton, G.R., Trigg, S.J. & Campbell L. M., 2013. Sussex and Byrock 1:100 000 geological sheets
8135 and 8136. Explanatory Notes. Geological Survey of New South Wales, Maitland.
Calver, C.R. and Walter, M.R. 2000. The Late Neoproterozoic Grassy Group of King Island, Tasmania:
correlation and palaeogeographic significance. Precambrian Research, 100, 299-312.
Calver, C. R., Black, L. P., Everard, J. L., Seymour, D. B. 2004. U-Pb zircon age constraints on late
Neoproterozoic glaciation in Tasmania. Geology 32, 893-896.
Caprarelli, G. and Leitch, E.C., 1998. Magmatic changes during the stabilisation of a Cordilleran fold
belt: the Late Carboniferous-Triassic igneous history of eastern New South Wales, Australia. Lithos
45, 413-430.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 171


Caprarelli, G. and Leitch, E.C., 2001. Geochemical evidence from Lower Permian volcanic rocks of
northeast New South Wales for asthenospheric upwelling following slab breakoff. Australian
Journal of Earth Sciences 48, 151–166.
Caprarelli, G. and Leitch, E.C., 2002. MORB-like rocks in a Palaeozoic convergent margin setting,
northeast New South Wales. Australian Journal of Earth Sciences 49, 367-374.
Carson, C.J., Hutton, L.J., Withnall, I.W., Perkins, W.G., Donchak, P.J.T., Parsons, A., Blake, P.R.,
Sweet, I.P., Neumann, N.L. and Lambeck, A., 2011. Summary of results: Joint GSQ-GA NGA
geochronology project, Mount Isa region, 2009–2010. Queensland Geological Record 2011/03.
Cas, R.A.F. (coordinator), O’Halloran, G.J., Long, J.A. and VandenBerg, A.H.M., 2003. Middle
Devonian to Carboniferous. Late to post-tectonic sedimentation and magmatism in an arid
continental setting. In Birch, W.D. (ed.) Geology of Victoria. Geological Society of Australia Special
Publication 23, pp. 157-193.
Cawood, P.A. 1976. Cambro-Ordovician strata in northern New South Wales. Search 7, 317-318.
Cawood, P.A. 1982. Structural relations in the subduction complex of the Palaeozoic New England
Fold Belt, eastern Australia. Journal of Geology 90, 381-392.
Cawood, P.A. 1983. Modal composition and detrital clinopyroxene geochemistry of lithic sandstones
from the New England Fold Belt (east Australia): A Palaeozoic forearc terrane. Geological Society
of America Bulletin 94, 1199-1214.
Cawood, P.A. 2005. Terra Australis Orogen: Rodinia breakup and development of the Pacific and
Iapetus margins of Gondwana during the Neoproterozoic and Palaeozoic. Earth Science Reviews
69, 249-279.
Cawood, P.A. and Flood, R. H. 1989. Geochemical character and tectonic significance of Early
Devonian keratophyres in the New England Fold Belt, eastern Australia. Australian Journal of Earth
Sciences 36, 297-311.
Cawood, P.A. and Leitch, E.C. 1985. Accretion and dispersal tectonics of the southern New England
Fold Belt, eastern Australia. In: Howell, D. G. (ed.) Tectonostratigraphic Terranes of the Circum-
Pacific Region, pp. 481-492. Circum-Pacific Council for Energy and Mineral Resources, Earth
Science Series 1.
Cawood, P.A., Pisarevsky, S.A. and Leitch, E.C., 2011. Unraveling the New England orocline, east
Gondwana accretionary margin. Tectonics 30, TC5002, doi:10.1029/2011TC002864.
Cayley, R.A., 2011. Exotic crustal block accretion to the eastern Gondwanaland margin in the late
CambrianTasmania, the Selwyn Block, and implications for the CambrianSilurian evolution of the
Ross, Delamerian, and Lachlan orogens. Gondwana Research 19, 628-649.
Cayley, R. A., Taylor, D. H., VandenBerg, A. H. M. and Moore, D. H. 2002. Proterozoic - Early
Palaeozoic rocks and the Tyennan Orogeny in central Victoria: the Selwyn Block and its tectonic
implications. Australian Journal of Earth Sciences 49, 225 - 254.
Cayley, R.A., Korsch, R.J., Moore, D.H., Costelloe, R.D., Nakamura, A., Willman, C.E., Rawling, T.J.,
Morand, V.J., Skladzien, P.B. and O'Shea, P.J., 2011. Crustal architecture of Central Victoria:
results from the 2006 deep crustal reflection seismic survey. Australian Journal of Earth Sciences
58. 113-156. DOI: 10.1080/08120099.2011.543151
Champion, D.C., 2013. Neodymium depleted mantle model age map of Australia: explanatory notes
and user guide. Record 2013/044. Geoscience Australia, Canberra.
http://dx.doi.org/10.11636/Record.2013.044.
Champion, D.C. and Bultitude, R.J. 2003. Granites of north Queensland. In: Blevin, P., Jones, M. and
Chappell, B. (eds) Magmas to Mineralisation: The Ishihara Symposium. Geoscience Australia
Record 2003/14, 19-23.
Champion, D.C. and Bultitude, R.J., 2013a. Kennedy Igneous Association. In P.A. Jell (ed.) Geology
of Queensland. Geological Survey of Queensland, pp. 473–514.
Champion, D.C. and Bultitude, R.J., 2013b. The geochemical and Sr-Nd isotopic characteristics of
Paleozoic fractionated S-types granites of north Queensland: Implications for S-type granite
petrogenesis. Lithos 162–163, 37–56.
Champion, D. C. and Chappell, B. W, 1992. Petrogenesis of felsic I-type granites: an example from
northern Queensland. Transactions of the Royal Society of Edinburgh: Earth Sci., 83, 115-126.

172 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Champion, D.C. and Heinemann, M.A., 1994. Igneous rocks of northern Queensland: 1:500 000 map
and explanatory notes. Australian geological Survey Organisation, Record 1994/11.
Champion, D.C., Kositcin, N., Huston, D.L., Mathews, E. and Brown, C., 2009. Geodynamic Synthesis
of the Phanerozoic of eastern Australia and Implications for metallogeny. Geoscience Australia
Record 2009/18.
Chappell, B. W. and White, A. J. R. 1992. I- and S-type granites in the Lachlan Fold Belt. Transactions
of the Royal Society of Edinburgh 83, 1-26.
Chappell, B. W., English, P. M., King, P. L., White, A. J. R. and Wyborn, D. 1991. Granites and related
rocks of the Lachlan Fold Belt. 1:1 250 000 Geological Map. Bureau of Mineral Resources,
Canberra.
Chappell, B. W., White, A. J. R. and Hine, R., 1988. Granite provinces and basement terranes in the
Lachlan Fold Belt, southeastern Australia. Australian Journal of Earth Sciences 35, 505–521
Chisholm, E.I., Blevin, P.L. and Simpson, C.J. 2014a. New SHRIMP U–Pb zircon ages from the New
England Orogen, New South Wales: July 2012–June 2014. Record 2014/52. Geoscience Australia,
Canberra; Report GS2014/0422. Geological Survey of New South Wales, Maitland.
http://dx.doi.org/10.11636/Record.2014.052.
Chisholm, E. I., Blevin, P. L., Downes, P. M. and Simpson, C. J. 2014b. New SHRIMP U–Pb zircon
ages from the central Lachlan Orogen and Thomson Orogen, New South Wales: July 2011–June
2012. Record 2014/32. Geoscience Australia, Canberra. Report GS2013/1837. Geological Survey
of New South Wales, Maitland. http://dx.doi.org/10.11636/Record.2014.032.
Chisholm, E. I., Fitzherbert, J. A., Deyssing, L. and Simpson, C. J. 2014c. New SHRIMP U–Pb zircon
ages from the Captains Flat area, Eastern Lachlan Orogen, New South Wales: July 2012–June
2013. Record 2014/07. Geoscience Australia, Canberra. Report GS2014/0090. Geological Survey
of New South Wales, Maitland. http://dx.doi.org/10.11636/Record.2014.007.
Collins, W.J. and Hobbs, B.E. 2001. What caused the Early Silurian change from mafic to silicic (S-
type) magmatism in the eastern Lachlan Fold Belt? Australian Journal of Earth Sciences 48, 25-41.
Collins, W.J. and Richards, S.W. 2008. Geodynamic significance of S-type granites in Circum-Pacific
orogens. Geology 36, 559-562.
Collins, W.J. and Vernon R.H., 1994. A rift-drift-delamination model of continental evolution:
Palaeozoic tectonic development of eastern Australia. Tectonophysics, 235, 249-275.
Collins, W.J., Beams, S.D., White, A.J.R. and Chappell, B.W., 1982. Nature and origin of A-type
granites with particular reference to SE Australia. Contributions to Mineralogy and Petrology 80,
189-200.
Collins, W. J., Offler, R., Farrel, T.R. and Landenberger, B. 1993. A revised Late Palaeozoic - Early
Mesozoic tectonic history for the southern New England Fold Belt. In: Flood, P. G. and Aitchison, J.
C. (eds.) New England Orogen, Eastern Australia, pp. 69-84. Department of Geology and
Geophysics, University of New England, Armidale.
Colquhoun, G.P., Meakin, N.S. and Cameron, R.G. (compilers). 2005. Cargelligo 1:250 000
rd
Geological Sheets SHI55-6, 3 edition, Explanatory Notes, Geological Survey of New South
Wales, Maitland, NSW, 291 pp.
Coney, P.J., 1992. The Lachlan belt of eastern Australia and Circum-Pacific tectonic evolution. In: C.L.
Fergusson and R.A. Glen (Editors), The Palaeozoic Eastern Margin of Gondwanaland: Tectonics of
the Lachlan Fold Belt, Southeastern Australia and Related Orogens. Tectonophysics, 14, l-25.
Cooney, P.M. and Mantaring, A.M., 2007. The petroleum potential of the Darling Basin. In Munson,
T.J. and Ambrose, G.J. (editors) Proceedings of the Central Australian Basins Symposium (CABS),
Alice Springs, Northern Territory, 16–18 August, 2005. Northern Territory Geological Survey,
Special Publication 2, pp. 1-20.
Cowley, W. 2007. Summary Geology of South Australia. Earth resources information sheet (M51).
Crawford, A. J. and Berry, R. F., 1992. Tectonic implications of Late Proterozoic - Early Palaeozoic
igneous rock associations in western Tasmania. Tectonophysics 214, 37-56.
Crawford, A. J. and Keays, R. R. 1987. Petrogenesis of Victorian Cambrian tholeiites and implications
for the origin of associated boninites. Journal of Petrology 28, 1075-1109.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 173


