17
PLEASE SCROLL DOWN FOR ARTICLE This article was downloaded by: [Retallack, G. J.][University of Oregon] On: 2 June 2009 Access details: Access Details: [subscription number 906605100] Publisher Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Australian Journal of Earth Sciences Publication details, including instructions for authors and subscription information: http://www.informaworld.com/smpp/title~content=t716100753 Cambrian, Ordovician and Silurian pedostratigraphy and global events in Australia G. J. Retallack a a Department of Geological Sciences, University of Oregon, Eugene, OR, USA Online Publication Date: 01 June 2009 To cite this Article Retallack, G. J.(2009)'Cambrian, Ordovician and Silurian pedostratigraphy and global events in Australia',Australian Journal of Earth Sciences,56:4,571 — 586 To link to this Article: DOI: 10.1080/08120090902806321 URL: http://dx.doi.org/10.1080/08120090902806321 Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf This article may be used for research, teaching and private study purposes. Any substantial or systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Australian Journal of Earth Sciences Cambrian, …blogs.uoregon.edu/gregr/files/2013/07/australianjournalofearth... · Australian Journal of Earth Sciences ... Victoria) show variation

Embed Size (px)

Citation preview

PLEASE SCROLL DOWN FOR ARTICLE

This article was downloaded by: [Retallack, G. J.][University of Oregon]On: 2 June 2009Access details: Access Details: [subscription number 906605100]Publisher Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,37-41 Mortimer Street, London W1T 3JH, UK

Australian Journal of Earth SciencesPublication details, including instructions for authors and subscription information:http://www.informaworld.com/smpp/title~content=t716100753

Cambrian, Ordovician and Silurian pedostratigraphy and global events inAustraliaG. J. Retallack a

a Department of Geological Sciences, University of Oregon, Eugene, OR, USA

Online Publication Date: 01 June 2009

To cite this Article Retallack, G. J.(2009)'Cambrian, Ordovician and Silurian pedostratigraphy and global events in Australia',AustralianJournal of Earth Sciences,56:4,571 — 586

To link to this Article: DOI: 10.1080/08120090902806321

URL: http://dx.doi.org/10.1080/08120090902806321

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

This article may be used for research, teaching and private study purposes. Any substantial orsystematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply ordistribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contentswill be complete or accurate or up to date. The accuracy of any instructions, formulae and drug dosesshould be independently verified with primary sources. The publisher shall not be liable for any loss,actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directlyor indirectly in connection with or arising out of the use of this material.

Cambrian, Ordovician and Silurian pedostratigraphyand global events in Australia

G. J. RETALLACK

Department of Geological Sciences, University of Oregon, Eugene, OR 97403, USA([email protected]).

Paleosols are used for stratigraphic correlation of Quaternary sediments because they reflectchronologically synchronous changes of glacial and interglacial paleoclimates. One obviouspaleoclimatic indicator is depth to carbonate nodules as a proxy for mean annual precipitation.Paleosol sequences of the Flinders Ranges (Cambrian, South Australia), Kalbarri (Ordovician, WesternAustralia) and Grampians (Silurian, Victoria) show variation in depth to carbonate through time, butoccasional transient excursions to very deep (450 cm) carbonate. These pedostratigraphic spikescorrespond to times of unusually high precipitation and temperature, but also to internationallyrecognised negative carbon isotope anomalies and mass extinctions, including the Late Ordovicianextinctions. These deep-calcic excursions can be correlated across Australia. Like magnetostratigraphyand chemostratigraphy, pedostratigraphic spikes provide calibration points for sequences less-precisely dated by paleontology or radiometry. Also like magnetostratigraphy and chemostratigraphy,pedostratigraphic correlations are best with numerous excursions spaced at irregular intervals andassuming constant local rate of sedimentation. Pedostratigraphy refines geological dating of Paleozoicand Precambrian redbeds and early terrestrial fossils.

KEY WORDS: Australia, Cambrian, Ordovician, paleoclimate, paleosol, pedostratigraphy, Silurian.

INTRODUCTION

Pedostratigraphy is stratigraphic correlation using

paleosols. Its fundamental soil stratigraphic unit is the

geosol (North American Commission on Stratigraphic

Nomenclature 1982) or pedoderm (Stratigraphic Nomen-

clature Committee 1973). The Sangamon Geosol of the

North American midwest is an unusually well-devel-

oped paleosol compared with glacial paleosols above and

below (Follmer 1978), and represents near-modern

interglacial conditions of ca 126 ka (Gradstein et al.

2004). Naming a buried landscape can be difficult,

because soils vary from hilltop to valley bottom, under

different vegetation, and with other soil-forming factors

(Follmer 1978). A simpler approach to pedostratigraphy

is to focus on recognisable features of laterally extensive

paleosols, such as violet horizons of the German

Triassic, Buntsandstein (Ortlam 1971), or clayey paleo-

sols of the North American Jurassic, Morrison Forma-

tion (Demko et al. 2004). Inflexions in time series of such

data can then be used to correlate sections in a manner

comparable with correlating carbon-isotopic chemostrat-

igraphic spikes (Retallack & Krull 2006), magnetostrat-

igraphic reversals (Morrison & Ellwood 1986; Schmidt &

Hamilton 1990), or geophysical logs (Evans et al. 2007).

Like these other forms of sedimentary correlation,

pedostratigraphic correlation is based on an assumption

of constant sedimentation rates, and this assumption is

only rejected when there is strong deviation from non-

linearity. Also like these other forms of correlation,

pedostratigraphy does not produced numerical age

estimates, but extends the geographic and temporal

reach of pre-existing radiometric and paleontological tie

points. This paper records variation in depth from the

top of a paleosol to its zone of calcareous nodules (Bk

horizon of Soil Survey Staff 2000) in sequences of many

successive paleosols. These simple measurements can

be made quickly in outcrop and core to reveal distinc-

tive patterns suitable for lithological correlation of

sparsely fossiliferous redbeds of Cambrian to Silurian

age in Victoria, South Australia and Western Australia

(Figure 1). Depth to Bk has been shown to be highly

correlated with mean annual precipitation in modern

soils (Retallack 2005a), and this relationship can be used

to estimate paleoprecipitation from paleosols once

corrected for compaction due to burial (Sheldon &

Retallack 2001). Recognition of the top of a paleosol is

not always straightforward, because some paleosols had

cumulic horizons (Soil Survey Staff 2000), in which

additional increments of sediment were at first incorpo-

rated into the soil by surviving plants, until these plants

were overwhelmed by additional sediment. Even if not

preserved as the usual abrupt discontinuity, the tops of

paleosols can be recognised by truncation of rhizocon-

cretions, drab-haloed root traces, soil cracks or

other indications of soil formation (Figures 2–4). Also

Australian Journal of Earth Sciences (2009) 56, (571–586)

ISSN 0812-0099 print/ISSN 1440-0952 online � 2009 Geological Society of Australia

DOI: 10.1080/08120090902806321

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

problematic for this measure of paleoprecipitation is

erosion of soil prior to burial, which is sometimes

visible as erosional microrelief, as calcareous nodules

from lower in paleosols redeposited in overburden, or

from carbonate benches of unusual thickness and

cementation found today in old (pre-Holocene) and

eroded soils (Retallack 2001). Some degree of erosion

may compromise precise paleoclimatic interpretations,

but this study emphasises exceptionally deep-calcic

paleosols, which would not be recognised if eroded.

Deep-calcic paleosols are rare in most redbed sequences.

A recent compilation of 3718 Permian to Holocene

paleosols from Utah and Montana found only 102 deep-

calcic paleosols at 40 distinct stratigraphic levels

correlative with times of crisis: mass extinctions, high

atmospheric CO2, carbon and oxygen isotopic excur-

sions, and fossil lagerstatten (Retallack 2005b, in press).

In Australia, Paleozoic deep-calcic paleosols correlate

with global environmental crises identified in marine

rocks by Talent et al. (1993), Saltzman (2005) and Zhu

et al. (2006).

MATERIALS AND METHODS

This work involved measuring stratigraphic sections,

usually formation type sections, and also observations

of drillcore. Thickness and size of soil features were

measured using a milliner’s tape. Munsell colour and

reaction with dilute acid were also recorded. Depth to

calcic horizon, as well as thickness of paleosol with

nodules and size of nodules were measured, because all

have paleoenvironmental significance (Retallack 2005a).

