Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Volcanology and Geothermal Research 377 (2019) 103–116

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

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

Nature and origin of tuff beds in Jurassic strata of the Surat Basin,
Australia: Implications on the evolution of the eastern margin of
Gondwana during the Mesozoic
Carmine C. Wainman a,⁎, Peter Reynolds a, Tony Hall b, Peter J. McCabe a, Simon P. Holford a
a
Australian School of Petroleum, University of Adelaide, SA 5005, Australia
b
Department of Earth Sciences, School of Physical Sciences, University of Adelaide, Adelaide, SA 5005, Australia

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

Article history: Volcanogenic rocks in the Great Australian Superbasin provide one of the principal records of contemporaneous
Received 20 November 2018 volcanism in eastern Australia during the Mesozoic. However, the paucity of primary Jurassic to Early Cretaceous-
Received in revised form 15 March 2019 age extrusive or intrusive igneous bodies on the Australian continent makes it particularly challenging to deduce
Accepted 18 March 2019
their source and character. This, in turn, makes it difficult to ascertain how the eastern margin of Gondwana
Available online 22 March 2019
evolved during this timeframe. Despite some studies of this enigmatic volcanism, there have been little or no de-
Keywords:
tailed analyses of these Jurassic to Early Cretaceous-aged sediments. Based on the multidisciplinary analyses of
Air-fall volcanic ash age-constrained air-fall tuffs (168 to 148 Ma) from the Jurassic Walloon Coal Measures of the Surat Basin, we sug-
Volcanogenic rocks gest from zircon grains of a similar age that the tuffs were erupted from volcanoes fed by intermediate to felsic
Jurassic magmas supported by their quartz and feldspars-rich composition, and from zircon grains with low to moderate
Surat Basin Nb values (0.5 to 100 ppm) and high U and Th values (30 to 1000 ppm). The mapping of tuff isopachs and a mean
Subduction zircon crystal size of 170 μm supports the source being from volcanoes approximately 280 to 1000 km from the
Whitsunday Igneous Association palaeoeast-southeast with a volcanic explosivity index (VEI) of 8. Our results indicate the tuffs originated from a
continental arc setting associated with the Whitsunday Igneous Association, and the long-lived (late Palaeozoic
to Cretaceous) westward subduction of the palaeo-Pacific oceanic crust beneath eastern Australia. Such a
tectono-magmatic environment would help constrain the timing of the transition of eastern Gondwana from a
convergent to a divergent margin.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction subsurface (Huff et al., 1992; Schutter, 2003; Dai et al., 2017; Watson
et al., 2017).
Analysis of ancient volcanic successions can provide vital informa- Volcanogenic rocks are important components of the Late Triassic to
tion on basin evolution, climate change and tectonic regimes in the geo- Early Cretaceous infill of the Great Australian Superbasin which encom-
logic past (Huff et al., 1992; Orton, 1996; Königer et al., 2002; Tucker et passes the Surat and Eromanga Basins. They originated mainly as air-fall
al., 2016). Pyroclastic and associated sediments can provide an abun- volcanic ash. However, the location of the contemporaneous volcanic
dant additional source of material to that derived by weathering and terrain remains poorly constrained owing to the lack of evidence for vol-
erosion (Orton, 1996). The largest super-eruptions (Volcanic canic centres in eastern Australia during the Jurassic and the Early Cre-
Explosivity Index, VEI ≥ 8, with volumes N500 km3) may deposit tephra taceous (Jell, 2013; Tucker et al., 2016). It is suspected that the relatively
beds of up to 4 cm thick over distances N2000 km from the source area unexplored submerged continental crust located to the east of Australia
(Mason et al., 2004; Matthews et al., 2012). Such tephra deposits pro- may hold significant records of Mesozoic volcanism (Tucker et al.,
vide insights on ancient column heights, ejecta volumes, mass eruption 2016). However, since there is uncertainty as to when the eastern Aus-
rates, and the environmental impacts of ancient explosive eruptions tralian margin transitioned from a convergent to a divergent margin
(Carey and Sparks, 1986; Pyle, 1989; Huff et al., 1992). The deposits (Bache et al., 2014), the tectonic setting in which the magmas were gen-
are also invaluable tools in event stratigraphy and palaeogeographic re- erated is unclear (Fielding, 1996; Jell, 2013).
constructions, and can even influence the flow of fluids in the There are two schools of thought on the nature and character of the
contemporaneous volcanic terrain that was responsible for the wide-
spread deposition of volcanogenic rocks in the region. Veevers et al.
⁎ Corresponding author. (1991), Veevers (2006), Bache et al. (2014), Tucker et al. (2016) and
E-mail address: carmine.wainman@adelaide.edu.au (C.C. Wainman). Babaahmadi et al. (2018), for instance, argue for the existence of a

https://doi.org/10.1016/j.jvolgeores.2019.03.012
0377-0273/© 2019 Elsevier B.V. All rights reserved.
104 C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116

Fig. 1. Summary of key tectonic events that occurred in eastern Australia between the Late Triassic (Rhaetian) and the Early Cretaceous (Albian) (Bryan et al., 2012; Jell, 2013; Tucker et al.,
2016).

long-lived continental arc on the eastern margin of Australia from the wedge and forearc sediments have also been used as evidence for the
Late Carboniferous (330 Ma) until the mid-Cretaceous (95 Ma). This is long-lived subduction (Cambrian to Late Cretaceous) on the southeast-
based largely on: detrital zircon U–Pb dating of sandstones (92 Ma) in ern oceanic margin of Gondwana (Mortimer et al., 2012).
the Mesozoic Australian Superbasin; the dating of clay gouges of the In contrast, Bryan (1996), Bryan et al. (1997), Bryan et al. (2000),
Demon Fault (163–152 Ma) suggesting intraplate tectonism related to Bryan et al. (2012) and Jell (2013) suggest volcanic rocks within the
subduction in eastern Gondwana, the interpretation of Zealandia-3 Great Australian Superbasin were sourced from a large igneous prov-
Megasequence (Jurassic to Early Cretaceous) on seismic profiles be- ince (The Whitsunday Igneous Association (WIA), also known as the
tween Australia, New Caledonia and New Zealand that comprise sub- Whitsunday Large Igneous Province or WSLIP) that formed during the
duction-related units and volcanic/volcaniclastic rocks; and reviewing onset of rifting from the Late Jurassic (162 Ma) and reached its peak in
patterns of seafloor spreading around Australia based on magnetic- the Early Cretaceous between 120 and 105 Ma. Based on the geochem-
anomaly isochrons. On the South Island of New Zealand, exhumed istry of basalts (E-MORB (enriched-mid ocean ridge basalts) affinities)
low-grade metamorphic rocks that were originally accretionary near Proserpine in Queensland, and the absence of any large
C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116 105

N1000 km from the eastern Australian margin. Bryan et al. (2012) and
Jell (2013) suggest that rifting commenced in the Middle Jurassic
(Callovian), but this is poorly constrained from analysis of tuffs in the
Walloon Coal Measures in the Surat Basin and contemporaneous alkali
basalts (162 ± 0.9 Ma) that were intersected on the Marion Plateau cor-
ing undertaken by the International Ocean Drilling Program (IODP)
Cruise 194 (Isern et al., 2002). They are thought to define the eastern
limit of the Whitsunday Igneous Association, compositionally similar
to other alkali basalts from an intraplate margin, and represent a back-
arc extensional event prior to the emplacement of the silicic large igne-
ous province (Bryan et al., 2012). Major tectonic events that are thought
to have occurred in eastern Australian from the Late Triassic (Rhaetian)
to the Early Cretaceous (Albian) are summarised in Fig. 1.
To throw light on the nature and character of the volcanic source,
this study analyses tuffs from the Jurassic Walloon Coal Measures of
the Injune Creek Group of the Surat Basin in southeast Queensland
(Fig. 2), a sub-basin of the regionally extensive Great Australian
Superbasin.