Crawford, A. J., Cameron, W. E. and Keays, R. R., 1984. The association boninite-low-Ti andesite-
tholeiite in the Heathcote Greenstone Belt, Victoria: ensimatic setting for the Early Lachlan Fold
belt. Australian Journal of Earth Sciences 31, 161-177.
Crawford, A. J., Donaghy, A. G., Black, L. P. and Stuart-Smith, P. G. 1996. Mount Read Volcanics
correlatives in western Victoria: a new exploration opportunity. Australian Institute of Geoscience
Bulletin 20, 97-102.
Crawford, A.J., Stevens, B.P.J. and Fanning, C.M. 1997. Geochemistry and tectonic setting of some
Neoproterozoic and Early Cambrian volcanics in western New South Wales. Australian Journal of
Earth Sciences 44, 831-852.
Crawford, A. J., Meffre, S. and Symonds, P. A. 2003a. 120 to 0 Ma tectonic evolution of the southwest
Pacific and analogous geological evolution of the 600–220 Ma Tasman Fold Belt System. In: Hillis,
R. R. and Müller, R. D. (eds) Evolution and Dynamics of the Australian Plate, pp. 383–403.
Geological Society of Australia Special Publication 22 and Geological Society of America Special
Paper 372.
Crawford, A.J., Cayley, R.A., Taylor, D.H., Morand, V.J., Gray, C.M., Kemp, A.I.S., Wohlt, K.E.,
VandenBerg, A.H.M., Moore, D.H., Maher, S., Direen, N.G., Edwards, J., Donaghy, A.G.,
Anderson, J.A. and Black, L.P. 2003b. Neoproterozoic and Cambrian continental rifting, continent-
arc collision and post-collisional magmatism. In: Birch, W.D. (ed) Geology of Victoria, pp. 73-94.
Geological Society of Australia Special Publication 23. Geological Society of Australia (Victoria
Division).
Crawford, A. J., Glen, R. A., Cooke, D. R. and Percival, I.G., Guest Editors, 2007a. Geological
evolution and metallogenesis of the Ordovician Macquarie Arc, Lachlan Orogen, New South Wales.
Australian Journal of Earth Sciences, 54, 137-141.
Crawford, A. J., Meffre, S., Squire, R. J., Barron, L. M. and Falloon, T. J. 2007b. Middle and Late
Ordovician magmatic evolution of the Macquarie Arc, Lachlan Orogen, New South Wales.
Australian Journal of Earth Sciences 54, 181-214.
Cross, A.J., Purdy, D.J., Bultitude, R.J., Dhnaram, C.R. and Von Gnielinski, F.E., 2009. Joint GSQ-GA
NGA geochronology project, New England Orogen and Drummond Basin, 2008. Queensland
Geological Record 2009/03.
Cross, A.J., Bultitude, R.J. and Purdy, D.J.,2012, Summary of results for the joint GSQ–GA
geochronology project: Ayr, Bowen, Eulo, Mount Coolon, Proserpine and Warwick 1:250 000 sheet
areas, record no. 2012/19, Geological Survey of Queensland, Brisbane.
Cross, A.J., Dunkley, D.J., Bultitude, R.J., Brown, D.D., Purdy, D.J., Withnall, I.W., von Gnielinski, F.E.
and Blake, P.R., 2015a. Summary of results Joint GSQ–GA geochronology project: Thomson
Orogen, New England Orogen and Mount Isa region, 2010–2012. Geological Survey of
Queensland, Queensland Geological Record 2015/01.
Cross, A.J., Purdy, D.J., Bultitude, R.J., Brown, D. and Carr, P.A., 2015b. Summary of results – joint
GSQ-GA geochronology project: Betoota, Bowen, Brighton Downs, Cape Weymouth, Cooktown,
Durham Downs, Jundah, Mossman, Surat, Tickalara and Winton 1:250 000 Sheet areas.
Geological Survey of Queensland, Queensland Geological Record.
Cross, K. C., Fergusson, C. L. and Flood, P. G. 1987. Contrasting structural styles in the Palaeozoic
subduction complex of the Southern New England Orogen, Eastern Australia. In Leitch E.C. and
Scheibner E. (eds.) Terrane Accretion and Orogenic Belts. American Geophysical Union
Geodynamics Series 19, 83-92.
Crouch, S.B.S., Blake, R.R. and Withnall, I.W. 1995. Geochemistry of the plutonic and volcanic rocks
of the southern Anakie Inlier. In: Withnall, I.W. et al. (eds) Geology of the Southern part of the
Anakie Inlier, central Queensland. Queensland Geology 7.
Davis, B.K. 1994. Synchronous syntectonic granite emplacement in the South Palmer River region,
Hodgkinson Province, Australia. Australian Journal of Earth Sciences 41, 91-103.
Davis, B.K., Lindsay, M. and Hippertt, J.F.M. 1996. Gold mineralization in the northern Hodgkinson
Province, North Queensland, Australia. James Cook University, Department of Earth Sciences,
Townsville, Queensland, Australia. Economic Geology Research Unit News 14.
Day, R. W., Murray, C. G. and Whitaker, W. G. 1978. The eastern part of the Tasman Orogenic Zone.
Tectonophysics 48, 327-364.

174 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Denaro, T.J., Kyriazis, Z., Fitzell, M.J., Morwood, D.A. and Burrows, P.E., 2004. Mines, mineralisation
and mineral exploration in the northern Drummond Basin, central Queensland. Queensland
Geological Record 2004/6.
Direen, N. G. and Crawford, A.J. 2003. The Tasman Line; where is it, what is it, and is it Australia's
Rodinian breakup boundary? Australian Journal of Earth Sciences 50, 491-502.
Donchak et al. 2013. Donchak, P.J.T., Purdy, D.J., Withnall, I.W., Blake, P.R. and Jell, P.A., with
contributions from Bultitude, R.J., Champion, D.C., Cranfield, L.C., Cross, A.J., Draper, J.J.,
Holcombe,R.J. and McKellar, J.L., 2013. Chapter 5. New England Orogen. In P.A. Jell (ed.)
Geology of Queensland. Geological Survey of Queensland, pp. 305–472.
Draper, J. J. (ed) 2002a. Geology of the Cooper and Eromanga Basins. Queensland. Queensland
Minerals and Energy Review Series Queensland Department of Natural Resources and Mines.
Draper, J. J. 2002b. Geological Setting. In: Draper, J.J. (ed.) Geology of the Cooper and Eromanga
Basins, Queensland. Queensland Department of Natural Resources, Mines and Energy, p. 27-29.
Draper, J.J. 2006. The Thomson Fold Belt in Queensland revisited. Australian Earth Sciences
Convention 2006. Melbourne, Australia, 2–6 July 2006. Extended abstracts (DVD ROM).
Draper, J.J. and Bain, J.H.C. 1997. A brief geological history of North Queensland. In: Bain, J.H.C.
and Draper, J.J (eds). North Queensland Geology. Bulletin of the Australian Geological Survey
Organisation 240, and Queensland Department of Mines and Energy Geology 9, p 547-550.
Draper, J. J. and McKellar, J.L. 2002. Cooper Basin tectonics. In: Draper, J.J. (ed.) Geology of the
Cooper and Eromanga Basins, Queensland. Queensland Department of Natural Resources, Mines
and Energy, p. 27-29.
Draper, J. J., Boreham, C.J., Hoffmann, K.L. and McKellar, J.L. 2004. Devonian petroleum systems in
Queensland. PESA’s Eastern Australasian Basins Symposium II, Adelaide.
Drexel, J.F. and Preiss, W.V. (eds), 1995. The Geology of South Australia. Volume 2, The
Phanerozoic. South Australia Geological Survey, Bulletin 54.
Evans, P.R., 1980. Geology of the Galilee Basin. In: R.A. Henderson and P.J. Stephenson (eds.) The
Geology and Geophysics of Northeastern Australia. Geological Society of Australia, Queensland
Division. pp. 299–305.
Evans, P.R., Hoffmann, K.L., Remus, D.A. and Passmore, V.L. 1990. Geology of the Eromanga sector
of the Eromanga-Brisbane Geoscience Transect. In: Finlayson, D.M. (ed.) The Eromanga-Brisbane
Geoscience Transect: a guide to basin development across Phanerozoic Australia in southern
Queensland. Bureau of Mineral Resources, Australia Bulletin 232, 83-104.
Fanning, C. M., Leitch, C. E. and Watanabe, T. 2002. An updated assessment of the SHRIMP U-Pb
Zircon dating of the Attunga eclogite in New South Wales, Australia: relevance to the Pacific
Margin of Gondwana. International Symposium on the Amalgamation of Precambrian Blocks and
the Role of the Palaeozoic Orogens in Asia (Sapporo, 2002).
Fenton, M.W. and Jackson, K.S. 1989. The Drummond Basin: low cost exploration in a high risk area.
The APEA Journal 29, 220-234.
Fergusson, C.L. 1991. Thin-skinned thrusting in the northern New England Orogen, central
Queensland, Australia. Tectonics 10, 797-806.
Fergusson, C.L. 2003. Ordovician-Silurian accretion tectonics of the Lachlan Fold Belt, southeastern
Australia. Australian Journal of Earth Sciences 50, 475-490.
Fergusson, C.L., 2009. Tectonic evolution of the Ordovician Macquarie Arc, central New South Wales:
arguments for subduction polarity and anticlockwise rotation. Australian Journal of Earth Sciences
56, 179–193.
Fergusson, C.L., 2010. Plate-driven extension and convergence along the East Gondwana active
margin: Late Silurian–Middle Devonian tectonics of the Lachlan Fold Belt, southeastern Australia,
Australian Journal of Earth Sciences 57, 627-649, DOI: 10.1080/08120099.2010.494767.
Fergusson, C.L., 2014. Late Ordovician to mid-Silurian Benambran subduction zones in the Lachlan
Orogen, southeastern Australia. Australian Journal of Earth Sciences 61, 587-606.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 175