Comprehensive petrographic and geochemical study of

Cambrian paleosols from South Australia is presented

elsewhere (Retallack 2008), but was not completed for

Ordovician and Silurian paleosols of Western Australia

and Victoria. For each area studied, previous studies of

geological age are summarised, and known compro-

mising burial alteration considered, before interpreting

new pedostratigraphic data and correlations. The time-

scale used here is Avalonian (Britain and Newfound-

land), because this is a coherent scheme established

over a single geological region, and because Cambrian

Figure 1 Australian onshore surface sedimentary basins, with localities and boreholes of Cambrian to Devonian paleosols

examined in this study.

572 G. J. Retallack

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

and Ordovician international stages are not formally

agreed (Gradstein et al. 2004).

PALEOSOL RECOGNITION

Paleosols were discovered largely in sedimentary facies

known for a suite of non-marine indicators: redbeds,

loessites, halite hopper-casts, paleokarst, wind-dissected

ripples (setulfs), climbing translatant cross stratifica-

tion, mud cracks, and arthropod trackways (Mawson &

Segnit 1949; Clarke 1990; Moore 1979, 1990; Hocking 1991;

Trewin & McNamara 1994; Cayley & Taylor 1997).

Vascular plant roots are a key feature for recognition

of Silurian and younger paleosols (Retallack 1992; Driese

et al. 1997, 2000), but were seen in few paleosols identified

in this study. Instead, paleosol recognition was based on

soil horizons and soil structures (Figures 2–4).

Soil horizons show gradational changes below shar-

ply truncated surfaces, regarded as ancient exposure

surfaces because of their more intense weathering and

mud cracks, and diffuse zones of carbonate nodules. In

essence, the top of a paleosol is recognised at the contact

between clear primary bedding and massive, hackly,

reddened rock, grading downward into increasingly

clear bedding again. Gradational change in clay content

due to hydrolytic weathering can be superficially like

graded bedding, but is less regular and redder than

Figure 2 Cambrian–Ordovician paleosols of South Australia. (a) Dawson Hill capped with Grindstone Range Sandstone

(white) on Pantapinna Sandstone (red), both dipping west (left) from central Balcoracana Creek (31.178368S, E138.917858E). (b,

c) Red sandy Aridisols (Adla pedotype) in (b) Grindstone Range Sandstone (27 m in local section) and (c) upper Pantapinna

Sandstone (0.3 m in local section), both in bluffs south of central Balcoracana Creek (31.172468S, 138.929638E). (d) Red sandy

Entisol (Wandara pedotype) in Trainor Hill Sandstone at 416.1 m in Byilkaoora 2 borehole (27.29994518S, 133.45075958E) ca

500 Ma , early Late Cambrian. (e) Red silty Vertisol (Viparri pedotype) in Moodlatana Formation at 456.6 m in Lake Frome 2

borehole (31.04681618S, 139.79473568E) ca 510 Ma, Middle Cambrian.

Early Paleozoic pedostratigraphy 573

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

marine or lacustrine turbidites. Soil formation depletes

easily weathered grains such as feldspar and rock

fragments, to produce an upward increase in resistant

minerals such as quartz, and products of weathering

such as clay and iron oxides. Evidence for soil oxidation

of iron in well-drained paleosols includes even distribu-

tion of hematite pigment in carbonate-cemented silt-

stones, sharp boundaries between red and green beds,

mottles only a few millimetres thick within texturally

homogenous siltstones, and red clasts at the base of

green-grey channel sandstone facies (Moore 1990;

Retallack 2008). Gamma-ray survey of one of these

sequences (Evans et al. 2007) showed that paleosols were

less radioactive than their enclosing sediments, pre-

sumably due to weathering of naturally radiogenic

minerals before burial.

Figure 3 Ordovician paleosols of the Tumblagooda Sandstone, Western Australia. (a) View south along Murchison River from

Z-bend Lookout, Kalbarri National Park (27.63538S, 114.456638E) showing end-Llanvirnian deep-calcic paleosol half-way-up

left slope. (b) Sandy Aridisol 3 km southeast of Ross Graham Lookout (27.8128698S, 114.498078E) at 447.1 m in composite

section (ca 472 Ma early Llanvirnian). (c) Sandy Aridisol at Red Bluff (27.747088S, 114.137708E) at 1108.1 m in composite section

(ca 443 Ma, Hirnantian). (d) Rhizomorph in Entisol overlain by red claystone breccia at 450.1 m in Wendy 1 core (28.299058S,

115.0166818E: ca 445 Ma, Hirnantian). (e) Drab top with bioturbation and calcareous nodules in Aridisol at 442.9 m in Wendy 1

core (ca 444 Ma, Hirnantian). (f) Drab-haloes and calcareous nodules in Aridisol at 514.4 m in Wendy 1 core (ca 451 Ma, late

Caradocian).

574 G. J. Retallack

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

Distinctive paleosol structures include nodules of

micrite, sand crystals (after gypsum), rosettes of anhy-

drite (replacing gypsum), and silica geodes pseudomor-

phous after gypsum rosettes. Such fossil soil structures

contribute to the hackly and rubbly appearance of

paleosols, in contrast with bedded sedimentary rocks.

Calcic horizons show considerable variation in develop-

ment from calcareous mottles to hard carbonate nodules

Figure 4 Silurian and Devonian paleosols of Grampians National Park, Victoria. (a) Case-hardened paleosols showing sharp

tops and graded lower horizons in the Serra Sandstone [same level as (d, e) in core] along Wonderland Trail (37.1509338S,

142.50548E) at about 3700 m in composite section (ca 417 Ma, Pridolian). (b) Sandy Aridisol in middle Wartook Sandstone in

roadcut 2 km south of Zumstein (37.107288S, 142.399318E) at 3901 m in composite section (ca 407 Ma, early Emsian). (c) Silty

Aridisol in middle Silverband Formation in roadcut 2 km south of Silverband–Dunkeld road junction (37.129808S,

142.531978E) at 2303 m in composite section (ca 426 Ma, Wenlockian). (d, e) Sandy Entisol, Inceptisol and Aridisol in Serra

Sandstone (respectively, left to right) at 56.1, 57.3 and 46.3 m in VIMP10 core, northern Grampians (37.178938S, 142.542088E)

*2800 m in composite section (ca 420 Ma, Ludlovian).

Early Paleozoic pedostratigraphy 575

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

of increasing size, and some carbonate benches, as in

aridland soils today (Gile et al. 1966, 1980). The most

spectacular soil structures are deep cracks and ridges

interpreted as vertic structure (mukkara) of swelling

clay soils (Vertisols: Driese & Foreman 1992). Unlike

gravity-driven load structures or mud volcanoes (Owen

2003), or seismically induced sand blows, contorted

lamination, dykes and breccia (Wheeler 2002), this

distinctive deformation fades downward from an un-

dulating, erosionally truncated surface to undeformed

layers. Some of these paleosols have burrows and track-

ways described by Trewin & McNamara (1994) and

Gouramanis et al. (2003), and are evidence of early colo-

nisation of land by arthropods and other invertebrates.

Cambrian paleosols of the Flinders Ranges were the

most varied of the paleosols studied (Retallack 2008).

Ordovician paleosols of the Tumblagooda Sandstone and

Silurian paleosols of the Grampians Group were com-

parable to a subset of the Cambrian paleosols (Grind-

stone Range Sandstone profiles only). The most common

paleosols in all areas examined were red sandy profiles

with cracking, clay enrichment and green mottling

near the surface over an horizon of sandy micrite

nodules 1–5 cm in diameter (Adla and Natala pedotypes

of Retallack 2008).

CAMBRIAN OF FLINDERS RANGES AND OFFICERBASIN

Cambrian strata flank ridges of Proterozoic sandstones

which form the high peaks of the Flinders Ranges of

South Australia (Figure 1). Mawson’s (1938, 1939a, b)

sections in Brachina and Parachilna Gorges, and on

Ten Mile and Balcoracana Creeks near Wirrealpa

Station (Figure 5) were re-assessed, taking into account

subsequent stratigraphic and structural mapping by

Dalgarno (1964), Dalgarno & Johnson (1966) and Clarke

(1990). This Cambrian sequence dips outward from the

Blinman Dome, a diapiric breccia in the Arrowie Basin,

between the Curnamona Craton to the east and the

Gawler Craton to the west (Jago et al. 2006). Paleocur-

rents in the Moodlatana, Pantapinna and Grindstone

Range Formation were to the north and east (Wopfner

1970; Stock 1974), indicating source terranes in the

Gawler Craton, rather than the distant eastern Curna-

mona Craton. Also examined were cores from Ungoolya

1 and Byilkaoora 2 from the central Officer Basin

(Lindsay & Leven 1996; Tingate & Duddy, 2002), an

intracratonic basin originally draining the eastern

Yilgarn Craton of northern South Australia and eastern

Western Australia (Figure 1).

