Vous êtes sur la page 1sur 26

Sedimentology (2001) 48, 559584

Palaeogeographical inuence on Late Eocene biosiliceous


sponge-rich sedimentation, southern Western Australia
P. R. GAMMON 1 . and N. P. JAMES
Department of Geological Sciences, Queen's University, Kingston, Ontario K7L 3N6, Canada
ABSTRACT

Late Eocene nearshore shallow-marine environments within the Bremer and


western Eucla Basins of southern Western Australia were characterized by the
thick deposition of spongolite and spiculite deposits. Epibenthic sponge
communities dominated estuaries and topographically complex basin margin
embayments-archipelagos, while cool-water carbonates with up to 10% sponges
accumulated in open-shelf environments. The transition from a biosiliceous
to calcareous epibenthos was related to the degree of palaeogeographical
`protection'. Within basement-protected embayments there was an offshore
gradation from shoreface spongolite and pure spiculite to a muddy spiculite
facies towards central embayment areas. Calcareous fossils are rare throughout
embayment facies, but rapidly increase in more open outer archipelago areas.
This depositional relationship occurred along 2000 km of the Late Eocene
southern Australian coastline. Palaeogeographical protection from strong
currents acted in concert with: (1) a planar, low-gradient inland topography
with sluggish run-off, supplying ne-grained sediment, nutrients, and abundant
dissolved silica; and (2) a microtidal setting, weak to moderate swells and
opposing wind and Coriolis surface current forcing, which inhibited water
exchange between embayments-estuaries and the open shelf. This situation led
to an embayment water chemistry that encouraged prolic sponge growth.
Calcareous spiculites record the mixing front between these embayment waters
and normal open-shelf waters supporting cool-water carbonates.
Keywords Australia, Eocene, palaeogeography, sedimentary facies, siliceous
sponge.

INTRODUCTION
Sponge-rich facies within the geological record
are usually interpreted as polar and/or deep
water deposits, in part because of the prolic
sponge communities in polar regions today, and
in part due to the latitudinal position of ancient
sponge-rich deposits (Reid, 1967, 1968; Koltun,
1970; Dayton et al., 1974; Krautter, 1987; van
Soest & Stentoft, 1988; van Wagoner et al., 1989;
Conway et al., 1991; Lang, 1991; Levi, 1991;
1

Present address: Department of Geology and Geophysics, Adelaide University, Adelaide 5005, Australia
(E-mail: paul.gammon@adelaide.edu.au).

2001 International Association of Sedimentologists

Pisera, 1991, 1997a; Trammer, 1991; Henrich


et al., 1992; Wiedenmayer, 1994; Beauchamp &
Desrochers, 1997). However, many siliceous
demosponges today live in warm, shallow waters
(Dayton et al., 1974; Reiswig, 1990; Rutzler,
1990; Reitner & Keupp, 1991; Ilan & Abelson,
1995), suggesting that at some stage(s) in the
ancient past, environmental conditions facilitating profuse warm, shallow-marine sponge
colonies may have occurred. Reports of such
shallow marine sponge-dominated sediments are
few, and mostly conned to localized, thin
horizons (e.g. Cavaroc & Ferm, 1968; McBride &
Folk, 1977; Lane, 1981). The depositional
environments and sedimentology of such
559

560

P. R. Gammon and N. P. James

shallow-marine sponge-dominated deposits are,


therefore, still poorly documented.
Biosiliceous sediments rich in sponge spicules and sponge body fossils were deposited
along the northern edge of a large east-west
oceanic gulf between Australia and Antarctica
during the Late Eocene (Fig. 1; Hinde, 1910;
Clarke & Phillipps, 1952; de Laubenfels, 1953;
Cockbain, 1968a; Pickett, 1982; Jones & Fitzgerald, 1986; McGowran & Beecroft, 1986;
McGowran, 1989a; Clarke, 1993; Benbow et al.,
1995; James & Bone, 2000). These deposits are
particularly well exposed in the onshore Bremer
and western Eucla Basins along the southern
coast of Western Australia. The purpose of this
paper is to document, interpret and model the

Fig. 1. Palaeogeographical reconstruction of Australia


and Antarctica at 40 Ma (redrawn from Lawver et al.,
1992). Shaded areas denote the Late Eocene southern
Australian basins with spiculitic sediments rimming
the margins in dark shade, and calcareous and terrigenous sediments in light shade. Symbols: STR, South
Tasman Rise continental block; BB, Bremer Basin; EB,
Eucla Basin; SVB, St. Vincents Basin; MB, Murray
Basin; GB, Gippsland, Otway and Bass Basins (basins
redrawn after Veevers, 1984). Square denotes study
area of the western Eucla Basin and Bremer Basin and
is enlarged in Fig. 2. Dashed lines are continental shelf
edge, dotted lines are latitudinal lines. Cities denoted
by black circles, with abbreviations: S, Sydney; M,
Melbourne; K, Kalgoorlie; E, Esperance; A, Albany; P,
Perth. Arrows indicate predominantly westerly winds
at this time (Kemp, 1978; Veevers, 1984); and Coriolis
force currents, discussed in text.