2. Geologic setting of the Surat Basin

The Late Triassic (Rhaetian) to Early Cretaceous (Albian) Surat Basin


(Fig. 2) is the third largest sub-basin within the Great Australian
Superbasin (Jell, 2013). Overlying the Permo-Triassic Bowen Basin
(Fig. 2), the Surat Basin covers approximately 270,000 km2 of southeast
Queensland and northern New South Wales (Jell, 2013). It connects
with the Eromanga Basin to the west and the Clarence-Moreton Basin
to the east (Exon, 1976). The basin is filled with between 1800 and
2700 m of nonmarine and marine strata that were deposited during
six sedimentary cycles in response to changes in eustatic sea level
(Exon and Burger, 1981; Goscombe and Coxhead, 1995). During the Ju-
rassic and Cretaceous, the Surat Basin was located in the polar latitudes
in the southern hemisphere, on the eastern margin of Gondwana
(Klootwijk, 2009).
One of the most enigmatic aspects of the geological evolution of the
basin concerns its subsidence history and the relatively unknown tec-
tonic character of the eastern Australian margin during the Jurassic
(Waschbusch et al., 2009; Jell, 2013). Incipient rifting (Fielding, 1996),
thermal sag after the cessation of volcanic arc activity (Korsch et al.,
1989), intracratonic sag resulting from extensional plate motion
(Green et al., 1997), subsidence related to sub lithospheric convection
as a result of subduction to the east (Gallagher et al., 1994), and dynam-
ically induced platform tilting (Korsch and Totterdell, 2009) have all
been suggested as mechanisms for basin subsidence.
The Callovian to Tithonian-age Walloon Coal Measures (sometimes
called the Walloon Subgroup) (Fig. 2) are regionally extensive and up
to 500 m thick (Jell, 2013). The formation consists of sandstones, silt-
stones, mudstones, coals, carbonaceous shales and tuffs deposited in
fluvial, lacustrine and estuarine environments (Martin et al., 2013;
Wainman and McCabe, 2018). In each core cut through the Walloon
Coal Measures, there are up to 10 tuff beds which are up to 2.13 m in
thickness. The location of tuff beds analysed is shown on a well cross-
section (Fig. 3). Previous work suggests the tuffs were derived from
extrabasinal volcanic centres towards the east, however, the petrogen-
esis of the parent magmas has not previously been constrained as they
Fig. 2. Map (a) showing the location of the Surat Basin in Queensland Australia, major
have only been described very briefly in field outcrops, or inferred
geological structures, and the wells used in this study, (b) a simplified structural cross-
section through the Surat Basin along line B'-B (adapted from Jell (2013)), and (c) The from petrographic analysis of volcanogenic sandstones (Boult, 1996;
stratigraphic column shows the age and continuity of the Walloon Coal Measures within Yago, 1996; Yago and Fielding, 2015). Wainman et al. (2015) and
the basin. (Adapted from Wainman et al. (2018)). Wainman et al. (2018) sampled and dated volcanic zircon from 27 tuff
beds in core across 10 wells that intersected the Eurombah Formation,
the Walloon Coal Measures and the Springbok Sandstone in the Surat
stratovolcanoes offshore near the Whitsunday Islands in Queensland, Basin, Queensland using the high-precision uranium–lead chemical
these authors suggest the Whitsunday Igneous Association was related abrasion-thermal ionization mass spectrometry (U–Pb CA-TIMS) tech-
to the breakup of Gondwana (and subsequently, the formation of the nique in the Surat Basin (Fig. 3). This study revealed that volcanism
modern eastern Australian margin). Bryan et al. (1997) suggest that was prevalent from the Middle Jurassic (Bathonian) through to the
even if a coeval subduction zone was present, it would have been Late Jurassic (Tithonian).
106 C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116

Fig. 3. Well cross-section and wireline logs across the Walloon Coal Measures in the Surat Basin of Queensland showing the stratigraphic position of tuff beds and types of analyses
undertaken in this study. The cause of the sequence boundary remains enigmatic; however, it suggested that this maybe be a result of topographic changes due to slab breakoff during
the process of subduction (Smith et al., 2019).

3. Macro-scale description of tuffs

From cores derived from 10 wells (Fig. 3 with the exception of the
Cameron 1 well), tuff beds were visually described with notes taken
on colour, bed thickness, fabric and stratigraphic contacts.
Measured tuff beds identified within the Walloon Coal Measures
vary from b0.01 to 2.13 m thick, with a median thickness of 0.04 m.
They vary in colour from beige to grey, massive in character and gener-
ally have sharp planar contacts with underlying beds (Fig. 4). In some
cases, the base of the tuff beds appear to be bioturbated; these are
characterised by irregular shaped, highly sinuous tubular structures
that are 1 to 5 mm wide and up to 3 cm in length. The tops of the tuffs
beds have sharp or gradational contacts with overlying beds over sev-
eral decimetres, and exhibit no signs of having been reworked with no
cross-lamination or erosional scours. Due to the intense alteration of
the tuffs and their high clay content, it was not possible to determine
their median grain size at the macroscale. The majority of the tuff
beds were not sufficiently thick (b0.04 m) to produce a characteristic
petrophysical response in wireline logs to aide their recognition –
these would typically give high gamma values, a distinct low- to moder-
ate-density and high neutron response (Watson et al., 2017).
The fact that the tuffs are found interbedded with coals and lacus-
trine rocks indicate that they were deposited in-situ. Once deposited,
they were not subject to currents that may have caused reworking.
Mire environments, in particular, tend to be isolated from active chan-
nel bodies that introduce siliclastics and terminate the accumulation
of peat (McCabe, 1984). Equally, the vegetation within mires tends to
baffle currents that may lead to the reworking of inorganics including
volcanic ash (Triplehorn and Bohor, 1986). In addition, any air-fall ash
tends to have a good chance of remaining undisturbed from deposition
through to diagenesis in mire and lacustrine environments (Triplehorn
and Bohor, 1986; Dai et al., 2017).

4. Petrographic description of tuffs and XRD analyses

Four tuffs (Stratheden 4 well depth, 239.46–239.39 m; Stratheden


60 well depth, 171.50–171.34 m; Alderley 1 well depth, 268.16–
286.09 m; and Kato 3 well depth, 105.23–104.64 m) at different strati-
graphic intervals in the Walloon Coal Measures across the Surat Basin
in Queensland were chosen for petrographic analysis under the micro-
Fig. 4. Photograph of two tuff beds from the Kato 3 well (from depths 105.68 to 103.68 m). scope (Fig. 3). These samples were chosen for being located between
C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116 107

Fig. 5. Thin-section images of a tuff in the Kato 3 well, taken from 105.23 to 104.64 m depth. Images A and C are shown in plane-polarised light; image B and D are shown in cross-polarised
light.

coal beds and thus unlikely to have been subject to currents that may against fluvial transport and reworking of the tuffs. The angular grain
have caused reworking. Micro-scale petrographic descriptions were un- morphologies are not interpreted to result from in-situ fragmentation
dertaken using rock thin-sections analysed under a Nikon petrographic since none of the clasts display a jigsaw-fit fragmentation (Best and
microscope in both plain polarised (PPL) and cross polarised (XPL) light. Christiansen, 1997). The mineralogy of the tuffs (dominated by quartz
Notes were taken on crystal size, sorting, composition, texture and alter- and feldspars that have altered to clays) suggests that the parent
ation for classification purposes. magmas were silicic in composition. These tuffs are similar in composi-
Petrographic analyses from these four samples reveal the tuffs are tion and texture to those found in other coal-bearing sequences includ-
poorly sorted, lightly compacted, lack any structure and show no evi- ing the Pennsylvanian Fire Clay Tonstein in the Appalachian Basin of the
dence of grading. They are comprised predominately of splintery, angu- USA that was soured from the Hercynian magmatic arc located in east-
lar quartz clasts (approx. 10–100 μm in diameter) supported in an ern North Carolina (Chesnut, 1983; Greb et al., 1999), and in Lopingian
amorphous, white-buff coloured matrix consisting of clay minerals coals across China that were sourced from felsic and rhyolitic magmas
(Fig. 5). Some of the quartz clasts have point contacts and are fractured. (Dai et al., 2017). Given the large amounts of alteration within the sam-
The clasts are massive, are not aligned and do not display any structure pled tuffs, we are unable to determine whether they represent a crystal,
or grading. vitric or lithic tuff (Schmid, 1981).
To support the macro-scale and petrographic observations, the min-
eralogy of 16 tuff samples (Figs. 3 and 6.) from different stratigraphic in- 5. Zircon crystal sizes and character
tervals in the Walloon Coal Measures across the Surat Basin in
Queensland were determined by X-ray diffraction (XRD) analysis con- In order to determine the grain size of zircon crystals and conse-
ducted qualitatively using a Bruker D8 ADVANCE Powder X-ray Diffrac- quently infer the transport distances of the tuffs, we measured 260
tometer with a Cu-radiation source. Data was processed using Bruker cathodoluminescence (CL) images of zircon crystals from 10 samples
DIFFRAC, EVA software and Crystallography Open Database reference available in Wainman et al. (2018) using the image analysis programme
patterns for identifying mineral phases. Quantification was calculated ImageJ (http://imagej.nih.gov/ij/) (Fig. 7). Only dated grains of a volca-
against an internal standard of zinc oxide at 10% using the USGS (United nic origin with ages between 149.08 and 168.07 Ma determined by ei-
States Geological Survey) RockJock software. ther CA-TIMS or laser ablation inductively coupled plasma mass
XRD analyses from these 16 samples confirm that the tuffs are dom- spectrometry (LA-ICPMS) methodologies were measured, thereby pro-
inantly comprised of quartz (≤40%) and clay (≥62%). Minor constituents viding us with an age-constrained suite of data for the measured sam-
include albite (≤16%), orthoclase (≤7%) and calcite (≤3%) (see Table 1; ples. Zircon textures were also analysed using the same
Fig. 6). Trace constituents (b0.05%) unobtainable by XRD analyses but cathodoluminescence (CL) images.
observed in thin-section include sericite, muscovite mica, biotite mica, CL images reveal that the zircon crystals have concentric zoning. Zir-
zircon and rutile. con cores vary from zoned to homogeneous with resorption patches.
The splintery, angular shape of the clasts supports our interpretation Additionally, some of the crystals have embayed rims. Grain shape anal-
that the clasts are a pyroclastcic fall deposit; such clast morphologies are ysis reveals that they have circularities (a function of the perimeter and
typical of fall deposits (Rosi et al., 1999). The shapes of the clasts argue area of any given grain) varying from 0.2 to 0.8 and a roundness value
108 C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116