Published age constraints

Several stratigraphic levels can be dated within the

section measured by Mawson (1939a) in Ten Mile Creek

(Figure 2a). A marked negative carbon isotope (d13Ccarb.)

excursion in the Woodendinna Dolomite, equivalent to

1152 m in Ten Mile Creek, has been correlated with a

538 Ma carbon isotope excursion in Siberia (by Tucker

1991; Kirschvink & Raub 2003). A tuff in the basal Billy

Creek Formation at 2389 m in Ten Mile Creek has

yielded a 206Pb/238U SHRIMP zircon (SL13 standard) age

of 522.8+ 1.8 Ma (Haines & Flottmann 1998; Gravestock

& Shergold 2001). The first appearance of the trilobite

Figure 5 Geological map of Cambrian rocks near Wirrealpa, eastern Arrowie Basin, Flinders Ranges, South Australia

(modified from Dalgarno & Johnson 1966) showing sections examined for paleosols.

576 G. J. Retallack

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

Redlichia guizhouensis in the basal Wirrealpa Lime-

stone at 3170 m in Ten Mile Creek has been correlated

with the late Lungwangmaioan R. guizhouensis zone of

China (Jago et al. 2006; Paterson & Brock 2007), which

is dated at 511.5 Ma by Gradstein et al. (2004). Two

other internationally correlatable trilobite occurrences

are the first appearance of Onaraspis rubra in the

Moodlatana Formation at 3533 m in Ten Mile Creek

(Jago et al. 2002), taken as coeval with the 509.5 Ma

Oryctocephalus indicus zone (Geyer & Shergold 2000;

Gradstein et al. 2004), and the 504 Ma agnostid Leiopyge

laevigata in the uppermost Balcoracana Formation in

Lake Frome 3 core (Daily & Forbes 1969), equivalent to

3882 m in Ten Mile Creek. The radiometric date and

Leiopgye have been considered controversial (Retallack

2008). A final constraint is folding and uplift of the

northern Flinders Ranges with intrusion of the British

Empire Granite, dated by U–Pb isochron at 446.3+ 2.2

to 441.1+ 2.1 Ma (Elburg et al. 2003; Foden et al. 2006),

which is Late Ordovician to Early Silurian (Gradstein

et al. 2004).

Age constraints on Cambrian rocks of the Officer

Basin come principally from the trilobite Abadiella

officerensis from the Ouldburra Formation from 967.6–

970.1 m in Manya 6 core (Jago et al. 2002). Also

suggestive of a mid-late Atdabanian age (ca 520 Ma:

Gradstein et al. 2004) for the Ouldburra Formation are

the acritarchs Ceratophyton dumufuntum, Redkinia sp.,

and Polygonium sp. (Jago et al. 2006). Unidentified Early

Cambrian redlichiid trilobites were also found in the

Observatory Hill Formation at 87.85–333 m in Marla 1

core (Jago & Youngs 1980) and a small shelly fossil

comparable with late Early Cambrian (ca 515 Ma)

Biconulites from the type section of the same formation

at Observatory Hill (Gatehouse 1976). Regional uncon-

formities bounding the Marla Group are apparent in

outcrop in the Mt Johns Range (Benbow 1982) and from

regional correlation of drillcore (Lindsay & Leven 1996;

Tingate & Duddy 2002).

Burial alteration

The 5 km of Cambrian strata in Ten Mile Creek (Mawson

1939a) include zircons with fission-track ages indicating

Late Paleozoic exhumation (temperatures51108C), as

well as post-mid-Cretaceous cooling (Mitchell et al. 1998,

2002). This is compatible with evidence from the

Triassic, Leigh Creek Coal Measures, which are over

610 m thick, and include sub-bituminous coals (Parkin

1969). Fluorescence colour and intensity of organic

matter indicates increasing thermal alteration down

the Cambrian section: equivalent to sub-bituminous to

bituminous vitrinite reflectance of 0.5–1.53% in the

Balcoracana Formation, 0.8–0.9% in Moodlatana Forma-

tion, 1.11% in Wirrealpa Limestone, and 1.16% in

Oraparinna Shale (Zang 2002). Organic-matter matura-

tion data are compatible with burial temperatures of

about 1008C at the top of the sequence, and about 2008C in

the Oraparinna Shale (Tissot & Welte 1984; Diessel 1992),

and these values extrapolate to 258C at 8569 m, implying

2.7 km of missing overburden. Comparable alteration is

apparent from clay crystallinity studies, which show

smectite remaining only in the Billy Creek and higher

formations, and progressive illitisation downsection to

lower greenschist facies in the Parachilna Formation

(Retallack 2008). An additional 3 km of rock above the

highest paleosol in the Grindstone Range Sandstone at

5599 m, would have created burial compaction to 58.9%

of original thickness at that level, and 52.3% at the base

of the Parachilna Formation at 1052 m, using the

compaction algorithm of Sheldon & Retallack (2001).

Burial alteration was less profound in the Officer

Basin where the sedimentary sequence is more gently

deformed. Reflectance-equivalents calculated from the

methyphenanthrene index for the Cambrian Ouldburra

Formation is 1.07%, suggesting a maximum tempera-

ture of 1458C before Permian (ca 280 Ma dated by

fission track) cooling to below 95–1008C, and subse-

quent minor fluctuation to current 36–388C at 450–500 m

depth (Tingate & Duddy 2002). Like the Ten Mile Creek

section, pedogenic carbonate nodules in the redbeds

are strongly reactive to acid, micritic and unrecrystal-

lised in core. An additional 3 km of overburden above

the top of the Ungoolya and Byilkaoora bores would

have created burial compaction to 53.8% of original

thickness at the top of the Trainor Hill Sandstone and

55.0% at the base of the Observatory Hill Formation,

using the compaction algorithm of Sheldon & Retallack

(2001).

Pedostratigraphic ages

Ten Mile Creek and Balcoracana Creek sections reveal a

pattern of multiple deep-calcic paleosol horizons within

the upper Moodlatana and Balcoracana Formations, but

few such deep-calcic paleosols in the upper and lower

parts of the sequence, a pattern repeated in the Officer

Basin drillholes (Figure 6). Each local sequence is

plotted at different scales reflecting local sedimentation

rate, but the proportional spacing of deep-calcic hor-

izons in each are too similar to be due to chance. This

irregularity of spacing and magnitude allows correla-

tion in a manner comparable with geophysical logs. In

addition, published ages show a linear relationship with

stratigraphic level of high statistical significance

(R2¼ 0.99), supporting an assumption of steady accumu-

lation rate (Figure 6).

New results of the age model (Figure 6) is extrapola-

tion to an Ordovician (Tremadocian) age of the basal

Grindstone Range Formation in the Arrowie Basin, and

an age no younger than mid-Late Cambrian for the

Trainor Hill Sandstone of the Officer Basin. This

approach thus supports ages implied by the controver-

sial identification of the trilobite Leiopyge laevigata, and

the disputed 206Pb/238U SHRIMP zircon date from the

Billys Creek Formation (Retallack 2008). Extrapolating

backward is unsurprising: both sequences fall short of

the Cambrian–Precambrian boundary, as already

known (Jago et al. 2006).

ORDOVICIAN OF KALBARRI AND OFFICER BASIN

The Tumblagooda Sandstone is red and white quartz-

rich sandstone exposed in gorges of the Murchison

River and coastal cliffs of Kalbarri National Park,

Early Paleozoic pedostratigraphy 577

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

Western Australia (Figure 1). Hocking’s (1991) com-

posite section of the formation was traversed from

Hardabut Pool to the Z-bend, The Loop, Bracken Point,

and then coastal sections of Red Bluff. The sequence is

well known for its non-marine trace fossils, including

trackways of large arthropods (Trewin & McNamara

1994), and a body fossil of perhaps the oldest known land

animal (probably amphibious), the euthycarcinoid Kal-

barria brimmellae (McNamara & Trewin 1993). The

sequence of paleosols is remarkably similar to that of

the Tumblagooda Sandstone in the Wendy 1 core 100 km

to the southeast (Victoria Petroleum N. L. 2004), and to

the Munda Group, unconformably overlying the Marla

Group in Ungoolya 1 core from the Officer Basin, South

Australia (Tingate & Duddy 2002).