sedimentary facies and depositional environments of the Western Australian sponge-rich


deposits.
METHODS AND TERMINOLOGY
Mapping of the lithofacies and biofacies distribution, and measurement of stratigraphical sections,
was accomplished during the Australian 199798
summer. Pallinup Formation sections occur along
the southern Western Australian coastal plain,
with the bulk of high-quality sections located in
Fitzgerald River National Park (Fig. 2). The nearest Wilson Bluff Limestone outcrop occurs at the
westernmost end of cliffs that rim the Great
Australian Bight (Figs 1 and 2). Seventy-four thin
sections were analysed to aid facies interpretations. All percentages of lithology components
were visually estimated, and are summarized in
Tables 1 and 2.
The specic gravity (SG) of sponge opal-A was
determined using the weight of dry sponge
spicules, and volumes determined from the
weight difference, within a constant volume,
between dry sponge spicules plus water, and
water alone. Sponge SG calculations were performed on 30 lithistid sponge fragments. Lithistid
sponge fragments were weighed dry (without
pore water) and wet (gently ultrasonied to
ensure complete water penetration into pores).
The weight difference yields the sponge pore
water weight, which was converted to sponge
pore volume using the SG of water at 20 C
(empirical constants used were from Weast et al.,
1972). The dry fragment weight was converted to
spicule volume using the SG of opal-A. To
calculate the SG of a sponge fragment, the wet
weight was divided by the combined spicule and
pore volumes. Freshwater SG was converted to
sea water SG prior to calculating sponge SG.
The term spiculite is used to indicate
sediments dominated by sponge spicules, and
spongolite to indicate sediments dominated by
rigid-bodied sponge skeletons.
GEOLOGICAL SETTING
The contiguous Bremer and Eucla basins developed during AustraliaAntarctic rifting in the
Late Mesozoic and Early Tertiary (Figs 1 and 2;
Veevers, 1984; Feary et al., 1994). Both were lled
with Eocene sediments. Subsequently, Oligocene
uplift and southerly tilting of the Yilgarn Craton

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation

561

Fig. 2. Lithofacies map illustrating distribution of Fitzgerald Member facies along the southern Western Australian
margin (Bremer and Eucla Basins). Present-day contours (dashed lines) turn from basin parallel and low-gradient in
the east (Eucla Basin), to coast parallel and high-gradient in the west (Bremer Basin). This reects the post-Eocene
subsidence of the Eucla Basin (Feary & James, 1998), and the Late Eocene palaeo-escarpment along the northern edge
of the Bremer Basin (discussed in text). The palaeo-escarpment is due to Late Eocene marine erosion and is inferred
to be the Late Eocene transgressive limit. Upper Eocene sediments inland from the palaeo-escarpment are contained
within palaeovalleys (e.g. Princess Royal Spongolite; Clarke, 1993). Stratigraphic units: PRS, Princess Royal
Spongolite; PF, Pallinup Formation. Black dots indicate the more important outcrops used during mapping.
Annotated outcrops that are mentioned in the text are: NRP, North Royal Pit; MH, Munglinup High; TR, Thomas
River; FRNP, Fitzgerald River National Park. N is Australian Map Grid North redrawn and simplied after Western
Australian 1:250 000 geology and topography map sheets: Albany-Mt. Barker, Bremer Bay, Newdegate, Ravensthorpe, Esperance-Mondrain Island, Kalgoorlie and Norseman.

has kept the Bremer Basin succession subaerially


exposed (Cope, 1975). In the Eucla Basin, Oligocene uplift was followed by renewed subsidence
that resulted in deposition of Oligocene and
Miocene neritic carbonates (Lowry, 1970; Veevers, 1984; Feary et al., 1994).