Fig. 6. X-Ray diffraction analysis of three tuffs from the Stratheden 60, Guluguba 2 and Kato 3 wells from the Walloon Coal Measures in the Surat Basin. Colour bars denote key peaks from
each mineral and correspond to highlighted peaks in the sample (see legend for each tuff). Listed order or height of the peak does not represent differences in the quantitative abundance of
minerals present in each of the tuffs.
C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116 109

(degree of smoothness on the perimeter of the grain) of 0.2–0.8. They

105.23 m
have aspect ratios of 2:1. The grains vary from euhedral to tabular in

100.0
Kato

16.0

26.9

73.1

73.1
8.1

2.8
shape. There is a trend towards crystals becoming more elongate as


they become larger (Fig. 7).

360.59 m
The size of the zircon crystals varies from 52 to 527 μm, whilst the

22.07

100.0
mean crystal size for all measured crystals is 170 μm. Measurements
Indy

40.0
12.4

59.7

18.3

40.3
7.2
3


of zircons from each of the wells are shown in Table 2. The largest
mean crystal size was found in the Turallin 1 well, located in the east
Guluguba

252.51 m
of the basin. The smallest crystals sizes were located in the Guluguba 2

100.0
21.3

27.4

12.9
43.8
15.9
72.6
6.1
well in the northern part of the basin. There is no systematic variation
2



in grain size across the basin or with depth through the wells.
Guluguba

268.53 m

These zircon crystals are thought to represent the juvenile compo-


nent of a pyroclastic fall deposit. The euhedral shape and aspect ratios

100.0
30.7

39.3

14.9
11.6
34.1
60.7
8.6

of the crystals support this origin and their angular shape is consistent
2


with deposition from an ash cloud (Lyons et al., 1992; Königer et al.,
Guluguba

269.07 m

2002; Dai et al., 2017). The zircons are interpreted to have crystallised

100.0
in one, or more likely, several parent magma chambers prior to their
29.3

37.6

17.0
13.7
31.8
62.4
8.2
2


eruption; the absence of anhedral cores suggests that they are not
inherited (Paterson et al., 1992; Hanchar and Miller, 1993; Dai et al.,
Guluguba

278.05 m

2017). Further evidence supporting the zircons representing a juvenile


100.0
30.0

38.1

14.3
17.8
29.7
61.9
8.1

component of a pyroclastic fall deposit comes from the high probability


2


of fit (N0.05) for individual tuff dates which adhere to the law of super-
Percentages of minerals in tuffs determined by XRD analysis. Quantification was calculated against an internal standard of zincite at 10% using the Rockjock software.

Guluguba

414.45 m

position up-strata in any given well (Fig. 3) (Wainman et al., 2018). This
100.0

strongly suggests a zircon crystallisation age that is likely to reflect the


35.8
55.0

95.8
1.6
2.6

4.2

5.0

age of the eruption event (Bowring and Schmitz, 2003).


2



Guluguba

527.39 m

36.26

100.0
24.0

28.5

11.0
24.3

71.5

6. Zircon geochemistry
4.5
2


Using an existing LA-ICPMS dataset from Wainman et al. (2018)


157.10 m
Alderley

from six samples (Figs. 3 and 8), trace element abundances in parts
100.0
71.4
25.7
0.05
97.1
2.9

2.9

per million (ppm) from dated zircon of ages between 149.08 and
1



168.07 Ma were plotted against each other using methods of


268.09 m
Alderley

Belousova et al. (2002) and Grimes et al. (2015) to indirectly determine


100.0

source rock type and tectono-magmatic provenance of the tuffs.


97.3

97.3
2.7

2.7
1




Based on the five trace element discrimination plots employed by


Belousova et al. (2002), we infer that the zircons from six samples
Stratheden

171.54 m

found in the Surat Basin were derived from granodiorite and possibly
100.0

syenite magmas, notably from the Alderley 1 well (GA 2254159 and
10.0

80.8

90.0
2.8
7.1

9.3
60


GA 2254160). One sample from the Stratheden 60 well (GA 2231589)


was maybe derived from lamproites. This is supported by moderate U
Stratheden

332.84 m

values (100 to 1000 pm) and elevated Y values (N500 pm; Fig. 8). Ta
100.0

Vs Nb discrimination plots, on the contrary, suggest a more mafic to


97.8

98.6
1.4

1.4

0.8
60



lamproitic source (Ta b1 ppm and Nb b 10 ppm). Based on seven trace


element discrimination plots by Grimes et al. (2015), we infer from
Stratheden

197.10 m

four of plots that the zircons are similar to those derived from pres-
100.0

ent-day continental arc settings such as Mt. St Helens in Washington


10.4
83.7

94.1
0.8
5.1

5.9
4


State in the USA and Taupo in New Zealand. This is supported by low
to moderate Nb values (0.5 to 100 ppm) and high U and Th values (30
239.46–39 m
Stratheden

to 1000 ppm). For three of the plots (Nb-Yb, Hf-U/Yb and Nb-Ti), the
data plotted outside of known geochemical limits associated with pres-
100.0
92.9

92.9
2.9

4.2

7.1

ent-day tectono-magmatic environments (Grimes et al., 2015). This


4


may be associated with unusual fractionation processes in the magma


Stratheden

chamber, or random variation between isolated melt pockets formed


371.43 m

late in the crystallisation sequence (Grimes et al., 2015). The inclusion


100.0
52.6
45.3

97.9
2.1

2.1

of such data may mask the primary characteristics of the parent


4



magma and clues relating to the tectonic environment.


Stratheden

Zircon crystals from four of the six samples in the Stratheden 60 and
396.69 m

Alderley 1 wells are thought to be derived from magmas on the inter-


100.0
10.6

14.3
75.1

89.4
4.2
6.4

mediate (granodiorites) to felsic (syenites) spectrum that were derived


4


from a continental arc setting, indicated by the analysis of similar zircon


Total non-clays

geochemistry results by Grimes et al. (2015). In sample GA 2180601


Kaolinite
Smectite

from the Stratheden 4 well, zircon may have been derived from a basal-
Orthoclase

Total clays

tic parent rock from a similar tectono-magmatic environment; this is in-


Quartz

Illite
Calcite

TOTAL
CLAYS
Albite
Table 1

dicated by high and variable Th/U (1 to 3) and variable Eu anomalies


ratios (0.1 to 0.9; Fig. 8) (Wainman et al., 2015).
110 C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116

Fig. 7. Graph showing the relationship between zircon crystal size and aspect ratio. The inset shows a CL image of a typical crystal; note the concentric zoning.