Published age constraints

The Tumblagooda Sandstone was considered Silurian

because of Ludlovian (421 Ma: Gradstein et al. 2004)

conodonts in the overlying Dirk Hartog Group in Dirk

Hartog 17B core (Philip 1969). A Silurian age was also

supported by Pb mineralisation ages of 434+ 16 Ma

(Llandoverian) in the nearby Northampton Block

(Richards et al. 1985), and assumed lack of mineralisa-

tion of the sandstones (Hocking 1990). My observations

of gossans associated with faults near Fourways,

Hardabut and Mooliabaanya suggest that the Tumbla-

gooda Sandstone may pre-date mineralisation. Gorter

et al. (1994) reported Cambrian–Ordovician conodonts

from the lower Tumblagooda Sandstone in Wandagee 1

core, but these rare fragments were considered re-

worked by Trewin & McNamara (1994). More conclu-

sive evidence for a pre-Silurian age comes from

conodonts (Ozarkodina broenlundi) in the Marron

Member (Ajana Formation, Dirk Hartog Group) over-

lying the Tumblagooda Sandstone in Coburn 1 core

(Yasin & Mory 1999) at a stratigraphic level of 907.3–

908.2 m (equivalent to 1265 m in Figure 7). This

conodont is within the Pterospathognathus celloni zone,

and mid-Telychian (432 Ma: Gradstein et al. 2004).

Another age constraint is a paleomagnetic reverse-to-

normal transition near the Gabba Gabba Member

conglomerate at 1115 m in coastal outcrops near Red

Bluff (Schmidt & Hamilton 1990), which can be corre-

lated with basal Silurian (ca 444 Ma: Gradstein et al.

2004) reversed to normal polarity at Ringgold Gap,

Georgia, USA (sites 24–25 of Morrison & Ellwood 1986;

Rindsberg & Chowns 1986).

Other age constraints apply to Ungoolya 1 in the

Officer Basin, South Australia, where the Mt Chandler

Sandstone overlies volcanics dated by K–Ar at

484+ 4 Ma (Table Hill Volcanics of Stevens & Apak

1999) and 483.6+ 20 Ma and 493.7+ 20 Ma (Kulyong

Volcanics of Major & Teluk 1967, corrected using

Dalrymple 1979). The Indulkana Shale was also dated

by Rb–Sr isochron at 460+ 15 Ma (Webb 1978), which is

compatible with correlation of the basal Indulkana

Shale with marine transgression of the Horn Valley

Siltstone of the Amadeus Basin and its Castlemainian

Figure 6 Cambrian age model and pedostratigraphic correlations of the Arrowie and Officer Basins, South Australia. Metre

scales differ according to local sediment accumulation rates, but proportional distribution of deep-calcic events is preserved

in each, and shows highly significant correlation.

578 G. J. Retallack

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

trilobite (Lycophron freemani) assemblage (Laurie 2006),

equivalent to Llanvirnian of Avalonia and dated at ca

468 Ma (Gradstein et al. 2004).

Burial alteration

The Tumblagooda Sandstone is up to 1400 m thick in

deep wells onshore (Iasky et al. 2002) and in outcrop

(Hocking 1991), though up to 4633 m has been inferred

from seismic lines offshore (Iasky et al. 2002). It is mostly

flat-lying, with exception of the basal 400 m near

Hardabut Pool, which dips at 98 west on strike 2118magnetic azimuth (for 2007). Pedogenic micrite nodules

in the Wendy 1 core reacted strongly with acid, but some

small nodules in outcrop showed weak reaction, because

of siliceous case hardening. Maximum conodont color

alteration index in the overlying Dirk Hartog Group in

Coburn 1 well is 1, equivalent to reflectance value of

about 0.8%, and temperatures5808C (Yasin & Mory

1999), and modelled alteration down into the upper

Tumblagooda Sandstone in that well is51% reflectance

(Iasky & Mory 1999). Apatite fission-track analyses of the

Ajana Sandstone in Coburn 1 are evidence of thermal

events of about 100–1058C in the Permian and Late

Jurassic, but5808C during the Tertiary (Yasin & Mory

1999). At a geothermal gradient of 258C/km, this gives an

additional 3 km of eroded section, and compaction of

59.6% at 315.4 m in Wendy 1 core, 60.6% in the highest

paleosol at Red Bluff, and 56.9% for the lowest paleosol

near Hardabut Pool, using algorithms of Sheldon &

Retallack (2001).

Pedostratigraphic ages

Paleosols in Kalbarri National Park, like those in

Wendy 1 core in the nearby Perth Basin and Ungoolya

1 core of the Officer Basin all show a similar pattern of

two upper deep-calcic paleosols above three more widely

spaced deep-calcic paleosols (Figure 8). Once again the

different scales for each section reflect different local

sedimentation rates, but the proportional spacing of

deep-calcic horizons in each are similar and validate the

assumption of near-constant local sedimentation rate.

The uppermost deep-calcic paleosol at Kalbarri is near

the paleomagnetic reversal at the Hirnantian–Llando-

verian boundary (Morrison & Ellwood 1986; Rindsberg

& Chowns 1986; Schmidt & Hamilton 1990), and the basal

Indulkana Shale at 685.5 m in Ungoolya 1 core is

correlated with the early Llanvirnian (ca 468 Ma)

marine transgression of the trilobite-bearing Horn

Valley Sandstone (Laurie 2006).

This pedostratigraphic study confirms a Late Cam-

brian age inferred for the lower Tumblagooda Sand-

stone on the basis of conodonts in the Wandagee 1 core

(Gorter et al. 1994). A Cambrian–Ordovician rock unit

below the Tumblagooda Sandstone has been suggested

for a high-amplitude reflection interpreted as an un-

conformity in seismic lines of the offshore southern

Carnarvon Basin (Iasky et al. 2002). The dip of 98 noted

in the lower part of the ‘Tumblagooda Sandstone’ near

Hardabut Pool declines to flat-lying by the stream

gauging station at 400 m in the section of Hocking

(1991). An angular unconformity was not found in this

area of poor exposure, but this lower part of the section

could be the sub-Tumblagooda interval inferred from

offshore seismic lines. This is stratigraphically below

the lowest of the deep-calcic paleosols, which is in the

lower part of the flat-lying Tumblagooda Sandstone.

This lowest deep-calcic paleosol falls near the Cam-

brian–Ordovician boundary, and others are near the top

of the Arenigian, Llanvirnian, Caradocian and Ashgil-

lian (Figure 8). Of the Avalonian Ordovician stages, only

the Tremadocian–Arenigian boundary was not marked

by deep-calcic paleosols.

The pedostratigraphic age model (Figure 8) now

allows dating of the various terrestrial trace fossils

and body fossils described by Trewin & McNamara

(1994). The lowest large arthropod trackway (Diplich-

nites gouldi form B) was seen at a stratigraphic level of

451 m, equivalent to 472 Ma in the age model, or late

Arenigian (of Gradstein et al. 2004). The oldest deep

burrows in paleosols (Heimdallia chatwini, Daedalus

Figure 7 Distribution of Cam-

brian–Silurian Tumblagooda

Sandstone in Kalbarri National

Park, Western Australia (after

Hocking 1991), showing localities

examined to construct a compo-

site section.

Early Paleozoic pedostratigraphy 579

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

sp. indet.) at a level of 446 m, were also 472 Ma, and

late Arenigian. McNamara (in Hocking & Mory 2006)

used the trackways as a dating constraint, assuming

they were from eurypterids known to be no older

than Arenigian. However, the track makers could also

have been euthycarcinoids, arthropleurids, or xipho-

surs, as well as eurypterids (Trewin & McNamara

1994). The euthycarcinoid body fossil Kalbarria

brimmellae was found near The Loop plateau level

(McNamara & Trewin 1993), at a stratigraphic level

of 851 m, or ca 454 Ma (mid-Caradocian: Gradstein

et al. 2004) (Figure 8). The Skolithos–Diplocraterion

assemblage of Red Bluff and Pencell Pool records a

marine transgression, though not with definitive

marine body fossils (Trewin & McNamara 1994;

Hocking & Mory 2006) at a stratigraphic level 10 m

above the Gabba Gabba Member conglomerate marker

beds at 1125 m, or 442 Ma in the age model (Figure 8),

just above the base of the Silurian (443.7 Ma: Gradstein

et al. 2004).

SILURIAN OF GRAMPIAN RANGES

The Grampian Ranges are rugged ridges of quartzite

rising 700 m above the cleared Wimmera Tableland, in

western Victoria (Figure 9). Grampians Group quart-

zites are poorly fossiliferous (Turner 1986; Warren

et al. 1986; Burrow & Turner 2000; Gouramanis et al.