Eocene sediments and their diagenesis


Tertiary deposits within the onland Bremer Basin
are encompassed within the Plantagenet Group,
which contains two formations, the mid-Upper
Eocene Werrilup Formation and the Upper
Eocene Pallinup Formation (Fig. 3; Clarke & Phillipps, 1952, 1955; Cockbain, 1968a,b; Darragh &
Kendrick, 1980). These two formations represent

two major transgressiveregressive cycles (Cockbain, 1968a; Clarke, 1993; Gammon et al., 2000a),
and are inferred to be equivalent to the widespread Middle to Late Eocene Tortachilla and
Tjuketja transgressiveregressive cycles of southern Australia (for more details see McGowran
et al., 1997; Gammon et al., 2000a). The earlier
Werrilup Formation deposits are heterolithic
shallowto
marginal-marine
siliciclastic
mudstones, siltstones, sandstones, lignites and
bryozoan limestones containing warm-water calcareous algae (Nanarup Limestone Member;
Cockbain, 1968a; Quilty, 1969). The younger
Pallinup Formation, which disconformably overlies the Werrilup Formation, contains siliciclastic
mudstones and sandstones with abundant

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Moderate to poorly
sorted mud and silt

Well-sorted spicules,
minor mud and silt

Poorly to well-sorted
pebble sized lithistids

Moderate to poorly
sorted mud and silt

Poorly to well-sorted
mud and minor silt

SP1

SP2

SP3

SP4

LM1 and
LM2

070%, detrital &


carbonate mud

2040%, silicied,?
ex-detrital

<2%, possibly
weathering introduced

<5%, detrital

1050%, detrital &


authigenic

>90%, authigenic

4070%, detrital &


authigenic

1060%, detrital &


authigenic

0%, possibly altered


during cement
diagenesis

Clay

<5%, pristine, angular

1020%, pristine,
angular

<2%, pristine, angular

<5%, pristine, angular

1020%, pristine,
angular

<5%, angular, pristine

1020%, angular,
pristine

1020%, pristine,
angular

<5%, rounded

Silt

0%

pristine 510%, very


ne grained, angular,
pristine

<2%, ne to coarse
grained, angular,

<2%, very ne grained,


angular, pristine

<2%, very ne grained,


angular, pristine

0%

<2%, very ne grained,


angular, pristine

1080%, angular, pristine,


moderate sorting

6080%, rounded,
polycrystalline, minor
intraclasts and pebbles

Sand

<2%, disseminated
groundmass alteration

~5%, moulds of peloids, angular


fragments & disseminated
alteration

<2%, occasional green-clay lined


burrows; probably leached

<2%, few peloids, mostly


angular fragments

510%, ovoid peloids & angular


fragments

>90%, ovoid peloids, matrix


& fossil pseudomorphs

525%, ovoid peloids, angular


fragments, fossil replacement

510%, ovoid peloids &


irregular fragments

2030%, authigenic
cement & ovoid peloids

Verdine

Pristine, monocrystalline, unstained grains that look fresh in both hand specimen and thin section. Terrigenous clastic silt and sand are quartz and muscovite,
except for TC1, which is only quartz. Verdine peloids and fragments are very ne to coarse sand sized.

Green-clay with sandsized verdine peloids

TC4

Moderate to well-sorted
muddy very ne sands

TC2

Moderate to poorly sorted


mud and silt

Well-sorted ne to
coarse sand and

TC1

TC3

Texture

Facies

Table 1. Facies composition and texture.

562
P. R. Gammon and N. P. James

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Spicule: 0%
Sponge: 0%

Spicule: 210%, whole & unabraded


Sponge: <2%, whole to large
unabraded fragments

Spicule: 540% whole & unabraded


Sponge: <5% Whole to large
unabraded fragments

Spicule: 0%
Sponge: <2%, verdinized fragments

Spicule: 1040%, whole & unabraded


Sponge: 5%, whole to large
unabraded fragments

Spicule: 5090%, whole-unabraded


to fragmented
Sponge: <550%, whole to unabraded
fragments

Spicule: 050%, whole-unabraded to


fragmented-abraded
Sponge: 50100%, whole-unabraded
to comminuted

Spicule: 3050%, moulds, whole


& unabraded
Sponge: 510%, moulds or siliciedrecrystallized whole-large
unabraded fragments

Spicule: 0%, possible early dissolution


Sponge: 020%, whole-unabraded to
weakly fragmented, silicied/
recrystallized

TC1

TC2

TC3

TC4

SP1

SP2

SP3

SP4

LM1 and LM2

40- > 90%, abundant bryozoa (arborescent, robust branching,


fenestrate), with subsidiary echinoderms, foraminifera,
pectinids, infaunal bivalves & serpulids; whole-unabraded
to abraded grains

Formerly 20-~40%, moulds of dominantly bryozoa (mostly


fenestrate), with subsidiary pectinids, echinoderms, robust
infaunalbivalves, & minor high-spired gastropods; whole
and unabraded

No calcareous moulds have been detected

Formerly Tr., very rare moulds of robust bivalves;


poorly preserved

Formerly Tr. to ~20%, moulds of epifaunal bryozoa, pectinids,


echinoiderms, minor high-spired gastropods, infaunal
bivalves, nautiloids; whole and unabraded