7. Tuff isopachs, prevailing palaeowind direction and transport eruption columns that reach the tropopause allow for increased fallout
distance in the distal regions 100's km from the volcano(es) (Brown et al.,
2012); and the influence of palaeotopography and/or the character of
The thickness of individual tuff beds (in metres) were extrapolated the ancient peat accumulation environment (Dai et al., 2017). Despite
from 68 well completion reports from wells that intersect the Walloon these inaccuracies, the isopachs strongly suggest the prevailing
Coal Measures of the Surat Basin. These were tallied for each well and palaeowind direction was from east-southeast to west-northwest.
plotted on a map irrespective of their stratigraphic position in the Volcanic ash deposits tend to decrease in grain size with distance
well, but most lie between the Middle Jurassic (Bathonian) and the from the vent (Fisher and Schmincke, 1984), although non-uniform fall-
Upper Jurassic (Tithonian) (Fig. 9). Using the total tuff thickness from out occurs when clasts within the ash cloud have varying hydraulic
each well, isopachs of equal thickness were contoured between the properties. For instance, minerals with high specific gravity such as zir-
data points. It should be noted that these are generalised tuff isopachs con are commonly transported shorter distances than minerals such as
because some wells do not precise intersect age-equivalent strata biotite, which have a higher “suspension ability” within ash clouds
from east to west, and not all the wells penetrate the entire succession. (Königer et al., 2002). Additionally, the maximum transport distances
Base level changes through time and the transient nature of erosion and for fine-grained (b1 mm size) particles are difficult to estimate, since
deposition in fluvial systems also means there are frequent sedimentary convective and thermal processes operate within the eruption column
hiati in any given stratigraphic succession. We use the methodology of and umbrella cloud (Königer et al., 2002). Nevertheless, the size of the
Königer et al. (2002) to assume that the lateral distribution of volcanic zircon crystals can be used to provide some estimate of the transport
air-fall ash (tuff) is controlled by the prevailing palaeowind direction distance of the tuffs (Barham et al., 2016).
and that ash/tuff beds thin away from the volcanic source. Total tuff In conjunction with other data acquired in this study, we use con-
thickness is variable across the basin; from b0.1 m in the western temporary eruption events to infer the conditions that may have led
Surat Basin to N5 m in the central Surat Basin. The Isopach mapping of to the deposition of the tuffs preserved in the Jurassic Walloon Coal
tuffs suggests that there are elongate lobes that thin from current day Measures including VEI and transport distance. The zircon crystals we
northeast to southwest (Fig. 9). When rotated to Australia's orientation describe have a mean size of 170 μm (Fig. 7 and Table 2). These crystals
in the Jurassic (~70° clockwise) (Torsvik and Cocks, 2016), this suggests are likely to be smaller than the average size of the ash shards (500 μm)
that the prevailing palaeowind direction was from east-southeast to recorded from the eruption of Mount St. Helens in 1980, due to the
west-northwest. This ties in with atmospheric circulation models from lower suspension ability of the zircon grains (Sarna-Wojcicki et al.,
this timeframe (Moore et al., 1992; Valdes et al., 1995; Sellwood and 1981). In deposits from the same eruption (VEI 5), clasts 320 and 110
Valdes, 1997). In the modern day high latitudes, prevailing winds μm in size are reported 220 and 310 km downwind from the vent
blow from high-pressure zones of the polar high, towards low-pressure (Sarna-Wojcicki et al., 1981). In comparison, the 74 ka Toba event
areas of the westerlies at the polar front in a westerly direction (the (VEI 8) deposited clasts with a mean size of 64 μm 2669 km from the
polar easterlies) (Golonka et al., 1994). A similar situation would have vent. These observations suggest that the tuffs we describe could have
occurred during the Middle to Late Jurassic (Golonka et al., 1994). This been produced from a vent N1000 km away from their site of deposi-
suggests that the source of volcanism was off the eastern margin of tion, provided that the eruption was of significantly high magnitude
Gondwana (Torsvik and Cocks, 2016). Alternative hypotheses for (VEI ≥8). Alternatively, the crystals could have been produced from a
these tuff isopach patterns include: the premature fallout due to the ac- much smaller magnitude eruption (VEI 5) at ~280 km distance. The
cretion of dominant small glass shards in the ash cloud (Lerbekmo, transport distances implied by the size of the zircon crystals and ash
2002); distal mass accumulation (or ‘secondary thickening’) because shards are likely to be minimum estimates.
C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116 111

Table 2
Crystal sizes (minimum, maximum, average and 2σ standard deviation) and aspect ratios (minimum, maximum, average and 2σ standard deviation) from 260 zircon grains across in ten
wells in the Jurassic Walloon Coal Measures across the Surat Basin, Queensland determined using the image analysis programme ImageJ.

Well Min size Max size Mean size Standard deviation Min aspect ratio Max aspect ratio Mean Standard deviation Number
(μm) (μm) (μm) (2σ) (AR) (AR) AR (2σ) (n)

Turallin 1 202.00 454.80 284.62 88.97 1.70 4.90 2.87 1.13 6


Stratheden 4 189.50 258.20 229.81 27.07 1.60 3.00 2.16 0.48 7
Stratheden 60 52.70 527.30 148.36 72.43 1.00 4.80 1.99 0.63 95
Wyalla 3 162.10 293.30 230.55 40.25 1.30 3.90 2.49 0.88 10
Alderley 1 82.60 376.20 179.58 62.58 1.20 5.90 2.38 1.03 93
Guluguba 2 141.70 337.50 141.70 69.26 1.90 4.00 2.68 0.86 9
Cameron 1 129.00 217.20 169.53 33.78 1.60 2.80 2.18 0.48 6
Pleasant Hills 83.30 184.80 134.64 28.35 1.20 3.10 2.06 0.56 18
25
Indy 4 96.50 297.60 140.38 77.29 1.50 5.90 2.62 1.63 6
Indy 3 102.80 202.00 146.01 35.18 1.70 4.00 2.60 0.84 10
Total number of zircon analysed 260

Similar to the grain size, the thickness of air-fall deposits tends to de- grain size, such as those produced during the Mt. St Helens eruption e.g.
crease with distance from the source volcano (Fisher and Schmincke, Sarna-Wojcicki et al. (1981).
1984). The mean thickness of the tuff beds in the wells from the Surat In summary, based on the bed thickness, tuff isopachs, clast size and
Basin is 7 cm. We note that this likely represents a minimum estimate the absence of any volcanic centre(s) within\ 200 km from our study
since we have not accounted for compaction, or variation through area, we suggest the tuffs most likely represent distal fall deposits up
time by taking into account their age. Therefore, calculated transport to 1000 km from the source, produced from large magnitude (~VEI 8)
distances are likely to be minimum estimates. Beds of comparable eruptions.
(and greater) thicknesses were deposited 500–1000 km from the vent
as the result of the 74 ka Toba and 340 ka Whakamaru eruptions (VEI
8). In comparison, ash fall from the Mt. St Helens eruption was 7 cm 8. Petrogenesis of the parent magmas
thick at 60 km distance. However, there is no evidence of coeval volca-
nic centre(s) within ~200 km of the wells in our study, and if the volca- Petrographic analyses of the tuffs, associated XRD data and the rela-
nic centre were located within 60 km the tuffs would likely have a larger tive abundance of zircon grains indicate their parent magma was pre-
dominately silicic (felsic). This is supported by observations of the
abundance of quartz and feldspars that have altered to clays, and the re-
covery of 100's of zircon grains per sample. The morphology of these
grains (sharply faceted, zoned to homogeneous and lacking inherited
cores) indicates these are eruption phase zircons. By using zircon geo-
chemistry data from volcanic grains in the tuff beds, first order determi-
nations on the source rock type and crystallisation environment could
be made enabling distinctions between the major tectonic settings
under which the parent magma was generated (Belousova et al.,
2002; Hawkesworth and Kemp, 2006; Grimes et al., 2015). This is par-
ticularly useful when the parent igneous body has not been located, as
is the case in this study. It is well established that the geochemistry of
zircon provides a sensitive monitor of its parental magma composition
despite variables (e.g. temperature and oxygen fugacity) that can affect
the trace element chemistry of zircon (Grimes et al., 2015). Generally,
rare earth elements in zircon increase in concentration from a mantle
origin through to pegmatites (i.e., element concentrations tend to be
higher in the most evolved, fractionated rock types) (Belousova et al.,
2002). Grimes et al. (2015) also demonstrate that there is a close asso-
ciation between zircon trace element and present-day tectono-mag-
matic environments. This may provide clues to interpreting past
tectono-magmatic environments in this region.
Although three of our analyses fell outside known geochemical
boundaries of studied tectono-magmatic settings from the modern
day (an unfortunate error often associated with analyses of zircon crys-
tals that can be related to unusual magmatic processes (Hawkesworth
and Kemp, 2006; Grimes et al., 2015)), four samples (2231589 and
2213590 from the Stratheden 60 well and 5554159 and 2254160 from
the Alderley 1 well) indicated that the zircons and their magmas were
derived from a continental arc that produced intermediate and felsic
magmas as suggested by Nb-U and U-Y geochemistry plots (Fig. 8).
Based on these analyses, we infer that the tuffs were erupted from
Fig. 8. Geochemical plots showing the relationship between U and Y (a) and Nb and U (b) magmas generated at an ancient subduction zone; such a setting is ca-
for the zircon crystals. The U-Y plot shows the tuffs were derived from granodiorites and
syenites (moderate U values (100 to 1000 pm) and elevated Y values (N500 pm)). The
pable of producing all the parent magma types we document
Nb-Y plots suggest they were derived from a continental arc setting (low to moderate (Tatsumi, 1989; Stern and Bloomer, 1992; Blakely et al., 2005; Rapp et
Nb values (0.5 to 100 ppm) and high U and Th values (30 to 1000 ppm)). al., 2008).
112 C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116

Fig. 9. Map showing the volcanic tuff isopachs from the Walloon Coal Measures across the Surat Basin in Queensland. Inferences on the east-southeast palaeowind direction are deduced
from decreasing total tuff thickness across the basin and rotating Australia ~70° clockwise.