Fergusson, C.L. and Leitch, E.C., C. 1993. Late Carboniferous to Early Triassic tectonics of the New
England Fold Belt, eastern Australia. In: Flood P.G. & Aitchison J.C. (eds.) New England Orogen,
Eastern Australia, NEO 93 Conference Proceedings, pp. 53–57. University of New England,
Armidale.
Fergusson, C.L., Henderson, R.A., Leitch, E.C. and Ishiga, H. 1993. Lithology and structure of the
Wandilla terrane, Gladstone-Yeppoon district, central Queensland, and an overview of the
Palaeozoic subduction complex of the New England Fold Belt. Australian Journal of Earth Sciences
40, 403-414.
Fergusson, C.L., Carr, P.F., Fanning, C.M. and Green, T.J. 2001. Proterozoic-Cambrian detrital zircon
and monazite ages from the Anakie Inlier, central Queensland; Grenville and Pacific-Gondwana
signatures. Australian Journal of Earth Sciences 48, 857-866.
Fergusson, C.L. and VandenBerg, A.H.M., 2003. Ordovician. The development of craton-derived
deep-sea turbidite successions. In Birch, W.D. (ed.) Geology of Victoria. Geological Society of
Australia Special Publication 23, pp. 95-115.
Fergusson, C. L., Henderson, R. A., Lewthwaite, K. J., Phillips, D. and Withnall, I. W. 2005. Structure
of the Early Palaeozoic Cape River Metamorphics, Tasmanides of north Queensland: evaluation of
the roles of convergent and extensional tectonics. Australian Journal of Earth Sciences 52, 261-
277.
Fergusson, C. L., Henderson, R. A., Withnall, I. W. and Fanning, C. M. 2007a. Structural history of the
Greenvale Province, north Queensland: Early Palaeozoic extension and convergence on the
Pacific margin of Gondwana. Australian Journal of Earth Sciences 54, 573-595.
Fergusson, C. L., Henderson, R. A., Withnall, I. W., Fanning, C. M., Phillips, D. and Lewthwaite, K. J.
2007b. Structural, metamorphic, and geochronological constraints on alternating compression and
extension in the Early Palaeozoic Gondwanan Pacific margin, northeastern Australia, Tectonics 26,
TC3008, doi:10.1029/2006TC001979.
Fergusson, C.L., Henderson, R.A., Fanning, C.M. and Withnall, I.W. 2007c. Detrital zircon ages in
Neoproterozoic to Ordovician siliciclastic rocks, northeastern Australia; implications for the tectonic
history of the east Gondwana continental margin. Journal of the Geological Society of London 164,
215-225.
Fergusson, C.L., Nutman, A.P., Kamiichi, T. and Hidaka, H., 2013. Evolution of a Cambrian active
continental margin: The Delamerian–Lachlan connection in southeastern Australia from a zircon
perspective. Gondwana Research 24, 1051–1066.
Fergusson, C.L., Offler, R. and Green, T.J., 2009. Late Neoproterozoic passive margin of East
Gondwana: Geochemical constraints from the Anakie Inlier, central Queensland, Australia.
Precambrian Research 168, 301–312.
Fergusson, C.L. and Henderson, R.A., with contributions from Blake, P.R., Bultitude, R.J., Champion,
D.C., Cross, A.J., Draper, J.J., Green, T.J., Hutton, L.J., Jell, P.A., Lister, G.S., McKellar, J.L.,
McKillop, M.D., Mulqueeny, L., Offler, R., Phillips, G., Richards, S., Withnall, I.W., Wood, D.G. and
Wormald, R., 2013. Chapter 3. Thomson Orogen. In Jell, P.A. (ed.) Geology of Queensland.
Geological Survey of Queensland.
Fielding, C.R., Falkner, A.J., Kassan, J. and Draper, J.J. 1990. Permian and Triassic depositional
systems in the Bowen Basin, eastern Queensland. In: Beeston, J.W. (compiler) Bowen Basin
Symposium 1990, Mackay, Queensland, Australia, Sept. 1990. Geological Society of Australia
Bowen Basin Geology Group Australia.
Finlayson, D.M., 1993. Crustal architecture across Phanerozoic Australia along the Eromanga-
Brisbane Geoscience Transect: evolution and analogues. Tectonophysics 219, 191–211.
Finlayson, D.M., Leven, J.H., Wake-Dyster, K.D. and Johnstone, D.W., 1990. A crustal image under
the basins of southern Queensland along the Eromanga–Brisbane Geoscience Transect. In:
Finlayson D.M. (ed.) The Eromanga—Brisbane Geoscience Transect: a guide to basin
development across Phanerozoic Australia in southern Queensland. Bureau of Mineral Resources,
Geology and Geophysics Bulletin 232, 153–175.
Flood, P. G. and Aitchison, J. C. 1988. Tectonostratigraphic terranes of the southern part of the New
England Orogen. In: Kleeman, J.D. (ed.) New England Orogen Tectonics and Metallogenesis, pp.
7-10. University of New England, Armidale.

176 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Flood, P. G. and Aitchison, J. C. 1992. Late Devonian accretion of the Gamilaroi terrane to eastern
Gondwana: provenance linkage suggested by the first appearance of Lachlan Fold Belt-derived
quartzite. Australian Journal of Earth Sciences 39, 539-544.
Foden, J., Barovich, K., Jane, M. and O’Halloran, G. 2001. Sr-isotopic evidence for Late
Neoproterozoic rifting in the Adelaide Geosyncline at 586 Ma: implications for a Cu ore forming
fluid flux. Precambrian Research 106, 291-308.
Foden, J.D., Elburg, M.A., Turner, S.P., Sandiford, M., O’callaghan, J. and Mitchell, S. 2002a. Granite
production in the Delamerian Orogen, South Australia. Journal of the Geological Society London
159, 557–575.
Foden, J., Hwan Song, S., Turner, S., Elburg, M., Smith, P.B., Van der Steldt, B. and Van Penglis, D.
2002b. Geochemical evolution of lithospheric mantle beneath S.E. South Australia. Chemical
Geology 182, 663–695.
Foden, J., Elburg, M.A., Dougherty-Page, J. and Burtt, A. 2006. The timing and duration of the
Delamerian Orogeny; correlation with the Ross Orogen and implications for Gondwana assembly.
Journal of Geology 114, 189-210.
Fraser, G.L., Gilmore, P.J., Fitzherbert, J.A., Trigg, S.J., Campbell, L.M., Deyssing, L., Thomas, O.D.,
Burton, G.R., Greenfield, J.E., Blevin, P.L. & Simpson, C.J. 2014. New SHRIMP U-Pb zircon ages
from the Lachlan, southern Thomson and New England orogens, New South Wales: February
2011–June 2013. Record 2014/53. Geoscience Australia, Canberra; Report GS2014/0829.
Geological Survey of New South Wales, Maitland. http://dx.doi.org/10.11636/Record.2014.053.
Fukui, S. 1991. K-Ar age study of metamorphic rocks from the New England Fold Belt in eastern
Australia. Bending of the Great Serpentine Belt and Tectonic History of the Arc-type Crust around
the Belt, Eastern Australia. Preliminary Report on the Geology of the New England Fold Belt 2, 23-
37.
Fukui, S., Watanabe, T., Itaya, T. and Leitch, E. C. 1995. Middle Ordovician high PT metamorphic
rocks in eastern Australia: evidence from K-Ar ages. Tectonics 14, 1014-1020.
Garrad, P.D. and Bultitude, R.J. 1999. Geology, mining history and mineralisation of the Hodgkinson
and Kennedy Provinces, Cairns Region, North Queensland. Queensland Minerals and Energy
Review Series, Queensland Department of Mines and Energy, 305p.
Gatehouse, C. G. 1986. The geology of the Warburton Basin in South Australia. Australian Journal of
Earth Sciences 33, 161-180.
Gatehouse, C. G. and Cooper, B.J. 1986. The subsurface Warburton Basin: an interpretation from
limited information. Sediments down-under; 12th international sedimentological congress,
Canberra, Australia, Aug. 24-30, 1986, p. 116. Bureau of Mineral Resources, Canberra, Australia.
Gatehouse, C.G., Fanning, C.M. and Flint, R.B. 1995. Geochronology of the Big Lake Suite,
Warburton Basin, northeastern South Australia. Quarterly Geological Notes - Geological Survey of
South Australia 128, 8-16.
Gibson, G.M., Morse, M.P., Ireland, T.R. and Nayak, G.K., 2011. Arc–continent collision and
orogenesis in western Tasmanides: Insights from reactivated basement structures and formation of
an ocean–continent transform boundary off western Tasmania. Gondwana Research 19, 608–627.
Gibson, G.M., Champion, D.C. and Ireland, T.R., 2015. Preservation of a fragmented late
Neoproterozoic–earliest Cambrian hyper-extended continental-margin sequence in the Australian
Delamerian Orogen. Geological Society, London, Special Publications, 413,
http://dx.doi.org/10.1144/SP413.8
Gilmore, P. Greenfield, J., Reid, W. and Mills, K. 2007. Metallogenesis of the Koonenberry Belt. In:
Lewis, P.C. (ed). Mines and Wines 2007 Extended Abstracts. Australian Institute of Geoscientists
Bulletin 46, 49-64.
Gilmore, P., Trigg, S., Campbell, L., Hegarty, R., Quinn, C. and Fraser, F., 2012. Coolabah, copper
and conodonts An update on regional mapping projects in western NSW. Exploration in the house
2012 presentation. Geological Survey of New South Wales.
Glass, L.M. and Phillips, D., 2006. The Kalkarindji continental flood basalt province: A new Cambrian
large igneous province in Australia with possible links to faunal extinctions. Geology 34, 461–464,
http://dx.doi.org/10.1130/G22122.1.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 177


Glen, R. A. 2004. Plate tectonics of the Lachlan Orogen: a framework for understanding its
metallogenesis. Geological Society of Australia Abstracts 74, 33–36.
Glen, R. A. 2005. The Tasmanides of eastern Australia. In: Vaughan A. P. M., Leat P. T. and
Pankhurst R. J. eds. Terrane Processes at the Margins of Gondwana. Geological Society of
London Special Publication 246, 23-96
Glen, R.A., 2013. Refining accretionary orogen models for the Tasmanides of eastern Australia.
Australian Journal of Earth Sciences 60, 315-370.
Glen, R.A., Walshe, J L., Barron, L.M. and Watkins, J.J., 1998. Ordovician convergent-margin
volcanism & tectonism in the Lachlan sector of east Gondwana. Geology 26, 751–754.
Glen, R.A., Stewart I. R. and Percival I. G. 2004. Narooma Terrane: implications for the construction of
the outboard part of the Lachlan Orogen. Australian Journal of Earth Sciences 51, 859-884.
Glen, R.A., Crawford, A. J., Percival, I. G. and Barron, L. M. 2007a. Early Ordovician development of
the Macquarie Arc, Lachlan Orogen, New South Wales. Australian Journal of Earth Sciences 54,
167-179.
Glen, R.A., Meffre, S. and Scott, R. J. 2007b. Benambran Orogeny in the Eastern Lachlan Orogen,
Australia. Australian Journal of Earth Sciences 54, 385-415.
Glen, R.A., Poudjom-Djomani, Y., Korsch, R.J., Costello, R.D. and Dick, S. 2007c. Thomson-Lachlan
seismic project - results and implications. In: Lewis, P.C. (Ed.), Mines and Wines 2007 Extended
Abstracts, Bulletin 46, p. 73-78, Australian Institute of Geoscientists, Perth.
Glen, R.A., Percival, I.G. and Quinn, C.D., 2009. Ordovician continental margin terranes in the
Lachlan Orogen, Australia: Implications for tectonics in an accretionary orogen along the east
Gondwana margin, Tectonics 28, TC6012, doi:10.1029/2009TC002446.
Glen, R.A., Saeed, A., Quinn, C.D. and Griffin, W.L., 2011. U–Pb and Hf isotope data from zircons in
the Macquarie Arc, Lachlan Orogen: Implications for arc evolution and Ordovician
palaeogeography along part of the east Gondwana margin. Gondwana Research 19, 670–685
Glen, R.A., Korsch, R.J., Hegarty, R., Saeed, A., Poudjom Djomanii, Y., Costelloe, R.D. and
Belousova, E., 2013. Geodynamic significance of the boundary between the Thomson Orogen and
the Lachlan Orogen, northwestern New South Wales and implications for Tasmanide tectonics.
Australian Journal of Earth Sciences 60, 371–412,
http://dx.doi.org/10.1080/08120099.2013.782899
Golding, S. D., Messenger, P. R., Dean, J. A., Perkins, C., Huston, D. A. and White, A. H. 1994. Mount
Morgan gold - copper deposit: geochemical constraints on the source of volatiles and lead and the
age of mineralization. In: Henderson, R. A. and Davis, B. K. (eds.) Extended Conference Abstracts
New Developments in Geology and Metallogeny: Northern Tasman Orogenic Zone, pp. 89-95.
James Cook University Economic Geology Research Unit Contribution 50.
Gravestock, D. I. and Gatehouse, C.G. 1995. Eastern Warburton Basin. In: Preiss, J.F. and Drexel, W.
V. (Eds) The geology of South Australia, Volume 2 - The Phanerozoic. Mines and Energy South
Australia.
Gravestock, D.I. and Jensen-Schmidt, B., 1998. Chapter 5. Structural Setting. In Gravestock, D.I.,
Hibburt, J. and Drexel, J.F. (eds.) Petroleum geology of South Australia Volume 4: Cooper Basin.
Primary Industries and Resources South Australia Report Book 98/9, 47-68.
Gray, A. R. G. and McKellar, J.L. 2002. Cooper Basin stratigraphy. In: Draper, J.J. (ed.) Geology of
the Cooper and Eromanga Basins, Queensland. Queensland Minerals and Energy Review Series,
Department of Natural Resources and Mines.
Gray, D.R. 1997. Tectonics of the southeastern Australian Lachlan Fold Belt: structural and thermal
aspects. In, Burg, J.-P. and Ford, M. (eds), Orogeny Through Time, Geological Society Special
Publication 121, pp. 149-177.
Gray, D.R. and Foster, D.A. 1997. Orogenic concepts - application and definition: Lachlan Fold Belt,
eastern Australia. American Journal of Science 297, 859-891.
Gray, D.R. and Foster, D.A. 2004. Tectonic review of the Lachlan Orogen, southeast Australia:
historical review, data synthesis and modern perspectives. Australian Journal of Earth Sciences 51,
773-817.