Formerly ~3040%, moulds of dominantly epifaunal bivalves,


mostly pectinids, oysters, subsidiary bryozoa, echinoderms;
whole, unabraded

Formerly Tr., very rare moulds of high-spired gastropods;


whole and unabraded

Formerly ~2%, moulds of fragile and robust infaunal bivalves,


with minor gastropods and pectinids; whole and unabraded,
rare double-valved shells

Formerly ~5%, moulds of dominantly robust infaunal grains


bivalves, minor epifaunal oysters, mussels, pectinids; whole
to comminuted grains

Carbonate percentage, assemblage and taphonomy

Chondrites
Thalassinoides
(45)

Thalassinoides
Planolites (5)

Thalassinoides
(difcult to identify,?35)

Thalassinoides
(difcult to identify,?45)

Thalassinoides (5)

Chondrites
Thalassinoides (5)

Thalassinoides
Planolites (5)

Skolithos (23)

Bioturbation (intensity)

Diatoms, dominantly benthic, unbroken and unabraded, have been identied in TC2, TC3, SP1 and SP2 facies. Bioturbation intensity based upon the scale of
Droser & Bottjer, 1986. Tr., trace.

Spicules and rigid-bodied sponges

Facies

Table 2. Faunal assemblage of each facies.

Late Eocene sponge-rich sedimentation


563

564

P. R. Gammon and N. P. James

Fig. 3. Regional and local


stratigraphy, plus correlations,
for sections of the Pallinup
Formation, Princess Royal
Spongolite and Wilson Bluff
Limestone (redrawn from
Gammon et al., 2000a). Lefthand black bar in Pallinup
Formation section indicates
meteoric silicication. Dashed
line indicates different portions of a relative sea-level
cycle; T, transgressive; H,
highstand; R, regressive. U1 to
U4 are the informal Pallinup
Formation units of Gammon
et al. (2000a).

sponges, and is capped by aerially extensive


spiculites and spongolites (Fig. 3; Cockbain,
1968a; Darragh & Kendrick, 1980; Gammon et al.,
2000a). The Pallinup Formation can be subdivided into ve discrete units, four lower informal
units (U1U4) and a formal, uppermost, spongedominated unit, the Fitzgerald Member (Fig. 3;
Gammon et al., 2000a). Units 1 and 2 are transgressive sandstones, whereas units 3, 4 and Fitz-

gerald Member are regressive muddy and/or


spiculitic deposits (Gammon et al., 2000a). Faunal
content is predominantly molluscs in Unit 1,
sponges and molluscs in Unit 2 and sponges in
all of the upper units. Gammon et al. (2000a)
interpret, on the basis of sediment provenance and
transport mechanisms, the high-stand point for the
Pallinup Formation sequence to be a little below
the unit 23 boundary.

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation


Palaeovalleys incised into Yilgarn Craton regolith (Fig. 2) and connected to the Eucla and
Bremer Basins, contain age-equivalent Eocene
strata, the mid-Eocene Norseman Limestone and
the Upper Eocene Princess Royal Spongolite
(Clarke & Phillipps, 1952; Clarke, 1993, 1994).
Coeval mid-Upper Eocene marine deposits within
the Eucla Basin proper are bryozoan marls and
limestones of the Wilson Bluff Limestone. Further
east, along the northern margin of the Eucla
Basin, palaeovalleys are lled by the mid-Upper
Eocene Pidinga Formation, with the spiculitic
Bring Member at the top (Benbow et al., 1995). In
contrast to the Bremer Basin, post-Eocene subsidence in the Eucla Basin resulted in later transgression and deposition of Oligo-Miocene
carbonates (Toolinna, Abrakurrie and Nullabor
limestones; Lowry, 1970; James & Bone, 1994;
Li et al., 1996).
Diagenesis considered critical to facies interpretation consists of clay neomorphism and
authigenesis, interpreted here as early submarine (near sea-oor verdinization). Despite general post-Eocene aridity (Kemp, 1978), late
meteoric diagenesis is locally severe, especially
towards the top of the succession at the
present land-surface. Late, meteoric diagenesis
includes porcellanitization and silicication,
lateritization, and leaching/oxidation (cf. Wise
& Weaver, 1973). Although rare, moulds of
calcareous organisms occur within all biosiliceous facies, indicating that the siliceous
sediments are not a diagenetic `remanie' deposit.