9. Implications for the origin of Mesozoic magmatism in eastern bodies and volcanogenic sediments of a similar age are present along
Australia the southern margin of Australia (Fig. 10), but are mostly associated
with intraplate volcanism (Young, 1987), the break-up of Australia
Our data suggest that tuffs within the Jurassic Walloon Coal Mea- from Antarctica (Mitchell et al., 1997), or the Karoo-Ferrar Large Igne-
sures of the Surat Basin were sourced from a magmatic arc/subduction ous Province (Elliot and Fleming, 2017).
zone system located to the east-southeast of the current day Queens- It has been suggested that the Whitsunday Igneous Association was
land coast (Fig. 10) that was active during this timeframe. A candidate an expression of a continental arc that was active until the Turonian
source is the Whitsunday Igneous Association. The Siliceous Large Igne- (~92 Ma) from Lu-Hf isotope analysis of detrital zircon from mid-Creta-
ous Province or (SLIP) to which the Whitsunday Igneous Association be- ceous sandstones of the Winton and Mackunda Formations in the
longs to stretched to the south beyond the Marion Plateau along the Eromanga Basin (Tucker et al., 2016). The Lu-Hf isotope data suggests
eastern seaboard of Australia (Blevin and Cathro, 2008; Barham et al., that they are from a juvenile source rather than from the melting of sig-
2016). This was the nearest volcanic province to the Surat Basin during nificantly older crust (Tucker et al., 2016). Furthermore, Jurassic to Early
the Mesozoic and is located 480 km to the east of the basin margin (Fig. Cretaceous granodiorites, dolerites and low grade metamorphic rocks
10). Eruptions within this province of VEI magnitude 8 or less could from New Zealand, of which the protolith was an accretionary prism
have deposited beds of tuff with similar characteristics to those we de- complex (a time when New Zealand was still attached to Gondwana),
scribe in the Surat Basin (0.07 m average thickness from intermediate to suggest subduction continued until at least 100 Ma on the southeastern
felsic sources). Other authors have reported similarly high magnitude oceanic margin of Gondwana (Scott and Palin, 2008; Mortimer et al.,
eruptions (VEI 8) from the Whitsunday Igneous Association, although 2009; Mortimer et al., 2012; Mortimer et al., 2015). This notion is also
these eruptions are thought to have occurred during the Cretaceous supported by Barham et al. (2016) who have suggested the 106 Ma zir-
(Barham et al., 2016). Other candidate sources include the Marion Pla- cons from the Madura Shelf (onshore Bight Basin) represent ultra-distal
teau, the Lord Howe Rise, or the Teremba, Koh Central and Boghen ter- ashfall from the Whitsunday Igneous Association.
ranes of New Caledonia where Jurassic-age rocks of similar affinities These previous findings together with our data support the notion
(alkali basalts, I-type granites. andesites, dacites, rhyolites and that the eastern Australian margin was located atop a subduction zone
plagiogranites) have been identified (Cluzel et al., 2012; Jell, 2013; that endured into the mid-Cretaceous. This hypothesis contrasts with
Mortimer et al., 2015). Comparisons of equivalent datasets between the works of Ewart et al. (1992), Bryan et al. (1997), Bryan et al.
tuffs of the Walloon Coal Measures in the Surat Basin and igneous (2000) and Bryan et al. (2012) who suggested that the Whitsunday Ig-
rocks analysed from candidate sources are shown in Table 3. Despite neous Association and associated intrusions in south-eastern Queens-
the paucity of data from igneous rocks from these candidate sources, fu- land (including the Noosa Quartz Diorite) represented a large igneous
ture ocean drilling in these locations may help to resolve their igneous province that was primarily active during the Cretaceous around 100
activity during this timeframe. It should be noted that other igneous to 85 Ma, and located in an intraplate setting despite its silicic character
C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116 113

Fig. 10. Palaeogeographic reconstruction of the convergent eastern margin of Gondwana during the Jurassic showing the age and distribution of ancient volcanic centres, subduction
complexes, palaeowind direction and sites of deposition for volcaniclastic/volcanogenic sediments. Dates obtained from Jurassic-Cretaceous volcanic centres in South Australia and
Victoria (Sutherland, 1966; Sutherland, 1978; Twidale, 1994); the Garrawilla Lavas in New South Wales (Dulhunty, 1965); the Mount Bauple Syenites of the Maryborough Basin
(Green and Webb, 1974; Jones and Veevers, 1983), the Noosa Quartz Diorite (Green and Webb, 1974); Mesozoic igneous rocks in northern New South Wales (Dulhunty et al., 1987;
Middlemost et al., 1992); the Warnie Volcanic Province in the Eromanga Basin (Hardman et al., 2018); Jurassic dolerites in Tasmania (Hergt et al., 1989); the Whitsunday Igneous
Association(Bryan et al., 1997); Cretaceous detrital zircon dates from the Madura Shelf (Barham et al., 2016); Cretaceous detrital zircon dates from the Bight Basin (MacDonald et al.,
2013); basalts from the Casterton Formation in the Otway Basin (Mitchell et al., 1997); plutonic rocks from eastern Fiordland, New Zealand (Scott and Palin, 2008); dolerites in the
Northland Basin of New Zealand (Mortimer et al., 2009); the dating of low grade metamorphic dates in accretionary prisms in New Zealand (Mortimer et al., 2012); basalts from the
Marion Plateau (Jell, 2013); fault gouge dates from the Sydney Basin (Och et al., 2014); fault gouge dates from the Demon Fault (Babaahmadi et al., 2018); granite and rhyolite clasts
from the Lord Howe Rise (Mortimer et al., 2015); detrital zircon dates from the Cretaceous Winton and Mackunda Formations in the Eromanga Basin Tucker et al. (2016); detrital
zircon grains from a Late Jurassic accretionary complex of the Boghen terrane New Caledonia which is related to the Permian–Mesozoic Southeast-Gondwana arc system (Cluzel and
Meffre, 2002); and from tuffs and volcanogenic sandstones of the Walloon Coal Measures and Birkhead Formation in the Eromanga, Surat and Clarence-Moreton basins (Wainman et
al., 2018).

and calc-alkaline source. Bryan et al. (1997) also suggested that forma- Australia despite the remoteness of the continental margin
tion of the province immediately pre-dated continental rifting and initi- (Waschbusch et al., 2009; Jell, 2013). The 500 m thick package of strata
ated ocean crust formation in the Tasman Sea. The presence of a comprising the Walloon Coal Measures (the thickest sedimentary pack-
subduction zone would also help explain the asymmetrical pattern ob- age in the basin) was likely a consequence of when subduction veloci-
served evident with the variable thickness of the Walloon Coal Mea- ties peaked, and the horizontal length of subsidence increased as the
sures and older strata in the Surat Basin from east to west dip of the subduction plate decreased in the Middle to Late Jurassic
(Waschbusch et al., 2009; Smith et al., 2019). This is thought to have (Waschbusch et al., 2009).
been controlled by dynamic platform tilting related to the viscous cor- Our hypothesis, that the tuffs in the Surat Basin were sourced from
ner flow driven by westward-directed subduction beneath eastern the Whitsunday Igneous Association above a subduction zone, has
114 C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116

Table 3
Comparison of similar igneous rock datasets from the Walloon Coal Measures of the Surat Basin (this paper and Wainman et al. (2018)); Whitsunday Igneous Association (Ewart et al.
(1992), Bryan et al. (1997), Bryan et al. (2000) and Jell (2013)); Marion Plateau (Isern et al. (2002) and Jell (2013)); Lord Howe Rise (Mortimer et al., 2015); and the Teremba and Boghen
terranes of New Caledonia (Cluzel and Meffre (2002), Cluzel et al. (2012), Campbell et al. (2018)).