178 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Gray, D.R., Foster, D.A. and Bucher, M., 1997. Recognition and definition of orogenic events in the
Lachlan Fold Belt. Australian Journal of Earth Sciences 44, 489–501.
Gray, C.M., Kemp, A.I.S., Anderson, A.C., Bushell, D.J., Ferguson, D.J., Fitzherbert, J. and stevenson,
M.D., 2002. Delamerian Glenelg tectonic zone, western Victoria: geology and metamorphism of
stratiform rocks. Australian Journal of Earth Sciences 49, 187–200.
Gray, D.R., Foster, D.A., Morand, V.J., Willman, C.E., Cayley, R.A., Spaggiari, C.V., Taylor, D.H.,
Gray, C.M., VandenBerg, A.H.M., Hendrickx, M.A. and Wilson, C.J.L., 2003. Structure,
metamorphism, geochronology and tectonics of Palaeozoic rocks – interpreting a complex, long
lived orogenic system. In Birch, W.D. (ed.) Geology of Victoria. Geological Society of Australia
Special Publication 23, pp. 15–71.
Green, P.M., Hoffmann, K.L., Brain, T.J. and Gray, A.R.G. 1997a. Project aims and activities,
exploration history and geological investigations in the Bowen and overlying Surat Basins,
Queensland. In: Green, P.M. (ed.) The Surat and Bowen Basins, south-east Queensland.
Queensland Minerals and Energy Review Series, Vol.1997, p.1–11. Queensland Department of
Minerals and Energy, Brisbane.
Green, P.M., Carmichael, D.C., Brain, T.J., Murray, C.G., McKellar, J.L., Beeston, J.W. and Gray,
A.R.G. 1997b. Lithostratigraphic units in the Bowen and Surat Basin, Queensland. In: Green, P.M.
(ed) The Surat and Bowen Basins, south-east Queensland. Queensland Minerals and Energy
Review Series, Vol.1997, p.41–108. Queensland Department of Minerals and Energy, Brisbane.
Green, T. J., Fergusson, C. L. and Withnall, I. W., 1998. Refolding and strain in the Neoproterozoic -
Early Palaeozoic Anakie Metamorphic Group, Central Queensland. Australian Journal of Earth
Sciences, 45, 915–924.
Greene, D.C., 2010. Neoproterozoic rifting in the southern Georgina Basin, central Australia:
Implications for reconstructing Australia in Rodinia. Tectonics 29, TC5010,
doi:10.1029/2009TC002543.
Greenfield, J.E., Gilmore, P.J. and Mills, K.J. compilers, 2010. Explanatory notes for the Koonenberry
Belt geological maps. Geological Survey of New South Wales Bulletin 35.
Greenfield, J.E., Musgrave, R.G., Bruce, M.C., Gilmore, P.K. and Mills, K.J., 2011. The Mount Wright
Arc: a Cambrian subduction system developed on the continental margin of East Gondwana,
Koonenberry Belt, eastern Australia. Gondwana Research 19, 650–669.
Grimes, K.G., Hutton, L.J., Law, S. R., McLennan, T.P.T. and Belcher, R., 1986. Geological mapping
in the Mount Coolon 1:250 000 Sheet area, 1985. Geological Survey of Queensland Record
1986/59.
Gust, D.A., Stephens, C.J. and Grenfell, A.T. 1993. Granitoids of the northern NEO: their distribution
in time and space and their tectonic implications. In: P.G. Flood and J. Aitchison (Eds), New
England Orogen, eastern Australia. University of New England, Armidale, Australia, pp. 565–571.
Haines, P.W. and Wingate, M.T.D., 2007. Contrasting depositional histories, detrital zircon
provenance and hydrocarbon systems: Did the Larapintine Seaway link the Canning and Amadeus
basins during the Ordovician? In Munson TJ and Ambrose GJ (eds.) Proceedings of the Central
Australian Basins Symposium (CABS), Alice Springs, Northern Territory, 16–18 August, 2005.
Northern Territory Geological Survey, Special Publication 2, pp. 36–51.
Haines, P.W., Hand, M. and Sandiford, M., 2001. Palaeozoic synorogenic sedimentation in central and
northern Australia: A review of distribution and timing with implications for the evolution of
intracontinental orogens. Australian Journal of Earth Sciences 48, 911-928, DOI: 10.1046/j.1440-
0952.2001.00909.x.
Hall, R.L. 1975. Late Ordovician coral faunas from north-eastern New South Wales. Journal and
Proceedings of the Royal Society of New South Wales 108, 75–93.
Hall, R.L. 1978. A Silurian (Upper Llandovery) coral fauna from the Woolomin beds near Attunga, New
South Wales. Proceedings of the Linnean Society of New South Wales 102, part 3.
Hamilton, D.D., 1993, Genetic stratigraphic framework. In Tadros, N.Z. (Editor): The Gunnedah Basin,
New South Wales. Geological Survey of New South Wales, Memoir Geology, 12, 145–164.
Harrington, H. J. and Korsch, R. J. 1985. Late Permian to Cainozoic tectonics of the New England
Orogen. Australian Journal of Earth Sciences 32, 181–203.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 179


Harvey, S. C. and J. E. Hibburt, 1999. Prospective Neoproterozoic to Middle Palaeozoic Basins,
Eastern Warburton Basin. Petroleum exploration and development in South Australia (12th edition)
South Australia., Petroleum Exploration Data Package, 004, Department of Primary Industries and
Resources.
Henderson, R.A., 1980. Structural outline and summary geological history for northeastern Australia.
In: Henderson, R.A. and Stephenson, P.J. (eds). The Geology and Geophysics of Northeastern
Australia. Geological Society of Australia, Queensland Division, Townsville. pp. 1-26.
Henderson, R.A. 1986. Geology of the Mt Windsor subprovince - a lower Palaeozoic volcano-
sedimentary terrane in the northern Tasman orogenic zone. Australian Journal of Earth Sciences
33, 343–364.
Henderson, R.A. 1987. An oblique subduction and transform faulting model for the evolution of the
Broken River Province, northern Tasman Orogenic Zone. Australian Journal of Earth Sciences 34,
237–249.
Henderson, R.A., Davis, B.K. and Fanning, C.M., 1998. Stratigraphy, age relationships and tectonic
setting of rift-phase infill in the Drummond Basin, central Queensland. Australian Journal of Earth
Sciences 45, 579–595.
Henderson, R.A., Innes, B.M., Fergusson, C.L., Crawford, A.J. and Withnall, I.W., 2011. Collisional
accretion of a Late Ordovician oceanic island arc, northern Tasman Orogenic Zone, Australia.
Australian Journal of Earth Sciences 58, 1–19.
Henderson, R.A., Donchak, P.J.T. and Withnall, I.W., with contributions from Adams, C.J., Bultitude,
R.J., Champion, D.C., Davis, B.K., Hutton, L.J. and Wormald, R., 2013. Chapter 3. Mossman
Orogen. In P.A. Jell (ed.) Geology of Queensland. Geological Survey of Queensland, pp. 225–304.
Hoffman, P.F. and Schrag, D.P., 2002. The snowball Earth hypothesis: testing the limits of global
change. Terra Nova, 14, 129–155.
Holcombe, R. J., Stephens, C. J., Fielding, C. R., Gust, D., Little, T. A., Sliwa, R., McPhie, J. and
Ewart, A. 1997a. Tectonic evolution of the northern New England Fold Belt: Carboniferous to Early
Permian transition from active accretion to extension. In: Ashley, P. M. and Flood, P. G. (eds.)
Tectonics and Metallogenesis of the New England Orogen. Geological Society of Australia Special
Publication 19, 66–79.
Holcombe, R. J., Stephens, C. J., Fielding, C. R., Gust, D., Little, T. A., Sliwa, R., Kassan, J., McPhie,
J. and Ewart, A. 1997b. Tectonic evolution of the northern New England Fold Belt: the Permian -
Triassic Hunter-Bowen event. In: Ashley, P. M. and Flood, P. G. (eds.) Tectonics and
Metallogenesis of the New England Orogen. Geological Society of Australia Special Publication 19,
52–65.
Holm, O.H., Crawford, A.J. and Berry, R.F. 2003. Geochemistry and tectonic settings of meta-igneous
rocks in the Arthur Lineament and surrounding area, northwest Tasmania. Australian Journal of
Earth Sciences 50, 903–918.
Huston, D.L., Champion, D.C., Mernagh, T.P., Downes, P.M., Jones, P., Carr, G., Forster, D. and
David, V., 2016. Metallogenesis and geodynamics of the Lachlan Orogen: New (and old) insights
from spatial and temporal variations in lead isotopes. Ore Geology Reviews 76, 257–267.
Hutton, L. J., Draper, J.J., Rienks, I.P., Withnall, I.W. and Knutson, J. 1997. Charters Towers region.
In: Bain, J.H.C. and Draper, J.J (eds). North Queensland Geology. AGSO Bulletin 240 and
Queensland Geology 9, 165–224.
Hutton, L.J., Grimes, K.G., Law, S.R. and McLennan, T.P.T., 1998a. Mount Coolon (SF55-7),
Queensland, 1:250 000 geological series explanatory notes, 2nd edition. Geological Survey of
Queensland, Brisbane.
Hutton, L., Fanning, C.M. and Withnall, I.W. 1998b. The Cape River area - evidence for Late
Mesoproterozoic and Neoproterozoic crust in north Queensland. Abstracts of the Geological
Society of Australia 48, 216.
Ickert, R.B. and Williams, I.S., 2011. U–Pb zircon geochronology of Silurian–Devonian granites in
southeastern Australia: implications for the timing of the Benambran Orogeny and the I–S
dichotomy. Australian Journal of Earth Sciences 58, 501–516.