Regional palaeogeography
Pallinup Formation biosiliceous sediments were
deposited within a complex palaeogeographical
setting. Basement granites and granitic gneisses of
the Archaean Yilgarn Craton and Proterozoic
Albany-Fraser Province (Thom & Chin, 1984)
underwent deep and prolonged chemical weathering during preceding Mesozoic and early Tertiary time, resulting in an etchplaininselberg
landscape, much of which is preserved today
(van de Graaff et al., 1977, 1981; see Twidale,
1982 for a review of etchplain formation; Ollier
et al., 1988; BMR Palaeogeographic Group, 1990;
Clarke, 1994; Twidale & Bourne, 1998; Sircombe
& Freeman, 1999).
The modern topography is dominated by a large
plateau over the Yilgarn Craton (hereafter termed
the upland plain, part of the Great Western
Plateau; Veevers, 1984). The southern termin-

565

ation of the upland plain is a relatively steep


slope, interpreted as a degraded Late Eocene
coastal escarpment (Figs 2, 4 and 5; see Bunting
et al., 1973). From the base of the escarpment
another plain slopes gently seaward to the modern coastline (Figs 4 and 5; hereafter termed the
coastal plain, although the term is used with no
genetic, depositional connotation). Oligocene
uplift (Cope, 1975) elevated Upper Eocene sediments of the Bremer Basin, and post-Eocene
highstands, have eroded the southern edge, but
not submerged or transgressed these sediments.
The coastal plain is, therefore, essentially the top
of the Upper Eocene succession, veneered by a
metre-thick cover of Quaternary aeolian sand,
with numerous inselbergs poking through
(Fig. 4). The onland Pallinup Formation, thus,
represents both the topographically and stratigraphically highest marine sediments in the
onshore areas of the Bremer Basin, and the
modern coastal plain land surface (i.e. top Pallinup stratigraphical surface), therefore preserves
Late Eocene facies-geographical relationships
intact (Figs 4 and 5). The seaward termination of
the coastal plain is a heterogeneous arrangement
of basement topographical highs, post-Eocene
escarpments cut into Palaeogene sediments and
Pleistocene sand dunes. Inselbergs continue offshore as a series of bedrock shoals and islands,
especially numerous in the Recherche Archipelago (Figs 2 and 4).

Depositional palaeogeography
The mid-Late Eocene climate was humid (Kemp,
1978; Carpenter & Pole, 1995) and run-off from
the low-gradient upland plain was probably
sluggish, resulting in a ne-grained sediment
supply (Gammon et al., 2000a). The Werrilup
and Pallinup sequences record the two highest
relative sea-levels in the Bremer Basin Cenozoic
succession (Cockbain, 1968a; Gammon et al.,
2000a). In both transgressions, shoreface erosion
stripped deeply weathered material to expose
fresh basement rock at a topographically irregular, inselberg-prone weathering front. Shoreface
erosion led to deposition of Pallinup sediments
within topographically complex inner-shelf
embayments, shoals and island archipelagos
formed by high-standing basement which represents a weathering front that was stripped by
shoreface erosion of overlying regolith (Fig. 4).
Topographically low areas were probably
smoothed and oored by Werrilup Formation
sediment. The inner margin environments were

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

566

P. R. Gammon and N. P. James

Fig. 4. Modern southern Western


Australia geography and the ancient
palaeogeography. (A) Inselberg
rising above the Eocene plain;
approximately 60 km east of Esperance. Inselberg rises approximately
60 m above the plain (Esperance
1:250 000 topographical sheet). (B)
The Esperance Bay with many
inselbergs islands and shoals, suggesting this is a close modern analogue for the Late Eocene geography.
The Bay is approximately 15 km
from the near to far shore. (C) Flat
Eocene coastal plain abutting coastal
inselbergs. WR, Whoogarup Range;
EP, Eocene coastal plain; WB,
Woolbernup Hill. Looking southwest across the foreground Hamersley River valley that is blackened
due to bushre. See Fig. 5 for
approximate scale, position and
direction of (C).

bordered to the north by a shoreface escarpment,


which gave way northwards to the upland etchplain overlying a deeply weathered regolith. The
general palaeogeography can be subdivided into
four local sectors (S14) within which Upper
Eocene sediments accumulated.

Albany-Hopeton embayments with


extensive archipelagos, sector 1
The coastal plain in this sector is punctuated by a
north-east trending line of basement highs composed of granulite-grade metamorphic quartzites,
gneisses, schists, migmatites and granites that
form the core complex of the Proterozoic AlbanyFraser orogen (Fig. 4; Clarke & Phillipps, 1952;
Thom & Chin, 1984). The quartz-rich metamorphics were probably more resistant to deep
chemical weathering (see Twidale, 1982), and

hence the core complex forms a line of extensive


inselbergs that determined a Late Eocene palaeogeography of small, inselberg-protected embayments, whereas offshore, basement inselbergs
formed shoals and an extensive island archipelago (Figs 2 and 5). The modern geography
suggests that there were broad gaps in the
archipelago chain, so although embayments were
the most protected for this region, there is no
palaeogeographical (or sedimentological) evidence for embayment isolation.