Rock unit/regions Age Igneous lithologies Parent magma type Phenocrysts Zircon present Tuff/ignimbrite Tectonic
present bed thickness environment
(where
present)

Walloon Coal 168 to 149 Ma Silicic tuffs and Calc-alkaline, Tuffs: quartz Yes, eruptive phase zircon Up to 2.13 m, Convergent
Measures, Surat (Middle to volcanogenic intermediate to felsic, and feldspars recorded across the basin, average 0.04 m plate boundary
Basin Late Jurassic) sandstones some tuffs possibly abundant (subduction)
sourced from a mafic
magma

Whitsunday 132 to 95 Ma Granite, rhyolite lavas Relatively young, Dacitic Some present in the Mount Typically 10 to Intraplate to
Igneous (Early and domes, non-radiogenic, ignimbrites: Salmon and Wycarbah 100 m and up continental rift
Association Cretaceous) basaltic lavas, calc-alkaline, tholeiitic Pyroxenes and Volcanics, character and to 1 km thick
volcanogenic sandstone basalt showing E-MORB Fe–Ti oxides abundance not recorded
and affinities
conglomerates, lithic Crystal-rich
breccias, coarse rhyolitic
lithic ignimbrites, ignimbrites:
dacitic to rhyolitic quartz,
ignimbrites and plagioclase,
phreatomagmatic hornblende,
deposits Fe–Ti oxides
and pyroxenes

Lavas (basaltic
to rhyolitic):
plagioclase and
pyroxene

Marion Plateau 162.0 Ma Reddish brown olivine Alkali basalt (bimodal) Basalt: Abundance and character Not recorded Rifting and
basaltic, tuffs, plagioclase, not recorded large igneous
volcaniclastic breccias olivine and province (LIP)
and pyroxene: magmatism (?)
conglomerates:

Lord Howe Rise 216 to 97 Ma Granite and vitrophyric I-type calc-alkaline Vitrophyric Yes, present in rhyolite Not recorded Convergent
rhyolite flows rhyolite flows: flows. Eruptive phase zircon plate boundary
plagioclase (subduction)

Teremba, Mid-Jurassic Teremba Terrane: Island arc tholeiite to Not recorded Detrital zircon, hundreds ~700 m (?) Convergent
Koh-Central (Teremba andesites, dacites, calc-alkaline plate boundary
Chain and Terrane) rhyolites and (subduction)
Boghen terranes Volcaniclastics
of New Caledonia Permian to
Early Koh-Central Terrane:
Cretaceous ophiolite, gabbro,
(Koh-Central dolerite, rare
Terrane) plagiogranite and
pillow
Early Jurassic basalts
to Early
Cretaceous Boghen Terrane:
(Boghen peridotite/serpentinite,
Terrane) pillow basalt and tuffs

potentially important implications for the tectonic setting of Gondwana intermediate to felsic composition from XRD data (dominated by quartz
during the Late Jurassic. Previous authors have suggested that the east- and feldspars that have altered to clays) and zircon geochemistry (mod-
ern margin of Gondwana was undergoing a period of extension and erate U values and elevated Y values). Analysis of tuff isopachs and zir-
rifting during the Late Jurassic (Bryan et al., 2012; Jell, 2013). However, con grain sizes indicate the tuffs originated from VEI 8 magnitude
our data instead supports the hypothesis of Bache et al. (2014) and eruptions and that the vents were between 280 and 1000 km away
Tucker et al. (2016) that the margin was an active subduction zone dur- from the Surat Basin to the palaeo east-southeast, from the Whitsunday
ing this timeframe until ~85 Ma when the seafloor spreading com- Igneous Association.
menced in the Tasman Sea. Despite the paucity of Late Triassic to Late Cretaceous magmatic
rocks along the eastern seaboard of Australia and the uncertainty on
10. Conclusions the nature of the eastern margin of Gondwana during this timeframe,
we argue that it is probable the tuffs originated from a long-lived (late
Using multiple lines of evidence, we demonstrate from our newly Palaeozoic to Cretaceous) continental arc related to the westward sub-
acquired data that the tuffs of the Jurassic Walloon Coal Measures of duction of palaeo-Pacific oceanic crust beneath eastern Australia. This
the Surat Basin represent pyroclastic fall deposits based on the angular also bears little or no relation to the eruption of a Large Igneous Province
shape of their clasts, and that their parent magmas were of an that produced substantial extrusive volumes, and those which are
C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116 115