180 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Jackson, K.S., Horvath, Z. and Hawkins, P.J. 1981. An assessment of the petroleum prospects for the
Galilee Basin, Queensland. The APEA Journal, Vol.21 (Part 1), p.172-186; Petroleum; back to
basics, Adelaide, Australia, April 5-8, 1981.
Jell, P.A. (editor), 2013. Geology of Queensland. Geological Survey of Queensland.
Jenkins, R.B., Landenberger, B. and Collins, W.J. 2002. Late Palaeozoic retreating and advancing
subduction boundary in the New England Fold Belt, New South Wales. Australian Journal of Earth
Sciences 49, 467-489.
Johnson, S.E. and Henderson, R.A. 1991. Tectonic development of the Drummond Basin, eastern
Australia: back arc extension and inversion in a Late Palaeozoic active margin setting. Basin
Research 3, 197-213.
Johnston, A. J., Offler, R. and Liu, S. 2002. Structural fabric evidence for indentation tectonics in the
Nambucca Block, southern New England Fold Belt, New South Wales. Australian Journal of Earth
Sciences 49, 407-421.
Kendall, B., Creaser, R.A, Calver, C.R. and Evans, D.A.D. 2007. Neoproterozoic paleogeography,
Rodinia breakup, and Sturtian glaciation: Constraints from Re-Os black shale ages from southern
Australia and northwestern Tasmania. Abstract, Geological Society of America.
Kennett, B.L.N., Fishwick, S., Reading, A.M. and Rawlinson, N., 2004. Contrasts in mantle structure
beneath Australia – relation to Tasman Lines? Australian Journal of Earth Sciences 51, 563–569.
Kent, A.J.R. 1994. Geochronology and geochemistry of Palaeozoic intrusive rocks in the Rockvale
region, southern New England Orogen, New South Wales. Australian Journal of Earth Sciences 41,
365-379.
Kimbrough, D.L., Cross, K.C. and Korsch, R.J. 1993. U-Pb isotope ages for zircons from the Pola
Fogal and Nundle granite suites, southern New England Orogen. In: Flood, P. and Aitchison, J. C.
(eds) New England Orogen, Eastern Australia Conference Proceedings, University of New
England, p. 403-412.
Kirkegaard, A.G. 1974. Structural elements of the northern part of the Tasman Geosyncline. The
Tasman Geosyncline; A Symposium, p.47-63. Geological Society of Australia, Queensland
Division, Brisbane.
Korsch, R.J. 1977. A framework for the Palaeozoic geology of the southern part of the New England
Geosyncline. Journal of the Geological Society of Australia 23, 339-355.
Korsch, R.J. 1984. Sandstone compositions from the New England Orogen, eastern Australia:
implications for tectonic setting. Journal of Sedimentary Petrology 54, 192-211.
Korsch, R.J. 2004. A Permian-Triassic retroforeland thrust system - the New England Orogen and
adjacent sedimentary basins, eastern Australia. In: McClay, K.R. (ed.) Thrust tectonics and
hydrocarbon systems. AAPG Memoir 82, 515-537.
Korsch, R.J. and Harrington, H.J. 1987. Oroclinal bending, fragmentation and deformation of terranes
in the New England Orogen, eastern Australia. In: Leitch, E. C. and Scheibner E. (eds). Terrane
Accretion and Orogenic Belts. American Geophysical Union Geodynamics Series 19, 129-139.
Korsch, R.J. and Totterdell, J.M. 1995. Structural events and deformational styles in the Bowen Basin.
In: Follington, I. W., Beeston, J. W. and Hamilton, L. H. (eds.) Bowen Basin Symposium 1995...150
Years On...Proceedings, pp. 27-35. Geological Society of Australia, Coal Geology Group,
Brisbane.
Korsch, R. J., Harrington, H. J., Murray, C. G., Fergusson, C. L. and Flood, P. G. 1990. Tectonics of
the New England Orogen. In: Finlayson, D. M. (ed.) The Eromanga- Brisbane Geoscience
Transect: a Guide to Basin Development across Phanerozoic Australia in southern Queensland.
Bureau of Mineral Resources Bulletin 232, 35-52.
Korsch, R.J., Wake-Dyster, K.D. and Johnstone, D.W. 1993. The Gunnedah Basin-New England
Orogen deep seismic reflection profile: implications for New England tectonic. In: Flood, P.G and
Aitchison, J.C. (eds) New England Orogen, eastern Australia conference. University of New
England, Department of Geology and Geophysics, Armidale, N.S.W., Australia, p. 85-100.
Korsch, R. J., Johstone, D. W. and Wake-Dyster, K. D. 1997. Crustal architecture of the New England
Orogen based on deep seismic reflection profiling. In: Tectonics and Metallogenesis of the New
England Orogen. Geological Society of Australia Special Publication 19, 29-51.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 181


Korsch, R.J., Boreham, C.J., Totterdell, J.M., Shaw, R.D. and Nicoll, M.G. 1998. Development and
petroleum resource evaluation of the Bowen, Gunnedah and Surat Basins, eastern Australia. The
APPEA Journal, 38, 199-236.
Korsch, R.J., Moore, D.H., Cayley, R.A., Costelloe, R.D., Nakamura, A., Willman, C.E., Rawling, T.J.,
Morand, V.J. and O’Shea, P.J. 2008. Crustal architecture of Central Victoria: results from the 2006
deep crustal reflection seismic survey. In: Australian Earth Sciences Convention (AESC) 2008.
th
New Generation Advances in Geoscience. Abstracts No 89 of the 19 Australian Geological
Convention, Perth Convention and Exhibition Centre, Perth WA, July 20-24, p. 155.
Korsch, R. J., Totterdell, J. M., Cathro, D. L. and Nicoll, M. G., 2009a. The Early Permian East
Australian Rift System. Australian Journal of Earth Sciences.
Korsch, R. J., Totterdell, J. M., Fomin, T. and Nicoll, M. G., 2009b. Contractional structures and
deformational events in the Bowen, Gunnedah and Surat Basins, eastern Australia. Australian
Journal of Earth Sciences.
Korsch, R.J., Adams, C.J., Black, L.P., Foster, D.A., Fraser, G.L., Murray, C.G., Foudoulis, C. and
Griffin, W.L., 2009c. Geochronology and provenance of the Late Palaeozoic accretionary wedge
and Gympie Terrane, New England Orogen, eastern Australia. Australian Journal of Earth
Sciences.
Korsch, R.J., Huston, D.L., Henderson, R.A., Blewett, R.S., Withnall, I.W., Fergusson, C.L., Collins,
W.J., Saygin, E., Kositcin, N., Meixner, A.J., Chopping, R., Henson, P.A., Champion, D.C., Hutton,
L.J., Wormald, R., Holzschuh, J. and Costelloe, R.D., 2012. Crustal architecture and geodynamics
of north Queensland, Australia: insights from deep seismic reflection profiling. Tectonophysics 572,
76–99.
Kositcin, N., Purdy, D.J., Brown, D.D., Bultitude, R.J. and Carr, P.A. 2015a. Summary of results —
Joint GSQ–GA Geochronology Project: Thomson Orogen and Hodgkinson Province, 2012–2013.
Queensland Geological Record 2015/02.
Kositcin, N., Bultitude, R.J., Purdy, D.J., Brown, D.D., Carr, P.A. and Lisitsin, V., 2015b. Summary of
Results — Joint GSQ–GA Geochronology Project: Kennedy Igneous Association, Mossman
Orogen, Thomson Orogen and Iron Range Province 2013–2014. Queensland Geological Record
2015/03.
Kositcin, N. and Bultitude, R.J., 2015. Summary of Results — Joint GSQ–GA Geochronology Project:
Hodgkinson Formation of the Mossman Orogen, 2014-2015. Queensland Geological Record
2015/04.
Krassay, A.A., Korsch, R.J. and Drummond, B.J., 2009. Meandarra Gravity Ridge: symmetry elements
of the gravity anomaly and its relationship to the Bowen–Gunnedah–Sydney basin system.
Australian Journal of Earth Sciences 56, 355–79.
Leitch, E.C. 1975. Plate tectonic interpretation of the Palaeozoic history of the New England Fold Belt.
Geological Society of America Bulletin 86, 141-144.
Leitch, E.C. 1988. The Barnard Basin and the early Permian Development of the southern part of the
New England Fold Belt. In: Kleeman, J. D. (ed.) New England Orogen - Tectonics and
Metallogenesis, pp. 61-67. Department of Geology and Geophysics, University of New England,
Armidale.
Leitch, E.C. and Cawood, P.A. 1987. Provenance determination of volcaniclastic rocks: The nature
and tectonic significance of a Cambrian conglomerate from the New England Fold Belt, eastern
Australia. Journal of Sedimentary Petrology 57, 630-638.
Leitch, E.C. and Scheibner, E. 1987. Stratotectonic terranes of the eastern Australian Tasmanides.
American Geophysical Union Geodynamic Series 19, 1-19.
Leitch, E.C., Fergusson, C.L. and Henderson, R.A., 1992. Geological note: the intra-Devonian
unconformity at Mt Gelobera, south of Rockhampton, central Queensland. Australian Journal of
Earth Sciences 39, 121–2.
Lewis, P.C., Glen, R.A., Pratt, G.W. and Clarke, I. 1994. Bega-Mallacoota 1:250 000 Geological Sheet
SJ/55-4, SJ/55-8: Explanatory Notes, Geological Survey of New South Wales, Sydney, 148 pp.
Li, P. and Rosenbaum, G., 2014. Does the Manning Orocline exist? New structural evidence from the
inner hinge of the Manning Orocline (eastern Australia). Gondwana Research 25, 1599–1613.

182 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Li, Z.X. and Powell, C.McA., 2001. An outline of the palaeogeographic evolution of the Australasian
region since the beginning of the Neoproterozoic. Earth-Science Reviews. 53, 237–277.
Li, Z.X., Bogdanova, S.V., Collins, A.S., Davidson, A., De Waele, B., Ernst, R.E., Fitzsimons, I.C.W.,
Fuck, R.A., Gladkochub, D.P., Jacobs, J., Karlstrom, K.E., Lu, S., Natapovm, L.M., Pease, V.,
Pisarevsky, S.A., Thrane, K. and Vernikovsky, V., 2008. Assembly, configuration, and break-up
history of Rodinia: A synthesis. Precambrian Research, 160, 179–210.
Little, T. A., Holcombe, R. J., Gibson, G. M., Offler, R., Gans, P. B. and McWilliams, M. O. 1992.
Exhumation of late Palaeozoic blueschists in Queensland, Australia, by extensional faulting.
Geology 20, 231-234.
Little, T. A., McWilliams, M. O. and Holcombe, R. J. 1995. 40Ar/39Ar thermochronology of epidote-
blueschists from the North D’Aguilar Block, Queensland, Australia: timing and kinematics of
subduction complex unroofing. Geological Society of America Bulletin 107, 520-535.
Lyons, P., Raymond, O.L. and Duggan, M.B. (compiling editors) 2000. Forbes 1:250 000 geological
nd
sheet SI55-7, 2 edition, Explanatory Notes. AGSO Record 2000/20, 230 pp.
Maher, S., Moore, D.H., Crawford, A.J., Twyford, R. and Fanning, C.M., 1997. Test drilling on the
southern margin of the Murray Basin. Victorian Initiative for Minerals and Petroleum Report 52.
Department of Natural Resources and Environment.
Maidment, D. W., Williams, I.S. and Hand, M. 2007. Testing long-term patterns of basin sedimentation
by detrital zircon geochronology, Centralian Superbasin, Australia. Basin Research 19, 335-360.
Mawson, R. and Talent, J.A., 2003. Conodont faunas from sequences on or marginal to the Anakie
Inlier (Central Queensland, Australia) in relation to Devonian transgressions. Bulletin of
Geosciences 78, 335–358.
McConachie, B.A., Dunster, J.N., Wellman, P., Denaro, T.J., Pain, C.F., Habermehl, M.A. and Draper,
J.J. 1997. Carpentaria Lowlands and Gulf of Carpentaria regions. In: Bain, J.H.C. and Draper, J.J.
(eds) North Queensland Geology. AGSO Bulletin 240, 365-397.
McKillop, M.D., McKellar, J.L., Draper, J.J. and Hoffmann, K.L., 2007. The Adavale Basin: Stratigraphy
and depositional environments. In Munson, T.J. and Ambrose, G.J. (editors) Proceedings of the
Central Australian Basins Symposium (CABS), Alice Springs, Northern Territory, 16–18 August,
2005. Northern Territory Geological Survey, Special Publication 2, pp. 1-26.
McPhie, J. 1987. Andean analogue for Late Carboniferous volcanic arc and arc flank environments of
the western New England Orogen, New South Wales, Australia. Tectonophysics 138, 269-288.
nd
Meakin, N.S. and Morgan, E.J. (compilers). 1999. Dubbo 1:250 000 geological sheet SI/55-4, 2
edition, Explanatory Notes. Geological Survey of New South Wales, Sydney, 504 pp.
Meffre, S., Direen, N.G., Crawford, A.J. and Kamenetsky, V. 2004: Mafic Volcanic rocks on King
Island, Tasmania: Evidence for 579 Ma break-up in east Gondwana. Precambrian Research, 135:
177 - 191.
Meffre, S., Scott, R.J., Glen, R.A. and Squire, R.J. 2007. Re-evaluation of contact relationships
between Ordovician volcanic belts and the quartz-rich turbidites of the Lachlan Orogen. Australian
Journal of Earth Sciences 54, 363-383.
Meixner, A.J., Boucher, R.K., Yeates, A.N., Frears, R.A., Gunn, P.J. and Richardson, L.M. 1999.
Interpretation of Geophysical and Geological data sets, Cooper Basin region, South Australia.
AGSO Record 1999/22.
Meixner, T.J., Gunn, P.J., Boucher, R.K., Yeates, T.N., Richardson, L.M., Frears, R.A. 2000. The
nature of the basement to the Cooper Basin region, South Australia. Exploration Geophysics
(Melbourne) 31, 24-32.
Middleton, M.F. and Hunt, J.W. 1989. Influence of tectonics on Permian coal-rank patterns in
Australia. In: Lyons, P.C. and Alpern, B. (eds), Coal; classification, coalification, mineralogy, trace-
element chemistry, and oil and gas potential. International Journal of Coal Geology 13, 391-411.
Miller, J.M. and Gray, D.R. 1997. Subduction-related deformation and the Narooma anticlinorium,
eastern Lachlan Fold belt, southeastern Australia. Australian Journal of Earth Sciences 44, 237-
251.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 183