Hopeton-Esperance, a vanished record, sector 2


A low-lying plain with extensive sand-dunes
suggests shoreface erosion during post-Eocene
highstands has stripped much of the Upper
Eocene succession. The few isolated Pallinup
Formation exposures (e.g. Munglinup High)

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation

567

Fig. 5. Detailed topographical map of a portion of the coastal area. Black dashed lines are interpreted Late Eocene
basement highs, discussed in the text. Note the continuity of basement highs along the modern coast (black line),
which emphasizes the considerable protection, but not isolation, of Late Eocene embayment and archipelago palaeoenvironments. Angle/eye symbol is the approximate position and direction of view for Fig. 4C. Each successive
50-m contour interval is incrementally darker shaded with increasing height above sea-level. The 0 and 200-m
contours are outlined. Facies variation can be seen from representative Pallinup Formation localities (black dots,
some of these localities are also mentioned in text and other gures): CB, Cheyne Bay; P, Pallinup type section at
Beaufort Inlet; B, Bremer Bay; T, Twertup; F, Fitzgerald Inlet; H, Hamersley River (see Gammon et al., 2000a for full
locality index). The post-Eocene rivers incised into Upper Eocene successions are PR, Pallinup River; FR, Fitzgerald
River. Landmarks noted in Fig. 4C: WB, Woolbernup Hill, WR, Whoogarup Range. Grid lines are Australian Map
Grid co-ordinates, and are in kilometres east (E) or north (N).

are within small embayments similar to those


of S1.

Esperance-Israelite Bay island archipelagos,


sector 3
As for sector 1, Upper Eocene sediments surround basement inselbergs. Erosion of Upper
Eocene sediments during post-Eocene high sealevels has probably returned the Recherche
islands to an archipelago geography similar to
that during Late Eocene deposition (Figs 2 and 4).

A heavily degraded Late Eocene palaeo-escarpment separates the upland and coastal plains. The
palaeo-escarpment occurs at 50100 m elevation
and 310 km inland from the coast (Fig. 2). The
S3 basement is dominated by granites and gneisses, and in comparison to S1 contains substantially less topographically high basement
inselbergs, presumably due to the lack of quartzrich metamorphic rocks. Late Eocene palaeoenvironments would have been less protected than
those in sector 1 (Fig. 2).

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

568

P. R. Gammon and N. P. James

East of Israelite Bay Eucla Basin open shelf,


sector 4
The eastern-most Eucla Basin proper sector is
dominated by subsidence, and palaeo-topography
was buried beneath the >300-m-thick Eocene
succession (Fig. 2; Lowry, 1970; Veevers, 1984;
Jones, 1990; Feary et al., 1994). Seismic proles
do not identify basement topographic inuences
(Feary et al., 1994; Feary & James, 1998), and
sector 4 represents an open-shelf depositional
environment with a typical, low-gradient, essentially at-shelf morphology.
FACIES ASSOCIATIONS
Within the Pallinup Formation and Wilson Bluff
Limestone three distinct facies associations have
been recognized: terrigenous, biosiliceous and

calcareous, with many gradational lateral and/or


vertical facies boundaries (Tables 13).
TERRIGENOUS CLASTIC FACIES
ASSOCIATION

TC1 cross-bedded sandstone


This facies is ne to coarse-grained, well-sorted,
quartzose sandstone that is distinctively trough
cross-bedded (Unit 1; Tables 13; Fig. 6). Trough
cross-bed sets are up to 3 m long by 1 m thick,
and below the angle of repose (27 maximum
measured angle). TC1 sandstones ne upward,
both overall and within individual cross-bed
sets. Sandstones locally form large dune complexes >100 m wide and >10 m high, and only
occur within unit 1 of the Pallinup Formation
(Fig. 3).

Table 3. Bedding and sedimentary structures of facies.


Facies

Bed thickness and bed contacts

Sedimentology structures and comments

TC1

20300 cm
Sharp top and basal contacts

Multi-directional trough cross-bedding, HCS, laminated,


rippled, aser-bedded, and massive beds occur

TC2

30250 cm;
basal and top contacts are gradational
over 510 cm (?bioturbation)

Alternating coarser (less muddy, less sponge-rich) ner


(more muddy, more sponge-rich) grained parallel
beds; burrowed rmgrounds in ne-grained beds at
Munglinup High

TC3

Vague 30100 cm;


contacts as for TC2

When fresh, vague medium (less sponge-rich) to dark


brown (more sponge-rich) parallel colour alternations

TC4

30 cm;
basal contact sharp, burrowed; top
contact gradational over 5 cm

Top of underlying rmground burrowed and inlled


by green clays

SP1

Vague 30100 cm;


contacts as for TC2

Vague medium to dark brown parallel colour


alternations when fresh; 10- to 100-cm-thick, variably
cemented, porcellanitized beds are at least partly
diagenetic artefacts

SP2

Massive;
basal contact sharp to gradational
(510 cm); top contact gradational
over ~13 m