2003), and were deposited largely by rivers, although

some intervals show intertidal facies and fossils

(Spencer-Jones 1965; George 1994; Gouramanis et al.

2003). The quartzites are structurally complex, with

numerous thrust faults and local granitic and rhyolitic

intrusions deformed during Benambran mountain-

building of the Lachlan Fold Belt (Cayley & Taylor

1997).

Published age constraints

Fish denticles (Turinia fuscina), spines (Sinacanthus?

sp.) and scales (Poracanthus sp. cf. P. qujingensis)

from the Silverband Formation (2730 m in Figure 10)

are Silurian (Ludlovian: Burrow & Turner 2000) or ca

419 Ma (Gradstein et al. 2004), not Devonian or

Carboniferous as previously thought (Spencer-Jones

1965). The Rocklands Rhyolite with a U–Pb zircon age

of 410+ 3 (VandenBerg 2003) supported a Devonian

age when considered sedimentary basement, but then

the rhyolite was found to have intruded the sequence

(Simpson & Woodfull 1994), supporting a Silurian age

of the fossil fish. This rhyolite and three granodior-

ites intruding the sedimentary sequence average

400+ 13.6 Ma, whereas nine granitoids pre-dating

the sedimentary sequence average 474+ 28 Ma

(VandenBerg 2003), which is roughly Ordovician to

Devonian (Emsian to Arenigian inclusive: Gradstein

et al. 2004).

Figure 8 Age model and pedostratigraphic correlation of the Tumblagooda Sandstone of Western Australia, and equivalents

in the Officer Basin of South Australia. Metre scales differ according to local sediment accumulation rates, but proportional

distribution of deep-calcic events is preserved in each, and shows highly significant correlation.

580 G. J. Retallack

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

Figure 9 Geological map and paleosol sites examined in Grampians National Park, Victoria (after Cayley & Taylor 1997;

VandenBerg 2003).

Early Paleozoic pedostratigraphy 581

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

Burial alteration

Quartzites of the Grampians have been overthickened to

some 6250 m (Spencer-Jones 1965), due to overthrust

repetition of an original thickness of about 4000 m

(Cayley & Taylor 1997). Evidence of metamorphic

alteration is obvious as cataclastic bands near thrust

faults (Cayley & Taylor 1997), and as biotite–sericite

metaquartzite and quartz–diopside hornfels near grani-

tic intrusions (Spencer-Jones 1965). Pedogenic micrite

nodules are well preserved in red siltstones in outcrop

and in core, but comparable nodules in red sandstone

commonly appear strongly silicified and are weakly

reactive to acid. Many of the siltstones of the Silverband

Formation are not metamorphically recrystallised

(George 1994), yet were intruded by coarse-grained

Devonian granites. Possible comagmatic Devonian

ignimbrites include the Rocklands Volcanics. In modern

Yellowstone National Park, Wyoming, granitic magma

is indicated at depths of 5–10 km by a seismically slow

zone of low density, with high heat flow and low

magnetotelluric resistivity beneath a caldera which

erupted voluminous rhyolitic tuffs (Iyer 1988). Simi-

larly, a Pliocene (2.83+ 0.2 Ma by 40Ar/39Ar) ignimbrite

(Chegem Tuff) was erupted some 5 km above Eltjurta

granite in the Northern Causasus Mountains of Georgia

(Lipman et al. 1993). These calderas provide a model for

ignimbrites of the Rocklands Volcanic Group (Cayley &

Taylor 1997) and nearby granites of igneous affiliation in

the Grampians (Mafeking, Mackenzie River, Victoria

Valley, and Mirranatwa Granites: White & Chappell

1988). All the paleosols observed were from a topo-

graphic range of 500 m, and considering thrust inter-

leaving of sediments above granitic intrusions (Cayley

& Taylor 1997), all underwent nearly equal maximal

burial depths of 5.0+ 0.5 km, giving burial compaction

of 55.3+ 0.7%, using an algorithm of Sheldon &

Retallack (2001).

Pedostratigraphic ages

Deep-calcic paleosols are well spaced in lower and upper

parts of the Grampians sequence, but show three closely

spaced climatic perturbations during deposition of the

Ludlovian–Pridolian Silverband Formation (Burrow &

Turner 2000). Each deep-calcic episode is a reflection of

unusually high precipitation (Retallack 2005a) and is

associated with an increase in stream power demon-

strated by onset of erosive conglomeratic sandstones

after long intervals of silty or flaggy sandstones

(Figure 10). These perturbations correlate very closely

(R2¼ 0.99) with a series of isotopic and biotic crises

recognised in the Australian Silurian and Devonian

(Talent et al. 1993): lower Murray Hill Sandstone with

443.7 Ma basal Llandoverian, lower Major Mitchell

Sandstone with 436.2 Ma mid-Llandoverian Sandvika

Event, basal Silverband Formation with 427.6 Ma

earliest Wenlockian Ireviken Event, middle Silverband

Formation with 424.4 Ma mid-Wenlockian Mulde Event,

uppermost Silverband Formation with 420.2 Ma mid-

Ludlovian to Pridolian Lau Events, upper Serra Sand-

stone with 416.0 Ma basal Devonian Klonk Event, upper

Figure 10 Age model and pedostratigraphy of Grampians National Park. Proportional distribution of deep-calcic events in

section shows highly significant correlation with biotic and isotopic events in time.

582 G. J. Retallack

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

Moora Moora Sandstone with 411.2 Ma basal Pragian

Prague–Lochov Event, and upper Wartook Sandstone

with 406.2 Ma basal Emsian Zlichov Event (event dating

from Gradstein et al. 2004).

This new time framework allows dating of a variety

of significant fossils from the Grampians. Impressions of

equisetalean stems, 5 mm wide, from Chimney Pot Gap

(Douglas 1965) are in the lowest part of the Serra

Sandstone (Cayley & Taylor 1997), here regarded as

about 419 Ma (latest Ludlovian), so about the same age

of comparably (and controversially) large fossil plants

from Yea (Garratt & Rickards 1984). Sandstone casts of

Prototaxites (Douglas 1981; Douglas & Kenley 1981) on

the summit of Mt William (37.292958S, 142.801508E) are

also unusually large (2.91 m long by 0.32 m diameter) for

this genus of basidiomycete fungus (Boyce et al. 2007),

and here dated at 433 Ma or mid-Telychian (late

Llandoverian: Gradstein et al. 2004). Large terrestrial

arthropod burrows (Taenidium barrettii) from Mt

William (Gouramanis et al. 2003) are the same age. This

study also indicates a Llandoverian age for trace

fossils elsewhere in the Major Mitchell Sandstone

(Gouramanis et al. 2003): more specifically around

430 Ma for Glenisla quarry and 431 Ma for Mt Bepcha

some 170 m stratigraphically lower in the formation. Mt

Bepcha includes the terrestrial arthropod (Taenidium

barretti) and worm burrows (Daedalus), and a variety

of marginal-marine burrows (Skolithos linearis,

Thalassinoides, Rhizocorallium) and grazing trails

(Cruziana problematica). Glenisla quarry includes mar-

ginal marine burrows (Paleophycus, Skolithos linearis)

and large arthropod trackways (Diplichnites cuithensis),

at one time considered evidence of tetrapods on land

(Warren et al. 1986).

CONCLUSIONS

Transient deep-calcic paleosols of Paleozoic redbeds in

Australia record continent-wide abrupt increases in

mean annual precipitation (Figure 11), considering what

is known about this feature in modern soils (Retallack

2005a). In South Australian sections, deep-calcic hor-

izons also correspond to times of climatic warming,

marine transgression, black shale deposition and

articulated trilobite preservation (Retallack 2008). Deep-

calcic paleosols also were coeval with transient declines

in carbon isotopic composition of Australian Silurian–

Devonian limestones and extinctions (Talent et al. 1993).

Deep-calcic paleosols also appear at times of global

Cambrian–Devonian bioevents (Barnes et al. 1996; Zhu

et al. 2006; Patzkowsky & Holland 2007), including the

fourth largest known mass extinction within the latest

Ordovician (end-Ashgillian, 445.6 Ma) and seventh lar-

gest in the Late Silurian (end-Ludlovian, 418.7 Ma:

Calner 2005; Stanley 2007). The Australian paleosol

Figure 11 Composite Australian record of depth to Bk horizon in paleosols, compared with carbon isotopic record of the Great

Basin, USA (Saltzman 2005), Siberia (Kouchinsky et al. 2007), and Early Paleozoic bioevents (Gradstein et al. 1984).