Miller, J.McL., Phillips,D., Wilson, C.J.L. and Dugdale, L.J., 2005. Evolution of a reworked orogenic
zone: The boundary between the Delamerian and Lachlan fold belts, southeastern Australia.
Australian Journal of Earth Sciences 52, 921–940.
Mills, K.J. and David, V. 2004 The Koonenberry Deep Seismic Reflection Line and Geological
Modelling of the Koonenberry Region, in Western New South Wales. Geological Survey of NSW,
GS 2004/185.
Morand, V. J. 1993a. Stratigraphy and tectonic setting of the Calliope Volcanic Assemblage,
Rockhampton area, Queensland. Australian Journal of Earth Sciences 40, 15-30.
Morand, V. J. 1993b. Structure and metamorphism of the Calliope Volcanic Assemblage: implications
for Middle to Late Devonian orogeny in the northern New England Fold Belt. Australian Journal of
Earth Sciences 40, 257-270.
Moresi, L., Betts, P.G., Miller, M.S. and Cayley, R.A., 2014. Dynamics of continental accretion. Nature
508, 245–248, doi:10.1038/nature13033.
Munker, C and Crawford, A.J., 2000. Cambrian arc evolution along the SE Gondwana active margin:
A synthesis from Tasmania-New Zealand-Australia-Antarctica correlations. Tectonics 19, 415-432.
Murgulov, V., Beyer, E., Griffin, W.L., O’Reilly, S.Y., Walters, S.G. and Stephens, D., 2007. Crustal
evolution in the Georgetown Inlier, North Queensland, Australia: a detrital zircon grain study.
Chemical Geology 245, 198–218.
Murgulov, V., Griffin, W.L. and O’Reilly, S.Y., 2009. Temporal and genetic relationships between the
Kidston gold-bearing Breccia Pipe and the Lochaber Ring Dyke Complex, North Queensland,
Australia: insights from in situ U–Pb and Hf-isotope analysis of zircon. Mineralogy and Petrology
95, 17-45.
Murray, C. G. 1986. Metallogeny and tectonic development of the Tasman Fold Belt System in
Queensland. Ore Geology Reviews 1, 315-400.
Murray, C.G. 1988. Tectonostratigraphic terranes in southeast Queensland. In: Hamilton, L.H. (ed)
th
Field Excursions Handbook for the 9 Australian Geological Convention. Geological Society of
Australia, Queensland Division, Brisbane, 19-51.
Murray, C., 1990. Tasman Fold Belt in Queensland. In F.E. Hughes (ed.) Geology of the Mineral
Deposits of Australia and Papua New Guinea. Monograph 14. Australasian Institute of Mining and
Metallurgy, Vol. 2, 1434-1450.
Murray, C.G. 1994. Basement cores from the Tasman Fold Belt System beneath the Great Artesian
Basin in Queensland. Queensland Geological Record, Dept. of Minerals and Energy, Queensland
Geological Survey, 96.
Murray, C.G. 1997a. Basement terranes beneath the Bowen and Surat Basins, Queensland. In:
Green, P.M. (ed.) The Surat and Bowen Basins south-east Queensland. Queensland Minerals and
Energy Review Series, Queensland Department of Mines and Energy, p. 13-40.
Murray, C.G. 1997b. From geosyncline to fold belt: a personal perspective on the development of
ideas regarding the tectonic evolution of the New England Orogen. In: Ashley, P. M. and Flood, P.
G. (eds.) Tectonics and metallogenesis of the New England Orogen. Geological Society of
Australia, Special Publication 19, 1-28.
Murray, C.G. 2003. Granites of the northern New England Orogen. In: Blevin, P., Jones, M. and
Chappell, B. (eds) Magmas to Mineralisation: The Ishihara Symposium, pp. 101-108. Geoscience
Australia Record 2003/14.
Murray, C.G. 2007. Devonian supra-subduction zone setting for the Princhester and Northumberland
Serpentinites: implications for the tectonic evolution of the northern New England Orogen.
Australian Journal of Earth Sciences 54, 899-925.
Murray, C. G. and Blake, P. R. 2005. Geochemical discrimination of tectonic setting for Devonian
basalts of the Yarrol Province of the New England Orogen, central coastal Queensland: an
empirical approach. Australian Journal of Earth Sciences 52, 993-1034.
Murray, C.G. and Kirkegaard, A.G. 1978. The Thomson Orogen of the Tasman Orogenic Zone.
Tectonophysics 48, 299-325.

184 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Murray, C. G., Fergusson, C. L., Flood, P. G., Whitaker, W. G. and Korsch, R. J. 1987. Plate tectonic
model for the Carboniferous evolution of the New England Fold Belt. Australian Journal of Earth
Sciences 34, 213-236.
Murray, C. G., Blake, P. R., Hutton, L. J., Withnall, I. W., Hayward, M. A., Simpson, G. A. and
Fordham, B. G. 2003. Discussion. Yarrol terrane of the northern New England Fold Belt: forearc or
backarc? Australian Journal of Earth Sciences 50, 271-278.
Neef, G. 2004. Stratigraphy, sedimentology, structure and tectonics of Lower Ordovician and
Devonian strata of South Mootwingee, Darling Basin, western New South Wales. Australian
Journal of Earth Sciences, 51, 15-29.
Neef, G. and Bottrill, R.S., 1991. Early Devonian (Gedinnian) nonmarine strata present in a rapidly
subsiding basin in far western New South Wales, Australia. Sedimentary Geology, 71, 195-212.
Nishiya, T., Watanabe, T., Yokoyama, K. and Kuramoto, Y. 2003. New isotopic constraints on the age
of the Halls Reward Metamorphics, North Queensland, Australia: Delamerian metamorphic ages
and Grenville detrital zircons. Gondwana Research 6, 241-249.
O’Brien, P.E. (coordinator), Bowen, R.L., Thomas, G.A., Craig, M.A. and Holdgate, G.R., 2003.
Permian. Sedimentation around a continental ice sheet. In Birch, W.D. (ed.) Geology of Victoria.
Geological Society of Australia Special Publication 23, pp. 195-215.
Och, D.J., Leitch, E.C. and Caprarelli, G. 2007. Geological units of the Port Macquarie-Tacking Point
tract, northeastern Port Macquarie Block, Mid North Coast region of New South Wales. Geological
Survey of NSW Quarterly Notes 126, 1-19.
Offler, R. 1999. Origin of blueschist “knockers,” Glenrock Station, NSW. In: Flood, P. G. (ed.) New
England Orogen regional geology, tectonics and metallogenesis. Armidale, Australia, University of
New England, p. 35- 44.
Offler, R. and Foster, D. A. 2008. Timing and development of oroclines in the southern New England
Orogen, New South Wales. Australian Journal of Earth Sciences 55, 331-340.
Offler, R. and Gamble, J. 2002. Evolution of an intra-oceanic island arc during the Late Silurian to Late
Devonian, New England Fold Belt. Australian Journal of Earth Sciences 49, 349-366.
Offler, R. and Murray, C.G., 2011. Devonian volcanics in the New England Orogen: tectonic setting
and polarity. Gondwana Research 19, 706–15.
Offler, R. and Shaw, S. 2006. Hornblende Gabbro Block in Serpentinite Melange, Peel-Manning Fault
System, New South Wales, Australia: Lu-Hf and U-Pb Isotopic Evidence for Mantle-Derived, Late
Ordovician Igneous Activity. The Journal of Geology 114, 211-230.
Olgers, F. 1972. Geology of the Drummond Basin, Queensland. Bureau of Mineral Resources,
Geology and Geophysics Bulletin 132, 1-78.
Oversby, B.S., Mackenzie, D.E., McPhie, J., Law, S.R. and Wyborn, D., 1994. The geology of
Palaeozoic volcanic and associated rocks in the Burdekin Falls Dam—“Conway” area, northeastern
Queensland. Australian Geological Survey Organisation Record 1994/21.
Owen, D.E. 2009. How to use stratigraphic terminology in papers, illustrations, and talks. Stratigraphy
6, 106-116,.
Percival, I. G. and Glen, R. A. 2007. Ordovician to earliest Silurian history of the Macquarie Arc,
Lachlan Orogen, New South Wales. Australian Journal of Earth Sciences 54, 143-165.
Percival. I.G., Quinn, C.D. and Glen, R.A., 2011. A review of Cambrian and Ordovician stratigraphy in
New South Wales. Quarterly Notes Geological Survey of New South Wales 137, 1-39.
Perkins, C., Hinman, M.C. and Walshe, J.L. 1994. Timing of mineralization and deformation, Peak Au
mine, Cobar, New South Wales. Australian Journal of Earth Sciences 41, 509-522.
Pogson, D.J. and Glen, R.A., 2006. Eastern Lachlan Orogen Time-Space Plots. In Eastern Lachlan
Orogen Geoscience Database (on CD-ROM) Version 2, Geological Survey of New South Wales,
Maitland, Australia.
Pogson, D.J. and Watkins, J.J. 1998. Bathurst 1:250 000 Geological Sheet SI/55-8: Explanatory
Notes. Geological Survey of New South Wales, Sydney, 430 pp.
Powell, W. and O’Reilly, S., 2007 Metasomatism and sulfide mobility in lithospheric mantle beneath
eastern Australia: Implications for mantle Re–Os chronology. Lithos 94, 132-147.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 185