10 cm thick, by 1- to 5-m-long lenses of rigid-bodied


sponges; uniform colour (oxidized red/orange or
white) inhibits bed, bed-structure & bioturbation
identication

SP3

Massive
basal contact gradational over ~13 m

Uniform red/orange-oxidized colour and massive


iron-lateritization inhibit bed, bed-structure &
bioturbation identication; aligned sponge fragments
in SP3b

SP4

1050 cm;
basal and top contacts gradational over
510 cm (?bioturbation)

Dened by variable cementation that appears inversely


related to terrigenous clastic percentage; cementation
also follows Thalassinoides burrows

LM1 and LM2 10- to 50-cm-thick beds in 3- to


7-m-thick cyclic packages; basal and
top contacts gradational (510 cm) to
sharp

Marls, wackestones and packstones contain 1050 cm


thick parallel beds within 37 m thick cyclic packages;
grainstones and rudstones are massive

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation

569

Fig. 6. Terrigenous clastic facies. (A) Iron cemented, cross-bedded sandstone facies (TC1). Scale bar (arrowed) is
15 cm long; unit 1 at Cheyne Bay. (B) Photomicrograph (plane polarized light) of muddy sandstone facies (TC2), with
spicules (S), angular quartz (Q) and verdine (V) fragments. Scale bar is 02 mm long; unit 2 at Beaufort Inlet. (C)
Preferentially silicied Thalassinoides burrows on a muddy sandstone facies bedding surface. Scale bar 15 cm; unit
2 at Cheyne Bay. (D) Intensely bioturbated mudstone facies (TC3) with abundant verdine peloids (black spots), rigidbodied sponge fragments (arrowed), and many white streaks that are probably remnants of sponge fragments. Scale
bar is in centimetres; unit 3 at Fitzgerald Inlet. (E) Sponge-rich mudstone facies bed that is interpreted to represent
storm activity. White patches are large and small sponge fragments (abundant verdine peloids are too small to see in
this photograph). This bed is laterally continuous for at least 1 km; unit 3 at Fitzgerald Inlet. Photograph localities
given on Figs 2 and 5.

Interpretation
The cross-bedded sandstone facies is interpreted
to represent upper shoreface depositional environments, probably shoreface dunes (Clifton
et al., 1971; Collins & Banner, 1980; Swift et al.,

1991; Walker & Plint, 1992; Davis et al., 1993).


Large trough cross-beds, along with local dune
morphology, both conrm strong unidirectional
currents (Clifton et al., 1971), although current
directions uctuated through time. Skolithos and
robust infaunal bivalve moulds (Table 2) are

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

570

P. R. Gammon and N. P. James

consistent with an energetic, mobile sand environment (Pemberton et al., 1992). Hummocky
cross-stratication (HCS) and wave-ripples indicate orbital wave currents, and together with the
overall ning-up, suggest that middle and lower
shoreface depositional environments are also
represented within this facies.

TC2 muddy sandstone


This thick-bedded, moderately sorted, ne to very
ne grained, muddy sandstone to sandy mudstone facies generally exhibits preferential silicication that accentuates ner-grained beds and
imparts a rhythmic cementeduncemented character to outcrop (Tables 13; Fig. 6B). This facies
is vertically gradational upward into the mudstone facies (TC3), and laterally gradational to
muddy calcareous spiculite facies (SP4; Fig. 7).
Soft demosponges and hexactinellid sponges are
common throughout.

Interpretation
Fauna, ichnology and sedimentology (Tables 13)
indicate a shallow to moderate-depth, sublittoral
depositional environment (inner to mid-shelf;
Nummedal, 1991; Swift et al., 1991; Nittrouer &
Wright, 1994). The bimodal mud and sand grainsize distribution suggests that these deposits are a
product of both suspension (fair weather) and
bedload (higher energy,?storm) transport processes (intense bioturbation generally precludes
storm-bed preservation; Clifton et al., 1971; Swift
et al., 1987, 1991; Nummedal, 1991; Nittrouer &
Wright, 1994). The relative inuence of these two
processes determines the sandmud ratio, and is
probably related to location within the embaymentarchipelago complex (see below).

TC3 mudstone
These sediments are moderately sorted, intensively burrowed, clayey siltstone to silty claystone to claystone, and are medium to dark brown
when fresh, but bleached off-white when weath-

ered (Tables 13; Fig. 6). The percentage of verdine usually increases with increasing clay per
cent, but also varies independently on a centimetre scale. Rare, verdine-replaced benthic
diatoms are usually whole and unabraded. Occasional, thin (10 cm) beds are locally continuous
(>05 km), and contain abundant, whole, rigidbodied sponge fossils without basal attachment
structures (Fig. 6E), along with abundant verdine
peloids. Rigid-bodied sponge growth morphology
is mostly digitate, with less abundant folded
palmate growth forms. TC3 grades laterally and
vertically into muddy spiculite (SP1), and grades
vertically into muddy sandstone (TC2) facies
(Fig. 7). Soft demosponges and hexactinellid
sponges are abundant within this facies.