Early Paleozoic pedostratigraphy 583

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

record also corresponds with global variation in marine

carbonate isotopic composition (Figure 11), best known

from long records in the North American Great Basin

(Saltzman 2005) and Siberia (Kouchinsky et al. 2007).

Marine records show not only transient negative carbon

isotopic excursions coeval with deep-calcic paleosols,

but protracted intervening intervals of very high (42 %d13CPDB) carbon isotope composition. The Late Cam-

brian (Steptoan) positive carbon isotopic interval

(SPICE of Saltzman 2005) corresponds with a protracted

interval of consistently shallow paleosol calcic depth.

Other positive isotopic excursions in the marine

carbonate record were more short-lived within the

Caradocian, Hirnantian, Wenlockian, Ludlovian and

Pridolian, and correspond with shallow calcic paleosols

of dry climates. One of these intervals, the Hirnantian,

coincides with a short-lived glaciation centered on West

Africa (Denis et al. 2007), and coarse gravelly sediments

in the Gabba Gabba Member of the Tumblagooda

Sandstone (Evans et al. 2007) may have been distal

glacial outwash. Also comparable between paleosol and

marine isotopic records is increased time intervals

between perturbations through time (Saltzman 2005),

from an unsettled Cambrian record (Figure 11). Deep-

calcic paleosol events thus reflect global atmospheric

perturbations in the carbon cycle.

What would cause such perturbations is unclear for

the Early Paleozoic, but has been widely discussed in

the case of the end-Permian carbon isotopic excursions

and mass extinctions. Permian negative carbon isotopic

excursions and deep-calcic zones associated with mass

extinctions have been attributed to bolide impact,

massive volcanism, oceanic overturn, biomass oxida-

tion, and methane emission from permafrost and

oceanic clathrates, or from thermal alteration of carbo-

naceous sediments during intrusion (Retallack 2005b;

Retallack & Krull 2006; Retallack & Jahren 2008). Marine

positive excursions on the other hand signal increased

biological productivity, due to increased oceanic fertili-

sation and ventilation, and carbon burial at sea (Saltz-

man 2005). Plate-tectonic rearrangements or volcanic

flareups could be responsible for such events (Bluth &

Kump 1991), but a more likely explanation is evolu-

tionary advancements on land which promoted chemi-

cal weathering, such as the Middle Cambrian advent of

nonvascular land plants, Late Ordovician evolution of

mosses, and Early Silurian appearance of vascular land

plants (Retallack 2000).

ACKNOWLEDGEMENTS

Barbara and Warren Fargher graciously allowed access

to Wirrealpa Station. Permission for research in Flin-

ders Ranges National Park was approved by Kate Wood,

Ken Anderson, Darren Crawford, Arthur Coulthard,

and Pauline Coulthard. Permission for research in

Kalbarri National Park was approved by Kieran

McNamara, Mike Paxman and Russell Asplin. Exam-

ination of cores was aided by Brian Logan and Michael

Willison at the Primary Industries and Resources South

Australia Core Library in Glenside, Avi Olchina and

Ken Sherring at the Geoscience Victoria Core Library in

Werribee, and Chris Brooks at the Geological Survey of

Western Australia Core Library in Carlisle. Ian John-

son, Diane Retallack and Christine Metzger helped with

fieldwork. Arthur Mory and Ken McNamara provided

helpful and detailed reviews. Funded by American

Chemical Society PRF grant 45257-AC8.

REFERENCES

BARNES C. R., FORTEY R. A. & WILLIAMS S. H. 1996. The pattern of

global bioevents during the Ordovician period. In: Walliser O. H.

ed. Global Events and Event Stratigraphy in the Phanerozoic, pp.

139–179. Springer, Berlin.

BENBOW M. C. 1982. Stratigraphy of the Cambrian–?Early Ordovi-

cian Mount Johns Range, NE Officer Basin, South Australia.

Transaction of the Royal Society of South Australia 106, 191–211.

BLUTH G. J. S. & KUMP L. R. 1991. Phanerozoic paleogeology.

American Journal of Science 291, 284–308.

BOYCE C. K., HOTTON C. L., FOGEL M. L., CODY G. D., HAZEN R. M.,

KNOLL A. H. & HUEBER F. M. 2007. Devonian landscape

heterogeneity recorded by a giant fungus. Geology 35, 399–402.

BURROW C. J. & TURNER S. 2000. Silurian vertebrates from Australia.

Courier Forschungsinstitut Senckenberg 22, 169–174.

CALNER M. 2005. A Late Silurian extinction event and anachronistic

period. Geology 4, 305–308.

CAYLEY R. A. & TAYLOR D. H. 1997. Grampians special area

geological report. Geological Survey of Victoria Report 107.

CLARKE J. D. A. 1990. An Early Cambrian carbonate platform near

Wilkawillina Gorge, South Australia. Australian Journal of

Earth Sciences 37, 471–483.

DAILY B. & FORBES B. G. 1969. Notes on the Proterozoic and

Cambrian, southern and central Flinders Ranges. In: Daily B.

ed. Geological Excursions Handbook, pp. 23–30. ANZAAS Con-

gress, Section 3, Adelaide.

DALGARNO C. R. 1964. Lower Cambrian stratigraphy of the Flinders

Ranges. Transactions of the Royal Society of South Australia 88,

129–144.

DALGARNO C. R. & JOHNSON J. E. 1966. Parachilna Sheet SH54-13,

1:2500 Geological Atlas Sheet. Geological Survey of South

Australia, Adelaide.

DALRYMPLE G. B. 1979. Critical tables for conversion of K–Ar ages

from old to new constants. Geology 7, 558–560.

DEMKO T. M., CURRIE B. S. & NICOLL K. A. 2004. Regional

paleoclimatic and stratigraphic implications of paleosols and

fluvial/overbank architecture in the Morrison Formation

(Upper Jurassic), Western Interior, USA. Sedimentary Geology

167, 115–135.

DENIS M., BUONCRISTIANI J-F., KONATE M. & GUIRAUD M. 2007. The

origin and glaciodynamic significance of sandstone ridge net-

works from the Hirnantian glaciation of the Djado Basin (Niger).

Sedimentology 54, 1225–1243.

DIESSEL C. F. K. 1992. Coal-Bearing Depositional Systems. Springer,

Berlin.

DOUGLAS J. G. 1965. Articulated stems from the Grampians. In:

Spencer-Jones D. ed. The Geology and Structure of the Grampians

Area, Western Victoria, p. 92. Geological Survey of Victoria

Memoir 25.

DOUGLAS J. G. 1981. Fossil algae, Victoria, Australia. Palaeobotanist

28/29, 8–14.

DOUGLAS J. G. & KENLEY P. R. 1981. Enigmatic organic structures in

the Grampians Group sediments. Victorian Naturalist 98, 65–67.

DRIESE S. G. & FOREMAN J. L. 1992. Paleopedology and paleoclimatic

implications of Late Ordovician vertic paleosols southern

Appalachians. Journal of Sedimentary Petrology 62, 71–83.

DRIESE S. G., MORA C. I. & ELICK J. M. 1997. Morphology and

taphonomy of root and stump casts of the earliest trees (Middle

to Late Devonian), Pennsylvania and New York. Palaios 12, 524–

537.

DRIESE S. G., MORA C. I. & ELICK J. M. 2000. The paleosol record of

increasing plant diversity and depth of rooting and changes in

atmospheric pCO2 in the Siluro-Devonian. In: Gastaldo R. A. &

Dimichele W. A. eds. Phanerozoic Terrestrial Ecosystems, pp. 47–

61. Paleontological Society Special Paper 6.

584 G. J. Retallack

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

ELBURG M. A., BONS P. D., FODEN J. & BRUGGER J. 2003. A newly

defined Late Ordovician magmatic–thermal event in the Mt

Painter Province, northern Flinders Ranges, South Australia.

Australian Journal of Earth Sciences 50, 611–631.

EVANS R., MORY A. J. & TAIT A. M. 2007. An outcrop gamma ray

study of the Tumblagooda Sandstone, Western Australia.

Journal of Petroleum Science and Engineering 57, 37–59.

FODEN J., ELBURG M. A., DOUGHERTY-PAGE J. & 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.

FOLLMER L. R. 1978. The Sangamon in its type area—a review. In:

Mahaney W. C. ed. Quaternary Soils, pp. 125–165. GeoAbstracts,

Norwich.