known are Cretaceous rather than Jurassic in age as indicated by U–Pb Dai, S., Ward, C.R., Graham, I.T., French, D., Hower, J.C., Zhao, L., Wang, X., 2017. Altered
zircon age data. It will also help better constrain the timing of when volcanic ashes in coal and coal-bearing sequences: a review of their nature and signif-
icance. Earth Sci. Rev. 175, 44–47.
eastern Gondwana transitioned to a divergent margin, and define future Dulhunty, J., 1965. The Mesozoic age of the Garrawilla lavas in the Coonabarabran-
targets for ocean drilling to locate the parent igneous bodies in the Tas- Gunnedah District, Journal of the Proceedings of the Royal Society of NSW, pp.
man Sea. 105–109.
Dulhunty, J., Middlemost, E., Beck, R., 1987. Potassium–argon ages, petrology and geo-
chemistry of some Mesozoic igneous rocks in northeastern New South Wales. Journal
Acknowledgements and Proceedings of the Royal Society of NSW 120, 71–90.
Elliot, D.H. and Fleming, T.H., 2017. The Ferrar large Igneous Province: field and geochem-
ical constraints on supra-crustal (high-level) emplacement of the magmatic system.
We thank Jim Crowley, Debra Pierce and Alexandra Edwards from Geol. Soc. Lond., Spec. Publ., 463: SP463. 461.
Boise State University who assisted with mineral separation, prepara- Ewart, A., Schon, R., Chappell, B., 1992. The Cretaceous volcanic-plutonic province of the
tion and the dating of zircon grains using CA-TIMS and LA-ICPMS. We central Queensland (Australia) coast—a rift related ‘calc-alkaline’ province. Trans. R.
Soc. Edinb. Earth Sci. 83 (1–2), 327–345.
also thank Abbas Babaahmadi for his helpful comments on the geology Exon, N.F., 1976. Geology of the Surat Basin in Queensland. Bulletin of the Bureau of Min-
of eastern Australia during the Jurassic. eral Resources and Geophysics, Australia 166, 57–69.
Exon, N.F., Burger, D., 1981. Sedimentary cycles in the Surat Basin and global changes of
sea level. BMR Journal of Australian Geology & Geophysics 6, 153–159.
Funding
Fielding, C.R., 1996. Mesozoic sedimentary basins and resources in eastern Australia, Geo-
logical Society of Australia Abstracts. Geological Society of Australia 180–185.
This research did not receive any specific grant from funding agen- Fisher, R.V., Schmincke, H.U., 1984. Pyroclastic Rocks. Springer. Berlin. Heidelberg, New
cies in the public, commercial or not-for-profit sectors. York.
Gallagher, K., Dumitru, T.A., Gleadow, A.J.W., 1994. Constraints on the vertical motion of
eastern Australia during the Mesozoic. Basin Res. 6 (2–3), 77–94.
References Golonka, J., Ross, M., Scotese, C., 1994. Phanerozoic Paleogeographic and Paleoclimatic
Modeling Maps.
Babaahmadi, A., Uysal, I.T., Rosenbaum, G., 2018. Late Jurassic intraplate faulting (Demon Goscombe, P., Coxhead, B., 1995. Clarence-Moreton, Surat, Eromanga, Nambour, and
Fault) in eastern Australia: a link to subduction in eastern Gondwana and plate tec- Mulgildie Basins. Geology of Australian Coal Basins 1, 489–511.
tonic reorganisation. Gondwana Res. 66, 1–12. Greb, S., Eble, C., Hower, J., 1999. Depositional history of the Fire Clay coal bed (Late
Bache, F., Mortimer, N., Sutherland, R., Collot, J., Rouillard, P., Stagpoole, V., Nicol, A., 2014. Duckmantian), eastern Kentucky. USA. International Journal of Coal Geology 40 (4),
Seismic stratigraphic record of transition from Mesozoic subduction to continental 255–280.
breakup in the Zealandia sector of eastern Gondwana. Gondwana Res. 26 (3), Green, D. and Webb, A., 1974. Geochronology of the northern part of the Tasman Geosyn-
1060–1078. cline, The Tasman Geosyncline; A Symposium: Geological Society of Australia,
Barham, M., Kirkland, C., Reynolds, S., O'Leary, M., Evans, N., Allen, H., Haines, P., Hocking, Queensland Division, Brisbane, pp. 275–293.
R., McDonald, B., Belousova, E., 2016. The answers are blowin' in the wind: Ultra-dis- Green, P., Hoffmann, K., Brian, T., Gray, A., Murray, C., Carmichael, D., McKeller, J., Beeston,
tal ashfall zircons, indicators of Cretaceous super-eruptions in eastern Gondwana. Ge- J., Price, P., Smith, M., 1997. The Surat and Bowen Basins, south-east Queensland.
ology 44 (8), 643–646. Queensland Minerals and Energy Review Series, Queensland Department of Mines
Belousova, E., Griffin, W.L., O'Reilly, S.Y., Fisher, N., 2002. Igneous zircon: trace element and Energy 238.
composition as an indicator of source rock type. Contrib. Mineral. Petrol. 143 (5), Grimes, C., Wooden, J., Cheadle, M., John, B., 2015. “Fingerprinting” tectono-magmatic
602–622. provenance using trace elements in igneous zircon. Contrib. Mineral. Petrol. 170
Best, M.G., Christiansen, E.H., 1997. Origin of broken phenocrysts in ash-flow tuffs. Geol. (5–6), 1–26.
Soc. Am. Bull. 109 (1), 63–73. Hanchar, J.M., Miller, C.F., 1993. Zircon zonation patterns as revealed by
Blakely, R.J., Brocher, T.M., Wells, R.E., 2005. Subduction-zone magnetic anomalies and cathodoluminescence and backscattered electron images: Implications for interpre-
implications for hydrated forearc mantle. Geology 33 (6), 445–448. tation of complex crustal histories. Chem. Geol. 110 (1), 1–13.
Blevin, J. and Cathro, D., 2008. Australian southern margin synthesis. Project GA707, Client Hardman, J., Holford, S., Schofield, N., Bunch, M., Gibbins, D., 2018. The Warnie Volcanic
report to Geoscience Australia. froG Tech Pty Ltd, Canberra. Province: A Jurassic Volcanic Province in Central Australia. EarthArXiv https://doi.
Boult, P.J., 1996. An Investigation of Reservoir/Seal Couplets in the Eromanga Basin; Impli- org/10.31223/osf.io/9zmty (December 12).
cations for Petroleum Entrapment and Production: Development of Secondary Mi- Hawkesworth, C., Kemp, A., 2006. Using hafnium and oxygen isotopes in zircons to un-
gration and Seal Potential Theory and Investigation Techniques, University of South ravel the record of crustal evolution. Chem. Geol. 226 (3), 144–162.
Australia. Adelaide, Australia, p. 255. Hergt, J., Chappell, B., McCulloch, M., McDougall, I., Chivas, A., 1989. Geochemical and iso-
Bowring, S.A., Schmitz, M.D., 2003. High-precision U-Pb zircon geochronology and the topic constraints on the origin of the Jurassic dolerites of Tasmania. J. Petrol. 30 (4),
stratigraphic record. Rev. Mineral. Geochem. 53 (1), 305–326. 841–883.
Brown, R., Bonadonna, C., Durant, A., 2012. A review of volcanic ash aggregation. Physics Huff, W.D., Bergström, S.M., Kolata, D.R., 1992. Gigantic Ordovician volcanic ash fall in
and Chemistry of the Earth, Parts A/B/C 45, 65–78. North America and Europe: Biological, tectonomagmatic, and event-stratigraphic sig-
Bryan, S., 1996. The Whitsunday Volcanic Province (central Queensland) and the Otway/ nificance. Geology 20 (10), 875–878.
Gippsland Basins (Victoria): a comparison of Early Cretaceous rift-related volcano- Isern, A.R., Anselmetti, F.S., Andresen, N., Birke, T.K., Gartner, G.L.B., Burns, S.J., Conesa, G.
sedimentary successions. Geological Society of Australia Abstracts Geological Society A., Delius, H., Dugan, B., Eberli, G.P., 2002. Covering Leg 194 of the cruises of the Dril-
of Australia 124–133. ling Vessel JOIDES Resolution Townsville. Australia, to Apra Harbor, Guam Sites
Bryan, S., Constantine, A., Stephens, C., Ewart, A., Schön, R., Parianos, J., 1997. Early Creta- 1192–1199 3 January–2 (March 2001).
ceous volcano-sedimentary successions along the eastern Australian continental Jell, P.A., 2013. Geology of Queensland/edited by Peter A Jell. Geological Survey of Queens-
margin: implications for the break-up of eastern Gondwana. Earth Planet. Sci. Lett. land, Brisbane, 599 pp.
153 (1), 85–102. Jones, J., Veevers, J., 1983. Mesozoic origins and antecedents of Australia's Eastern High-
Bryan, S., Ewart, A., Stephens, C., Parianos, J., Downes, P., 2000. The Whitsunday Volcanic lands. J. Geol. Soc. Aust. 30 (3–4), 305–322.
Province, Central Queensland, Australia: lithological and stratigraphic investigations Klootwijk, C., 2009. Sedimentary basins of eastern Australia: Paleomagnetic constraints
of a silicic-dominated large igneous province. J. Volcanol. Geotherm. Res. 99 (1–4), on geodynamic evolution in a global context. Aust. J. Earth Sci. 56 (3), 273–308.
55–78. Königer, S., Lorenz, V., Stollhofen, H., Armstrong, R.A., 2002. Origin, age and stratigraphic
Bryan, S.E., Cook, A.G., Allen, C.M., Siegel, C., Purdy, D.J., Greentree, J.S., Uysal, I.T., 2012. significance of distal fallout ash tuffs from the Carboniferous–Permian continental
Early-mid Cretaceous tectonic evolution of eastern Gondwana: from silicic LIP Saar–Nahe Basin (SW Germany). Int. J. Earth Sci. 91 (2), 341–356.
magmatism to continental rupture. Episodes 35 (1), 142–152. Korsch, R.J., Totterdell, J.M., 2009. Subsidence history and basin phases of the Bowen,
Campbell, M.J., Shaanan, U., Rosenbaum, G., Allen, C.M., Cluzel, D., Maurizot, P., 2018. Gunnedah and Surat Basins. eastern Australia. Australian Journal of Earth Sciences
Permian rifting and isolation of New Caledonia: evidence from detrital zircon geo- 56 (3), 335–353.
chronology. Gondwana Res. 60, 54–68. Korsch, R., O'Brien, P., Sexton, M., Wake-Dyster, K., Wells, A., 1989. Development of Meso-
Carey, S., Sparks, R., 1986. Quantitative models of the fallout and dispersal of tephra from zoic transtensional basins in easternmost Australia. J. Geol. Soc. Aust. 36 (1), 13–28.
volcanic eruption columns. Bull. Volcanol. 48 (2–3), 109–125. Lerbekmo, J., 2002. The Dorothy bentonite: an extraordinary case of secondary thickening
Chesnut, D.R., 1983. Source of the volcanic ash deposit (flint clay) in the Fire Clay coal of in a late Campanian volcanic ash fall in central Alberta. Can. J. Earth Sci. 39 (12),
the Appalachian basin, Compte rendu, Tenth International Congress for Carboniferous 1745–1754.
Stratigraphy and. Geology 145–154. Lyons, P.C., Outerbridge, W.F., Triplehorn, D., Evans, H.T., Congdon, R.D., Capiro, M., Hess, J.,
Cluzel, D., Meffre, S., 2002. L'unité de la Boghen (Nouvelle-Calédonie, Pacifique sud- Nash, W.P., 1992. An Appalachian isochron: a kaolinized Carboniferous air-fall volca-
ouest): un complexe d'accrétion jurassique. Données radiochronologiques nic-ash deposit (tonstein). Geol. Soc. Am. Bull. 104 (11), 1515–1527.
préliminaires U Pb sur les zircons détritiques. Compt. Rendus Geosci. 334 (11), MacDonald, J.D., Holford, S.P., Green, P.F., Duddy, I.R., King, R.C., Backé, G., 2013. Detrital
867–874. zircon data reveal the origin of Australia's largest delta system. J. Geol. Soc. 170 (1),
Cluzel, D., Maurizot, P., Collot, J., Sevin, B., 2012. An outline of the geology of New Caledo- 3–6.
nia; from Permian-Mesozoic Southeast Gondwanaland active margin to Cenozoic Martin, M.A., Wakefield, M., MacPhail, M.K., Pearce, T., Edwards, H.E., 2013. Sedimentol-
obduction and supergene evolution. Episodes-Newsmagazine of the International ogy and stratigraphy of an intra-cratonic basin coal seam gas play: Walloon Subgroup
Union of Geological. Sciences 35 (1), 72. of the Surat Basin. eastern Australia. Petroleum Geoscience 19 (1), 21–38.
116 C.C. Wainman et al. / Journal of Volcanology and Geothermal Research 377 (2019) 103–116