Preiss, W.V., 2000. The Adelaide Geosyncline of South Australia and its significance in
Neoproterozoic continental reconstruction. Precambrian Research 100, 21–63.
Prendergast, E. and Offler, R., 2012. Underplated seamount in the Narooma accretionary complex,
NSW, Australia. Lithos 154, 224-234.
40 39
Prendergast, E., Offler, R., Phillips, D. and Zwingman, H., 2012. Ar/ Ar and K–Ar ages: Early
Paleozoic metamorphism and deformation in the Narooma accretionary complex, NSW. Australian
Journal of Earth Sciences 58, 21–32.
Purdy, D.J., Carr, P.A. and Brown, D.D., 2013: A review of the geology, mineralisation and geothermal
energy potential of the Thomson Orogen in Queensland. Queensland Geological Record 2013/01.
Quinn, C.D., Percival, I.G., Glen, R.A. and Xiao, W.-J., 2014. Ordovician marginal basin evolution near
the palaeo-Pacific east Gondwana margin, Australia. Journal of the Geological Society, London.
http://dx.doi.org/10.1144/jgs2012-034
Richards, J.R., White, D.A., Webb, A.W. and Branch, C.D. 1966. Radiometric ages of acid igneous
rocks in the Cairns hinterland, north Queensland. Bureau of Mineral Resources Bulletin 88.
Roberts, J. and Engel, B. A. 1980. Carboniferous palaeogeography of the Yarrol and New England
Orogens, eastern Australia. Journal of the Geological Society of Australia 27, 167-186.
Roberts, J. and Geeve, R. 1999. Allochthonous forearc blocks and their influence on an orogenic
timetable for the Southern New England Orogen. In: Flood, P. G. (ed.) New England Orogen -
Regional Geology, Tectonics and Metallogenesis, pp. 105-114. Earth Sciences, University of New
England, Armidale.
Roberts, J., Claoue-Long, J. and Jones, P. J. 1993. Revised correlation of Carboniferous and Early
Permian units of the southern New England Orogen, Australia. Newsletter on Carboniferous
Stratigraphy 11, 23-26.
Roberts, J., Claoue-Long, J., Jones, P. J. and Foster, C. B. 1995. SHRIMP zircon age control of
Gondwana sequences in late Carboniferous and Early Permian Australia. In Dunnay, R. E. and
Hailwood, E. A. (eds.) Nonbiostratigraphical methods of dating and correlation. Geological Society
Special Publication 89, 145-174.
Roberts J., Claoue-Long, J. C. and Foster, C. B. 1996. SHRIMP zircon dating of the Permian System
of eastern Australia. Australian Journal of Earth Sciences 43, 401-421.
Roberts, J., Offler, R. and Fanning, M. 2004. Upper Carboniferous to Lower Permian volcanic
successions of the Carroll-Nandewar region, northern Tamworth Belt, southern New England
Orogen, Australia. Australian Journal of Earth Sciences 51, 205-232.
Roberts, J., Offler, R. and Fanning, M. 2006. Carboniferous to Lower Permian stratigraphy of the
southern Tamworth Belt, southern New England Orogen, Australia: boundary sequences of the
Werrie and Rouchel blocks. Australian Journal of Earth Sciences 53, 249-284.
Rose, G., 1966. Wollongong 1:250 000 Geological Sheet SI/56-9, second edition. Geological Survey
of New South Wales, Sydney.
Rosenbaum, G., Li, P. and Rubatto, D., 2012. The contorted New England Orogen (eastern Australia):
New evidence from U-Pb geochronology of early Permian granitoids. Tectonics 31, TC1006,
doi:10.1029/2011TC002960.
Rosenbaum, G., Li, P. and Rubatto, D., 2012. The contorted New England Orogen (eastern Australia):
New evidence from U-Pb geochronology of early Permian granitoids. Tectonics 31, TC1006,
doi:10.1029/2011TC002960.
Rossiter, A.G. 2003. Granitic rocks of the Lachlan Fold Belt in Victoria – potential keys to the
composition of the lower-middle crust. In Birch, W.D. (ed.) Geology of Victoria. Geological Society
of Australia Special Publication 23, 217-237.
Sano, S., Offler, R., Hyodo, H. and Watanabe, T. 2004. Geochemistry and chronology of tectonic
blocks in serpentinite melange of the southern New England Fold Belt, NSW, Australia. Gondwana
Research 7, 817-831.
Scheibner, E. 1985. Suspect terranes in the Tasman Fold Belt system, eastern Australia. Circum-
Pacific Council for Energy and Mineral Resources, Earth Sciences Series 1, 493-514.
Scheibner, E. and Basden, H. (eds) 1996. Geology of New South Wales - Synthesis. Volume 1
Structural Framework. Geological Survey of New South Wales, Memoir Geology 13, 295 pp.

186 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Scheibner, E. and Basden, H. (eds) 1998. Geology of New South Wales: synthesis. Vol. 2. Geological
evolution. Sydney, Geological Survey of New South Wales, 666 p.
Scheibner, E. and Veevers, J.J., 2000. Tasman fold belt system. In Veevers, J.J. (Ed.) Billion-Year
Earth History of Australia and Neighbours in Gondwanaland. GEMOC Press, Sydney, pp. 154–234.
Scott, S. G., Beeston, J.W. and Carr, A.F. 1995. Galilee Basin. In: C. R. Ward, H. J. Harrington, C. W.
Mallett and J. W. Beeston (eds), Geology of Australian Coal Basins, Sydney. Geological Society of
Australia (Coal Geology Group) Special Publication 1, 341-352.
Seymour, D.B. and Calver, C.R (compilers). 1995. Explanatory notes for the time-space diagram and
stratotectonic elements map of Tasmania: TASGO NGMA project, sub-project 1: geological
synthesis. Mineral Resources Tasmania. Tasmanian Geological Survey record 1995/01, 62p.
Seymour, D.B. and Calver, C.R. 1998. Proterozoic rock sequences of western Tasmania. 14th
Australian geological convention, Townsville, Queensland, Australia, July 6-10, 1998. Geological
Society of Australia 49, 399.
Shaw, R. D. 2002. Bowen and Surat Basins, petroleum data package. R. D. Shaw (editor), Geological
Survey of New South Wales.
Shaw, S. E. and Flood, R. H. 1981. The New England Batholith, eastern Australia: geochemical
variations in space and time. Journal of Geophysical Research 86, B10530-B10544.
Sivell, W. J. and McCulloch, M. T. 1997. Geochemistry and Sm-Nd isotope systematics of Early
Permian basalts from Gympie province and fault basins in southeast Queensland: implications for
mantle sources in a backarc setting at the Gondwana rim. In: Ashley, P. M. and Flood, P. G. (eds)
Tectonics and Metallogenesis of the New England Orogen. Geological Society of Australia Special
Publication 19, 148-160.
Sivell, W. J. and McCulloch, M. T. 2001. Geochemical and Nd-isotopic systematics of the Permo-
Triassic Gympie Group, southeast Queensland. Australian Journal of Earth Sciences 48, 377-393.
Sivell, W. J. and Waterhouse, J. B. 1988. Petrogenesis of Gympie Group volcanics: evidence for
remnants of an early Permian Volcanic arc in eastern Australia. Lithos 21, 81-95.
Sliwa, R. and Draper, J. 2005. Solid geology of the Northern Bowen Basin, an interpretation based on
airborne geophysics. GIS product. http://www.glassearth.com/downloads/nbbmap.pdf
Soesoo, A., Bons, P.D., Gray, D.R. and Foster, D.A. 1997. Divergent double subduction; tectonic and
petrologic consequences. Geology 25, 755-758.
Spaggiari, C. V., Gray, D. R., Foster, D. A. and McKnight, S. 2003. Evolution of the boundary between
the western and central Lachlan Orogen: implications for Tasmanide tectonics. Australian Journal
of Earth Sciences 50, 725-749.
Spaggiari, C.V., Gray, D.R. and Foster, D.A. 2004. Ophiolite accretion in the Lachlan Orogen,
Southeastern Australia. Journal of Structural Geology 26, 87-112.
Squire, R.J., Wilson, C.J.L., Dugdale, L.J., Jupp, B.J. and Kaufman, A.L. 2006. Cambrian backarc-
basin basalt in western Victoria related to evolution of a continent-dipping subduction zone.
Australian Journal of Earth Sciences 53, 707-719.
Stephens, C.J., Holcombe, R.J. Fielding, C.R. and Gust, D.A. 1996. Tectonic evolution of the eastern
Gondwanaland margin during the late Palaeozoic and early Mesozoic. In: Mesozoic 96. Mesozoic
Geology of the Eastern Australia Plate Conference, Brisbane, 1996. Geological Society of
Australia, Extended Abstracts 43, 517-518.
Stephens, C.J., Schon, R.W.S. and Ewart, A. 1993. Mesozoic crustal extension in the northern NEO:
geochemical and isotopic evidence from large scale silicic magmatism. In: Flood, P.G. and
Aitchison, J.C. (eds). New England Orogen, Eastern Australia, NEO 93 conference proceedings,
pp 637-642. University of New England.
Stewart, I. 1995. Cambrian age for the Pipeclay Creek Formation, Tamworth Belt, northern New South
Wales. Cour. Forschungsinst. Seckenb. 182, 565-566.
Stokes, N., Fergusson, C.L. and Offler, R., 2015. Backarc basin and ocean island basalts in the
Narooma Accretionary Complex, Australia: setting, geochemistry and tectonics, Australian Journal
of Earth Sciences 62, 37-53, DOI: 10.1080/08120099.2015.985715.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 187