GARRATT M. J. & RICKARDS R. B. 1984. Graptolite biostratigraphy of

early land plants from Victoria, Australia. Proceedings of the

Yorkshire Geological Society 44, 377–384.

GATEHOUSE C. G. 1976. A fossil in the Observatory Hill Beds, South

Australia. Geological Survey of South Australia Quarterly Notes

60, 5–8.

GEORGE A. D. 1994. Tidal sedimentation in part of the Late Silurian

Grampians Basin, southeastern Australia. Journal of Sedimen-

tary Research B64, 311–325.

GEYER G. & SHERGOLD J. 2000. The quest for internationally

recognized divisions of Cambrian time. Episodes 23, 188–195.

GILE L. H., HAWLEY J. W. & GROSSMAN R. B. 1980. Soils and

geomorphology in the Basin and Range of southern New Mexico:

Guidebook to the desert project. New Mexico Bureau of Mines &

Mineral Resources Memoir 39.

GILE L. H., PETERSON F. F. & GROSSMAN R. B. 1966. Morphological

and genetic sequences of carbonate accumulation in desert soils.

Soil Science 101, 341–360.

GORTER J. D., NICOLL R. S. & FOSTER C. B. 1994. Lower Palaeozoic

facies in the Carnarvon Basin, Western Australia: stratigraphy

and hydrocarbon prospectivity. In: Purcell P. G. & Purcell R. R.

eds. The Sedimentary Basins of Western Australia, pp. 373–396.

Petroleum Exploration Society of Australia, Perth.

GOURAMANIS C., WEBB J. A. & WARREN A. A. 2003. Fluviodeltaic

sedimentology and ichnology of part of the Silurian Grampians

Group, western Victoria. Australian Journal of Earth Sciences

50, 811–825.

GRADSTEIN F. M., OGG J. G. & SMITH A. G. 2004. A Geologic Time Scale

2004. Cambridge University Press, Cambridge.

GRAVESTOCK D. I. & SHERGOLD J. H. 2001. Australian Early and

Middle Cambrian sequence biostratigraphy with implications

for species diversity and correlation. In: Zhuravlev A. Y. &

Riding R. eds. The Ecology of the Cambrian Radiation, pp. 107–

136. Columbia University Press, New York.

HAINES P. W. & FLOTTMANN T. 1998. Delamerian orogeny and

potential foreland sedimentation: a review of age and strati-

graphic constraints. Australian Journal of Earth Sciences 45,

559–570.

HOCKING R. M. 1990. Carnarvon Basin. In: Geology and Mineral

Resources of Western Australia, pp. 457–494. Geological Survey of

Western Australia Memoir 3.

HOCKING R. M. 1991. The Silurian Tumblagooda Sandstone, Western

Australia. Geological Survey of Western Australia Report 27.

HOCKING R. M. & MORY A. J. 2006. Geology of the Kalbarri area—a field

guide. Geological Survey of Western Australia Record 2006/19.

IASKY R. P., D’ERCOLE C., GHARI K. A. R., MORY A. J. & LOCKWOOD A.

2002. Structure and petroleum prospectivity of the Gascoyne

Platform, Western Australia. Geological Survey of Western

Australia Report 87.

IASKY R. P. & MORY A. J. 1999. Geology and petroleum potential of

the Gascoyne Platform, southern Carnarvon Basin, Western

Australia. Geological Survey of Western Australia Report 69.

IYER H. M. 1988. Seismological detection and delineation of magma

chambers beneath intraplate volcanic centers in western USA.

In: King C-Y. & Scarpa R. eds. Modeling of volcanic processes, pp.

1–56. Friedrick Vieweg & Sohn, Braunschweig.

JAGO J. B., LIN T. R. & DUNSTER J. N. 2002. A new species of the

trilobite Abadiella from the Lower Cambrian of the eastern

Officer Basin, South Australia. Acta Palaeontologica Sinica 41,

428–433.

JAGO J. B. & YOUNGS B. C. 1980. Early Cambrian trilobites from the

Officer Basin, South Australia. Transactions of the Royal Society

of South Australia 104, 197–199.

JAGO J. B., ZANG W. L., SUN X. L., BROCK G. A., PATERSON J. R. &

SKOVSTED C. B. 2006. A review of the Cambrian biostratigraphy

of South Australia. Palaeoworld 15, 406–423.

KIRSCHVINK J. L. & RAUB T. D. 2003. A methane fuse for the Cambrian

explosion: carbon cycles and true polar wander. Geoscience 335,

65–78.

KOUCHINSKY A., BENGTSON S., PAVLOV V., RUNNEGAR B., TORSSANDER

P., YOUNG E. & ZIEGLER K. 2007. Carbon isotope stratigraphy of

the Precambrian–Cambrian Sukharikha River section, north-

western Siberian Platform. Geological Magazine 144, 609–618.

LAURIE J. R. 2006. Ordovician trilobites from the Horn Valley

Siltstone and basal Stairway Sandstone, Amadeus Basin, North-

ern Territory. In: Paterson J. R. & Laurie J. R. eds. Cambro-

Ordovician studies II, pp. 287–345. Association of Australasian

Palaeontologists Memoir 32.

LINDSAY F. F. & LEVEN J. A. 1996. Evolution of a Neoproterozoic to

Palaeozoic intracratonic setting, Officer Basin, South Australia.

Basin Research 8, 403–424.

LIPMAN P. W., GOGATIKOV O. A., TSVETKOV A. A., GAZIS C.,

GURBANOV A. G., HON K., KORONOVSKY N. V., KOVALENKO V. I.

& MARCHEV P. 1993. 2.8-Ma ash-flow caldera at Chegem River in

northern Caucasus Mountains (Russia), contemporaneous gran-

ites, and associated ore deposits. Journal of Volcanology and

Geothermal Research 57, 85–124.

MAJOR R. B. & TELUK J. A. 1967. The Kulyong Volcanics. Geological

Survey of South Australia Quarterly Geology Notes 22, 8–11.

MAWSON D. 1938. Cambrian and sub-Cambrian formations at

Parachilna Gorge. Transactions of the Royal Society of South

Australia 62, 255–262.

MAWSON D. 1939a. The Cambrian sequence in the Wirrealpa Basin.

Transactions of the Royal Society of South Australia 63, 331–347.

MAWSON D. 1939b. The late Proterozoic sediments of South

Australia. ANZAAS Meeting Report 24(1), 79–88.

MAWSON D. & SEGNIT E. R. 1949. Purple slates of the Adelaide

System. Transactions of the Royal Society of South Australia 72,

276–280.

MCNAMARA K. J. & TREWIN N. H. 1993. A euthycarcinoid arthropod

from the Silurian of Western Australia. Palaeontology 36, 319–335.

MITCHELL M. M., KOHN B. P. & FOSTER D. A. 1998. Post-orogenic

cooling history of eastern South Australia from apatite FT

thermochronology. In: Van den Haute P. & de Corte F. eds.

Advances in fission-track geochronology, pp. 207–224. Kluwer,

Amsterdam.

MITCHELL M. M., KOHN B. P., O’SULLIVAN P. B., HARTLEY M. J. &

FOSTER D. A. 2002. Low-temperature thermochronology of the Mt

Painter Province, South Australia. Australian Journal of Earth

Sciences 49, 551–563.

MOORE P. S. 1979. Stratigraphy and depositional environments of the

Billy Creek Formation (Cambrian) central and northern Flin-

ders Ranges, South Australia. Transactions of the Royal Society of

South Australia 103, 197–211.

MOORE P. S. 1990. Origin of redbeds and variegated sediments,

Cambrian, Adelaide Geosyncline, South Australia. In: Jago J. B.

& Moore P. S. eds. The Evolution of a Late Precambrian–Early

Palaeozoic Rift Complex: the Adelaide Geosyncline, pp. 334–350.

Geological Society of Australia Special Publication 16.

MORRISON J. & ELLWOOD B. B. 1986. Paleomagnetism of Silurian-

Ordovician sediments from the Valley and Ridge Province,

northwest Georgia. Geophysical Research Letters 13, 189–192.

NORTH AMERICAN COMMISSION ON STRATIGRAPHIC NOMENCLATURE

1982. North American stratigraphic code. AAPG Bulletin 67, 841–

875.

ORTLAM D. 1971. Paleosols and their significance in stratigraphy and

applied geology in the Permian and Triassic of southern Germany.