Mason, B.G., Pyle, D.M., Oppenheimer, C., 2004. The size and frequency of the largest ex- Stern, R.J., Bloomer, S.H., 1992. Subduction zone infancy: examples from the Eocene Izu-
plosive eruptions on Earth. Bull. Volcanol. 66 (8), 735–748. Bonin-Mariana and Jurassic California arcs. Geol. Soc. Am. Bull. 104 (12), 1621–1636.
Matthews, N.E., Smith, V.C., Costa, A., Durant, A.J., Pyle, D.M., Pearce, N.J., 2012. Ultra-distal Sutherland, F., 1966. Considerations on emplacement of Jurassic Dolerites. Pap. Proc. R.
tephra deposits from super-eruptions: examples from Toba, Indonesia and Taupo Soc. Tasmania 133–146.
Volcanic Zone. New Zealand. Quaternary International 258, 54–79. Sutherland, F., 1978. Mesozoic-Cainozoic volcanism of Australia. Tectonophysics 48 (3),
McCabe, P.J., 1984. Depositional environments of coal and coal bearing strata. In: 413–427.
Rahmani, R., Flores, R. (Eds.), Sedimentology of Coal and Coal-bearing Sequences. In- Tatsumi, Y., 1989. Migration of fluid phases and genesis of basalt magmas in subduction
ternational Association of Sedimentologists, Hoboken, New Jersey, pp. 13–42. zones. Journal of Geophysical Research: Solid Earth 94 (B4), 4697–4707.
Middlemost, E., Dulhunty, J. and Beck, R., 1992. Some Mesozoic igneous rocks from north- Torsvik, T.H., Cocks, L.R.M., 2016. Earth History and Palaeogeography. Cambridge Univer-
eastern New South Wales and their tectonic setting, Journal and Proceedings of the sity Press.
Royal Society of NSW, pp. 1–11. Triplehorn, D., Bohor, B., 1986. Volcanic ash layers in coal: origin, distribution, composi-
Mitchell, M., Duddy, I., O'sullivan, P., 1997. Reappraisal of the age and origin of the tion and significance. Mineral Matter and Ash in Coal 301, 90–98.
Casterton Formation, western Otway Basin, Victoria. Aust. J. Earth Sci. 44 (6), Tucker, R.T., Roberts, E.M., Henderson, R.A., Kemp, A.I., 2016. Large igneous province or
819–830. long-lived magmatic arc along the eastern margin of Australia during the cretaceous?
Moore, G.T., Hayashida, D.N., Ross, C.A., Jacobson, S.R., 1992. Paleoclimate of the Insights from the sedimentary record. Geological Society of America Bulletin 128 (9–
Kimmeridgian/Tithonian (Late Jurassic) world: I. Results using a general circulation 10), 1461–1480.
model. Palaeogeogr. Palaeoclimatol. Palaeoecol. 93 (1), 113–150. Twidale, C.R., 1994. Gondwanan (Late Jurassic and Cretaceous) palaeosurfaces of the Aus-
Mortimer, N., Raine, J., Cook, R., 2009. Correlation of basement rocks from Waka Nui 1 and tralian craton. Palaeogeogr. Palaeoclimatol. Palaeoecol. 112 (1), 157–186.
Awhitu1, and the Jurassic regional geology of Zealandia. N. Z. J. Geol. Geophys. 52 (1), Valdes, P., Sellwood, B., Price, G., 1995. Modelling Late Jurassic Milankovitch climate var-
1–10. iations. In: House, M.R., Gale, A.S. (Eds.), Orbital Forcing Timescales and
Mortimer, N., McLaren, S., Dunlap, W., 2012. Ar-Ar dating of K-feldspar in low grade meta- Cyclostratigraphy. Geological Society, London. Special Publications, London, UK,
morphic rocks: example of an exhumed Mesozoic accretionary wedge and forearc, pp. 115–132.
South Island. New Zealand. Tectonics 31 (3). Veevers, J., 2006. Updated Gondwana (Permian–cretaceous) earth history of Australia.
Mortimer, N., Turnbull, R., Palin, J., Tulloch, A., Rollet, N., Hashimoto, T., 2015. Triassic–Ju- Gondwana Res. 9 (3), 231–260.
rassic granites on the Lord Howe Rise, northern Zealandia. Aust. J. Earth Sci. 62 (6), Veevers, J., Powell, C.M., Roots, S., 1991. Review of seafloor spreading around Australia. I.
735–742. Synthesis of the patterns of spreading. Aust. J. Earth Sci. 38 (4), 373–389.
Och, D., Offler, R., Zwingmann, H., 2014. Constraining timing of brittle deformation and Wainman, C.C., McCabe, P.J., 2018. Evolution of the depositional environments of the Ju-
fault gouge formation in the Sydney Basin. Aust. J. Earth Sci. 61 (3), 337–350. rassic Walloon Coal measures, Surat Basin, Queensland. Australia. Sedimentology
Orton, G., 1996. Volcanic environments. Sedimentary environments 485–567. 19, 1–26.
Paterson, B.A., Rogers, G., Stephens, W.E., 1992. Evidence for inherited Sm−Nd isotopes in Wainman, C.C., McCabe, P.J., Crowley, J.L., Nicoll, R.S., 2015. U–Pb zircon age of the Wal-
granitoid zircons. Contrib. Mineral. Petrol. 111 (3), 378–390. loon Coal measures in the Surat Basin, southeast Queensland: implications for
Pyle, D.M., 1989. The thickness, volume and grainsize of tephra fall deposits. Bull. palaeogeography and basin subsidence. Aust. J. Earth Sci. 62 (7), 807–816.
Volcanol. 51 (1), 1–15. Wainman, C.C., McCabe, P.J., Crowley, J.L., 2018. Solving a tuff problem: defining a new
Rapp, R.P., Irifune, T., Shimizu, N., Nishiyama, N., Norman, M.D., Inoue, T., 2008. Subduc- chronostratigraphic framework for Middle to Upper Jurassic strata in eastern Austra-
tion recycling of continental sediments and the origin of geochemically enriched res- lia using U-Pb CA-TIMS zircon dates. AAPG Bull. 102 (6), 1141–1168.
ervoirs in the deep mantle. Earth Planet. Sci. Lett. 271 (1), 14–23. Waschbusch, P., Korsch, R.J., Beaumont, C., 2009. Geodynamic modelling of aspects of the
Rosi, M., Vezzoli, L., Castelmenzano, A., Grieco, G., 1999. Plinian pumice fall deposit of the Bowen, Gunnedah, Surat and Eromanga Basins from the perspective of convergent
Campanian Ignimbrite eruption (Phlegraean Fields, Italy). J. Volcanol. Geotherm. Res. margin processes. Aust. J. Earth Sci. 56 (3), 309–334.
91 (2), 179–198. Watson, D., Schofield, N., Jolley, D., Archer, S., Finlay, A.J., Mark, N., Hardman, J., Watton, T.,
Sarna-Wojcicki, A.M., Shipley, S., Waitt Jr., R.B., Dzurisin, D., Wood, S.H., 1981. Areal distri- 2017. Stratigraphic overview of Palaeogene tuffs in the Faroe–Shetland Basin, NE At-
bution, thickness, mass, volume, and grain size of air-fall ash from the six major erup- lantic margin. Journal of the Geological Society: jgs2016–2132.
tions of 1980. Yago, J.V.R., 1996. Basin Analysis of the Middle Jurassic Walloon Coal Measures in the
Schmid, R., 1981. Descriptive nomenclature and classification of pyroclastic deposits and Great Artesian Basin, Australia. PhD Thesis, University of Queensland. Brisbane, Aus-
fragments: Recommendations of the IUGS Subcommission on the Systematics of Ig- tralia, p. 276.
neous Rocks. Geology 9 (1), 41–43. Yago, J.V., Fielding, C., 2015. Depositional Environments and Sediment Dispersal patterns
Schutter, S.R., 2003. Hydrocarbon occurrence and exploration in and around igneous of the Jurassic Walloon Subgroup in Eastern Australia. In: Lodwick, W. (Ed.), Eastern
rocks. Geol. Soc. Lond., Spec. Publ. 214 (1), 7–33. Australasian Basins Symposium, a Powerhouse Emerges; Energy for the Next Fifty
Scott, J., Palin, J., 2008. LA-ICP-MS U-Pb zircon ages from Mesozoic plutonic rocks in east- Years, Petroleum Exploration Society of Australia Special Publication. PESA. Perth,
ern Fiordland. New Zealand. New Zealand Journal of Geology and Geophysics 51 (2), Australia, pp. 141–150.
105–113. Young, I.F., 1987. Late Jurassic channels and associated fill, Eromanga Basin (ATP 259P),
Sellwood, B.W., Valdes, P.J., 1997. Geological evaluation of climate General Circulation Queensland: a seismic stratigraphic study. Explor. Geophys. 18 (2), 238–242.
Models and model implications for Mesozoic cloud cover. Terra Nova 9 (2), 75–78.
Smith, T., Bianchi, V., Capitanio, F., 2019. Subduction geometry controls on dynamic to-
pography: implications for the Jurassic Surat Basin. Aust. J. Earth Sci. 1–11.

You might also like