Stolz, A.J. 1994. Geochemistry and Nd isotope character of the Mt Windsor Volcanics; implications for
the tectonic setting of Cambro-Ordovician VHMS mineralization. Contributions of the Economic
Geology Research Unit, Vol.50, p.13-16; New developments in geology and metallogeny; northern
Tasman orogenic zone, Townsville, Queensl., Australia, Feb. 21-22 1994, edited by R.A.
Henderson and B.K. Davis. James Cook University of North Queensland.
Stolz, A.J. 1995. Geochemistry of the Mount Windsor Volcanics; implications for the tectonic setting of
Cambro-Ordovician volcanic-hosted massive sulfide mineralization in northeastern Australia.
Economic Geology 90, 1080-1097.
Stratford, J. M. C. and Aitchison, J. C. 1995. Lower Permian fauna from Manning facies rocks along
the Peel–Manning Fault System, Glenrock Station, New England orogen. Proceedings of the
Linnean Society of New South Wales 114, 239-246.
Stratford, J. M. C. and Aitchison, J. C. 1997. Lithostratigraphy of the Gamilaroi Terrane, upper Barnard
region, northeastern New South Wales. In: Ashley, P. M. and Flood, P. G. (eds.) Tectonics and
metallogenesis of the New England Orogen. Geological Society of Australia Special Publication 19,
178-187.
Stokes, N., Fergusson, C.L. and Offler, R., 2015. Backarc basin and ocean island basalts in the
Narooma Accretionary Complex, Australia: setting, geochemistry and tectonics, Australian Journal
of Earth Sciences 62, 37–53, DOI: 10.1080/08120099.2015.985715.
Sun, X. and Gravestock, D.I., 2001. Potential hydrocarbon reservoirs in upper levels of the eastern
Warburton Basin, South Australia. South Australia. Department of Primary Industries and
Resources Report Book, 2001/016.
Tadros, N.Z. 1995. Sydney-Gunnedah Basin Overview. In: Ward, C.R., Harrington, H.J., Mallet, C.W.
and Beeston, J.W. (eds) Geology of Australian Coal Basins. Geological Society of Australia Coal
Geology Group Special Publication, 1, 163-175.
Thalhammer, O.A.R., Stevens, B.P.J., Gibson, J.H. and Grum, W. 1998. Tibooburra Granodiorite,
western New South Wales: emplacement history and geochemistry. Australian Journal of Earth
Sciences 45, 775-787.
Totterdell, J.M., Moloney, J., Korsch, R.J. and Krassay, A.A., 2009. Sequence stratigraphy of the
Bowen–Gunnedah and Surat basins in New South Wales. Australian Journal of Earth Sciences 56,
433–59.
Turner, N.J., Black, L.P. and Kamperman, M. 1998. Dating of Neoproterozoic and Cambrian orogenies
in Tasmania. Australian Journal of Earth Sciences 45, 789-806.
Turner, S.P., Adams, C.J., Flottmann, T. and Foden, J.D., 1993. Geochemical and geochronological
constraints on the Glenelg River Complex, western Victoria. Australian Journal of Earth Sciences
40, 275-292.
Turner, S., Haines, P., Foster, D., Powell, R., Sandiford, M. and Offler, R., 2009. Did the Delamerian
Orogeny Start in the Neoproterozoic? Journal of Geology 117, 575–583. DOI: 10.1086/600866.
van Heeswijck, A 2004. The structure and hydrocarbon potential of the northern Drummond Basin and
northeastern Galilee Basin, central Queensland. In Boult, P.J., Johns, D.R and Lang, S.C. (eds)
Eastern Australian Basins Symposium II, Petroleum Exploration Society of Australia Special
Publication, pp. 319–30.
van Noord, K.A.A. 1999. Basin development, geological evolution and tectonic setting of the
Silverwood Group. In: Flood, P.G (editor). New England Orogen. Regional Geology, Tectonics and
Metallogenesis, pp 163-180. University of New England, Armidale.
Vandenberg, A.H.M. 2003. Silurian to Early Devonian – the Lachlan fold Belt at its most diverse. In
birch, W.D. (ed.) Geology of Victoria. Geological Society of Australia Special Publication 23, 117-
156.
VandenBerg, A.H.M., Willman, C.E., Maher, S., Simons, B.A., Cayley, R.A., Taylor, D.H., Morand,
V.J., Moore, D.H and Radojkovic, A. (eds.), 2000. The Tasman Fold Belt in Victoria. Geological
Survey of Victoria, Special Publication, Melbourne, 462 pp.
Veevers, J.J. (Ed.) 2000. Billion-Year Earth History of Australia and Neighbours in Gondwanaland.
GEMOC Press, Sydney, 400 pp.

188 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Veevers, J.J. 2004. Gondwanaland from 650–500 Ma assembly through 320 Ma merger in Pangea to
185–100 Ma breakup: supercontinental tectonics via stratigraphy and radiometric dating. Earth-
Science Reviews, 68, 1–132.
Veevers, J.J., 2013. Pangea: Geochronological correlation of successive environmental and strati-
tectonic phases in Europe and Australia. Earth-Science Reviews 127, 48–95.
von Gnielinski, F.E., Denaro, T.J., Wellman, P. and Pain, C.F., 1997. Torres Strait Region. In Bain,
J.H.C. and Draper, J.J (eds). North Queensland Geology. Bulletin of the Australian Geological
Survey Organisation 240, and Queensland Department of Mines and Energy Geology 9, 159-164.
Walter, M.R., Veevers, J.J., Calver, C.R. and Grey, K., 1995. Neoproterozoic stratigraphy of the
Centralian Superbasin, Australia. Precambrian Research 73, 173–195.
Walter, M.R., Veevers, J.J., Calver, C.R., Gorjan, P. and Hill, A.C., 2000. Dating the 840–544 Ma
Neoproterozoic interval by isotopes of strontium, carbon, and sulfur in seawater, and some
interpretative models. Precambrian Research, 100, 371–433.
Warren, A.Y.E, Gilligan, L.B. and Raphael, N.M. 1995. Geology of the Cootamundra 1:250 000 map
sheet. Geological Survey of New South Wales, Sydney, 160 pp.
Wass, R. and Dennis, D.M. 1977. Early Palaeozoic faunas from the Warwick-Stanthorpe region,
Queensland. Search 8, 207-208.
Watanabe, T., Leitch, E. C. and Fukui, S. 1993. Early Ordovician high P/T metamorphic inclusions
from serpentinite bodies in the southern New England Fold Belt. In: Flood, P. G. and Aitchison, J.
C. (eds.) New England Orogen, eastern Australia, pp. 181-186. University of New England,
Armidale.
Watanabe, T., Fanning, C. M. and Leitch, E. 1998. Neoproterozoic Attunga eclogite in the New
England Fold Belt. Australian Geological Society Convention, 14th (Townsville, Australia, 1998),
Abstracts 49, p. 458.
Watkins, J. 2007. New frontiers; minerals. Minfo 84, p.6-12. New South Wales Department of Primary
Industries, Sydney, N.S.W., Australia.
Watkins, J.J. and Meakin, N.S. 1996. Nyngan and Walgett 1:250 000 Geological Sheets SH/55-15 and
SH/55-11: Explanatory Notes, Geological Survey of New South Wales, Sydney, 112 pp.
Webb, A. W. and McDougall, I. 1968. The geochronology of the igneous rocks of eastern Queensland.
Journal of the Geological Society of Australia 15, 313-346.
Webby, B. D. 1978. History of the Ordovician continental platform shelf margin of Australia. Journal of
the Geological Society of Australia, 25, 41-63.
Willman, C.E., VandenBerg, A.H.M. and Morand, V.J., 2002. Evolution of the southeastern Lachlan
Fold Belt in Victoria. Australian Journal of Earth Sciences 49, 271–289.
Withnall, I.W. and Hutton, L.J., with contributions from Armit, R.J., Betts, P.G., Blewett, R.S.,
Champion, D.C. and Jell, P.A., 2013. Chapter 2. North Australian Craton. In P.A. Jell (ed.) Geology
of Queensland. Geological Survey of Queensland, pp. 23–112.
Withnall, I.W. and Lang, S.G. (eds) 1993. Geology of the Broken River Province, North Queensland.
Queensland Geology 4.
Withnall, I.W., Black, L.P. and Harvey, K.J. 1991. Geology and geochronology of the Balcooma area:
Part of an early Palaeozoic magmatic belt in North Queensland. Australian Journal of Earth
Sciences 38, 15-29.
Withnall, I.W., Blake, P.R., Crouch, S.B.S., Tennison Woods, K., Grimes, K.G., Hayward, M.A., Lam,
J.S., Garrad, P. and Rees, I.D. 1995. Geology of the southern part of the Anakie Inlier, central
Queensland, Queensland Geology, 7.
Withnall, I. W., Golding, S.D., Rees, I.D. and Dobos, S.K. 1996. K-Ar dating of the Anakie
Metamorphic Group; evidence for an extension of the Delamerian Orogeny into central
Queensland. Australian Journal of Earth Sciences 43, 567-572.
Withnall, I.W., Mackenzie, D.E., Denaro, T.J., Bain, J.H.C., Oversby, B.S., Knutson, J., Donchak,
P.J.T., Champion, D.C., Wellman, P., Cruikshank, B.I., Sun, S.S. and Pain, C.F. 1997a.
Georgetown Region. In: Bain, J.H.C. and Draper, J.J (eds). North Queensland Geology. Bulletin of
the Australian Geological Survey Organisation 240, and Queensland Department of Mines and
Energy Geology 9, 19-69.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 189


Withnall, I.W., Lang, S.C., Lam, J.S., Draper, J.J., Knutson, J., Grimes, K.G and Wellman, P. 1997b.
Clarke River Region. In: Bain, J.H.C. and Draper, J.J (eds). North Queensland Geology. Bulletin of
the Australian Geological Survey Organisation 240, and Queensland Department of Mines and
Energy Geology 9, 327-347.
Withnall, I.W, Hutton, L.J, Rienks, I.P, Bultitude, R.J, von Gnielinski, F.E, Lam, J.S, Garrad, P.D and
John, B.H. 1998. Queensland geological record 1998/1; South Connors-Auburn-Gogango Project;
progress report on investigations during 1997. Queensland Government Mining Journal 99, 46-47.
Wormald, R.J., Price, R.C. and Kemp, A.I.S. 2004. Geochemistry and Rb-Sr geochronology of the
alkaline-peralkaline Narraburra Complex, central southern New South Wales; tectonic significance
of Late Devonian granitic magmatism in the Lachlan Fold Belt. Australian Journal of Earth Sciences
51, 369-384.
Wyborn, D., 1992. The tectonic significance of Ordovician magmatism in the eastern Lachlan Fold
Belt. Tectonophysics, 214, 177-192
Wyborn, L. 1997. Australian mineral systems; a National Geoscience Mapping Accord Project.
Australian Geological Survey Organisation Record 1997/56.
Yarrol Project Team 1997. New insights into the geology of the northern New England Orogen in the
Rockhampton -Monto region, central coastal Queensland: progress report on the Yarrol Project.
Queensland Government Mining Journal 98, 11-26.
Yarrol Project Team 2003. Central Queensland Regional Geoscience Data-Yarrol GIS. Queensland
Department of Natural Resources and Mines, Brisbane.
Zucchetto, RG, Henderson, RA, Davis, BK. and Wysoczanski, R., 1999. Age constraints on
deformation of the eastern Hodgkinson Province, north Queensland: new perspectives on the
evolution of the northern Tasman Orogenic Zone’, Australian Journal of Earth Sciences 46, 105–
14.

190 Geodynamic Synthesis of the Phanerozoic of eastern Australia


Acknowledgements

The report was greatly assisted by detailed discussions on state geology and metallogeny, as well as
reviews by, personnel of the Queensland, New South Wales, Victoria and Tasmanian Geological
Surveys. We particularly wish to thank C. Murray, J. Draper, L. Hutton, I. Withnall (Geological Survey
of Queensland), R. Glen, and P. Blevin (Geological Survey of New South Wales), R, Cayley, C.
Willman (Geoscience Victoria), R. Bottrill, C. Calver, G. Green, M. McClenaghan, D. Seymour and J.
Taheri (Mineral Resources Tasmania). We also thank W. Collins (James Cook University) and R.
Korsch (Geoscience Australia), who along with R. Glen (Geological Survey of New South Wales) gave
invited presentations at Geoscience Australia on their views of Tasmanides Geology. We also
acknowledge the 2008 Tasmanides Workshop, Melbourne and the 2013 Thomson Workshop,
Brisbane, and the speakers involved who summarized various aspects of Tasmanides Geology. We
also acknowledge past and present colleagues at Geoscience Australia, including George Gibson,
Paul Henson, Andrew Cross, Simon Bodorkos, Russell Korsch, Michael Doublier and Geoff Fraser.
We also thank Russell Korsch, Simon Van Der Wielen and Subhash Jaireth who provided detailed
internal reviews at Geoscience Australia for the first edition and Natalie Kositcin and Andrew Cross for
the second edition.

Geodynamic Synthesis of the Phanerozoic of eastern Australia 191

You might also like