In: Yaalon D. H. ed. Paleosols: Origin, Nature and Dating of

Paleosols, pp. 321–327. Israel Universities Press, Jerusalem.

OWEN G. 2003. Load structures: gravity-driven sediment mobilization

in the shallow subsurface. In: Van Rensbergen P., Hillis R. R.,

Maltman A. J. & Morley C. K. eds. Subsurface Sediment

Mobilization, pp. 21–34. Geological Society of London Special

Publication 216.

PARKIN L. W. 1969. Handbook of South Australian Geology. Geologi-

cal Survey of South Australia, Adelaide.

PATERSON J. R. & BROCK G. A. 2007. Early Cambrian trilobites from

Angorichna, Flinders Ranges, South Australia, with a new

assemblage from the Pararaia bunyerooensis zone. Journal of

Paleontology 81, 116–142.

Early Paleozoic pedostratigraphy 585

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009

PATZKOWSKY M. E. & HOLLAND S. M. 2007. Diversity partitioning of a

Late Ordovician marine biotic invasion: controls on diversity in

a regional ecosystem. Paleobiology 33, 295–309.

PHILIP G. M. 1969. Silurian conodonts from the Dirk Hartog

Formation, Western Australia. Proceedings of the Royal Society

of Victoria 82, 287–297.

RETALLACK G. J. 1992. What to call early plant formations on land.

Palaios 7, 508–520.

RETALLACK G. J. 2000. Ordovician life on land and early Paleozoic

global change. In: Gastaldo R. A. & Dimichele W. A. eds.

Phanerozoic terrestrial ecosystems, pp. 21–45. Paleontological

Society Special Paper 6.

RETALLACK G. J. 2001. Soils of the Past. Blackwell, Oxford.

RETALLACK G. J. 2005a. Pedogenic carbonate proxies for amount and

seasonality of precipitation in paleosols. Geology 33, 333–336.

RETALLACK G. J. 2005b. Permian greenhouse crises. In: Lucas S. G. &

Ziegler K. E. eds. The Nonmarine Permian, pp. 256–269. Bulletin

New Mexico Museum of Natural History and Science 30.

RETALLACK G. J. 2008. Cambrian paleosols and landscapes of South

Australia. Australian Journal of Earth Sciences 55, 1083–1106.

RETALLACK G. J. Greenhouse crises of the geological past. Geological

Society of America Bulletin (in press).

RETALLACK G. J. & JAHREN A. H. 2008. Methane release from igneous

intrusion of coal during Late Permian extinction events. Journal

of Geology 116, 1–20.

RETALLACK G. J. & KRULL E. S. 2006. Carbon isotopic evidence for

terminal-Permian methane outbursts and their role in extinc-

tions of animals, plants, coral reefs and peat swamps. In: Greb S.

& DiMichele W. A. eds. Wetlands Through Time, pp. 249–268.

Geological Society of America Special Paper 399.

RICHARDS J. R., BLOCKLEY J. G. & DE LAETER J. R. 1985. Rb–Sr and Pb

isotopic data from the Northampton Block, Western Australia.

Australasian Institute of Mining and Metallurgy Proceedings 290,

43–55.

RINDSBERG A. K. & CHOWNS T. M. 1986. Ringgold Gap: progradational

sequences in the Ordovician and Silurian of northwest Georgia.

In: Neathery T. L. ed. Centennial Field Guide, Southeast Section,

vol. 6, pp. 159–162. Geological Society of America, Boulder.

SALTZMAN M. R. 2005. Phosphorus, nitrogen, and the redox evolution

of the Paleozoic oceans. Geology 33, 573–576.

SCHMIDT P. W. & HAMILTON P. J. 1990. Palaeomagnetism and age of

the Tumblagooda Sandstone, Western Australia. Australian

Journal of Earth Sciences 37, 381–385.

SHELDON N. D. & RETALLACK G. J. 2001. Equation for compaction of

paleosols due to burial. Geology 29, 247–250.

SIMPSON C. J. & WOODFULL C. J. 1994. New field evidence resolving

the relationship between the Grampians Group and Rocklands

Rhyolite, western Victoria. Australian Journal of Earth Sciences

41, 621–624.

SOIL SURVEY STAFF 2000. Keys to Soil Taxonomy. Pocahontas Press,

Blacksburg.

SPENCER-JONES D. 1965. The geology and structure of the Grampians

area, western Victoria. Geological Survey of Victoria Memoir 25.

STANLEY S. M. 2007. An analysis of the history of marine animal

diversity. Paleobiology Memoir 33(4), 1–55.

STRATIGRAPHIC NOMENCLATURE COMMITTEE.1973. Australian code of

stratigraphic practice. Journal of the Geological Society of

Australia 20, 105–112.

STEVENS M. K. & APAK S. N. 1999. Empress 1 and 1A well completion

report, Yowalga Sub-basin, Officer basin, Western Australia.

Geological Survey of Western Australia Record 1999/4.

STOCK E. C. 1974. The clay mineralogy, petrology and environments

of deposition of the Cambrian Lake Frome Group, Flinders

Ranges, South Australia. MSc thesis, University of Adelaide,

Adelaide (unpubl.).

TALENT J. A, MAWSON R., ANDREW A. S., HAMILTON P. J. & WHITFORD

D. J. 1993. Middle Palaeozoic extinction events: faunal and

isotopic data. Palaeogeography Palaeoclimatology Palaeoecology

104, 139–152.

TINGATE P. R. & DUDDY I. R. 2002. The thermal history of the eastern

Officer Basin (South Australia): evidence from apatite fission

track analysis and organic maturity data. Tectonophysics 349,

251–275.

TISSOT B. P. & WELTE D. H. 1984. Petroleum Formation and

Occurrence. Springer, Berlin.

TREWIN N. H. & MCNAMARA K. J. 1994. Arthropods invade the land:

trace fossils and paleoenvironments of the Tumblagooda Sand-

stone (?late Silurian) of Kalbarri, Western Australia. Transac-

tions of the Royal Society of Edinburgh: Earth Sciences 85, 177–

210.

TUCKER M. D. 1991. Carbon isotopes and Cambrian–Precambrian

boundary geology South Australia: ocean basin formation,

seawater chemistry and organic evolution. Terra Nova 1, 573–

582.

TURNER S. 1986. Vertebrate fauna of the Silverband Formation,

Grampians, western Victoria. Proceedings of the Royal Society of

Victoria 98, 53–62.

VANDENBERG A. H. M. 2003. Silurian to Early Devonian. In: Birch W.

D. ed. Geology of Victoria, pp. 117–155. Geological Society of

Australia Special Publication 23.

VICTORIA PETROLEUM N. L. 2004. Final well report Wendy 1

stratigraphic drilling project EP 427. Geological Society of

Western Australia Statutory Petroleum Exploration Report

S20948A1 (unpubl.).

WARREN A., JUPP R. & BOLTON B. 1986. Earliest tetrapod trackway.

Alcheringa 10, 183–186.

WEBB A. W. 1978. Geochronology of stratigraphically significant

rocks. South Australia Department of Mines and Energy Amdel

Progress Report 24.

WHEELER R. L. 2002. Distinguishing seismic from no-seismic soft-

sediment structures: criteria from seismic-hazard analysis. In:

Ettensohn F. R., Rast N. & Brett C. E. eds. Ancient Seismites, pp.

1–11. Geological Society America Special Paper 359.

WHITE A. J. R. & CHAPPELL B. W. 1988. Granites. In: Douglas J. G. &

Ferguson J. A. eds. Geology of Victoria, pp. 427–439. Geological

Society of Australia, Victorian Division, Melbourne.

WOPFNER H. 1970. Early Cambrian paleogeography, Frome Embay-

ment, South Australia. AAPG Bulletin 54, 2395–2409.

YASIN A. R. & MORY A. J. 1999. Coburn well completion report,

Gascoyne Platform, southern Carnarvon Basin Western Austra-

lia. Geological Survey of Western Australia Record 1999/5.

ZANG W-L. 2002, Sequence analysis and petroleum potential in the

Arrowie Basin, South Australia. Minerals and Energy Resources

South Australia Report Book 2002/024.

ZHU M-Y., BABCOCK L. E. & PENG S-C. 2006. Advances in Cambrian

stratigraphy and paleontology: integrating correlation techni-

ques, paleobiology, taphonomy and paleoenvironmental recon-

struction. Paleoworld 2, 217–222.

Received 4 March 2008; accepted 15 January 2009

586 G. J. Retallack

Downloaded By: [Retallack, G. J.][University of Oregon] At: 17:11 2 June 2009