Interpretation
The mudstone facies is interpreted to represent
protected sublittoral (inner to mid-shelf) environments, with minor storm inuence. The Pallinup
Formation unit 23 sandstone (TC2) to mudstone
(TC3) transition (Fig. 3, Fitzgerald Inlet section) is
essentially an end to sand supply. Sand was
supplied to the embayments from eroding sea
cliffs (shoreface escarpment erosion), a process
which ended at the start of regression (see
Gammon et al., 2000a; for full provenance and
sequence stratigraphical details). Mud was continually supplied via palaeovalleys. Regression
thus started sometime prior to the unit 23
boundary (Fig. 3). This suggests that mudstone
water depths were probably similar to or possibly
even shallower than that of the coarser-grained
muddy sandstones (TC2). This interpretation is
supported by the local presence of rigid-bodied
sponge and verdine-rich beds that are interpreted
as surfaces of reworking, presumably due to
higher-energy currents, possibly storms. Irregular,
angular verdine and sponge fragments also suggest reworking, although at least some breakage
would be due to intense bioturbation. The
numerous upright sponges probably bafed
wave-currents. Interlocked, meshed soft demosponge spicule mats, similar to those in modern

Fig. 7. Gradational relationships


between Pallinup Formation facies.
Dashed lines are inferred lateral
correlatives.
2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation


high-latitudes (van Wagoner et al., 1989; Conway
et al., 1991), may have provided resistance to
wave-reworking.

TC4 green-clay
Poorly sorted green-clay (verdine) beds, are only
observed at the Munglinup High section (Fig. 2;
Tables 13). Epifaunal and infaunal mollusc
moulds are common, whereas sponges are sparse
within this facies. This facies grades upward to
mudstone facies (TC3).

571

Interpretation
This facies is interpreted to represent periods of
slow, condensed, authigenic green clay accumulation in subshoreface environments (see Odin,
1988). Such deposits only occur on submerged
topographical highs isolated from terrigenous
clastic sedimentation.
BIOSILICEOUS FACIES ASSOCIATION

SP1 muddy spiculite


Moderately sorted muddy spiculites are the most
common lithology encountered within the study
area (Tables 13; Fig. 8A). Soft demosponge spicules are diverse but dominantly monaxons.
Hexactinellids and possibly lithistid sponges are
less numerous than in other facies. Average
terrigenous clastic grain-size increases with
increasing spicule per cent, whereas calcareous
fossils increase laterally basinwards. This facies
grades laterally and vertically into calcareous
spiculite (SP4), pure spiculite (SP2) and mudstone (TC3) facies (Fig. 7).

Interpretation
The muddy spiculite facies is interpreted to
represent accumulation in sublittoral depositional
environments. Increasing terrigenous clastic
grain-size and spicule percentage is interpreted
as evidence of increasing mud bypassing in
comparison to TC3 mudstone, probably the result
of increasing hydrodynamic energy with decreasing water depth. Spicules, with a ne to very ne
sand modal grain size, are conceptually equivalent to terrigenous sand particles. The unbroken,
Fig. 8. Biosiliceous facies. (A) Photomicrograph (plane
polarized light) of spicules (S), minor terrigenous
clastic quartz silt (Q) and mud within a muddy spiculite lithology (SP1). Scale bar 02 mm; unit 4 at
Hamersley River. (B) Photomicrograph (plane polarized
light) of pure spiculite facies (SP2), with minor terrigenous clastic silt (examples arrowed), but very little
clay. Scale bar 02 mm; Fitzgerald Member at Hamersley River. (C) Vertical cross-section of whole, rigidbodied lithistid sponges supported by a pure spiculite
matrix (spiculitic spongolite subfacies SP3a). Note the
overturned vase-shaped sponge (arrowed) with spicules concentrated in the cavity below, suggesting that
current activity was intermittently too strong for spicule accumulation at the bed surface; Fitzgerald
Member at Hamersley River. Photograph localities are
on Fig. 5.
2001 International Association of Sedimentologists, Sedimentology, 48, 559584

572

P. R. Gammon and N. P. James

unabraded nature of the spicules suggests


minimal transport, probably due to bafing by
the sponge meadows, and/or wave-resistant spicule mat formation (see van Wagoner et al., 1989;
Conway et al., 1991).

SP2 pure spiculite


These poorly to well-sorted pure spiculite deposits are up to 6 m thick (Tables 13; Fig. 8B). Soft
demosponge spicules dominate and are mostly
unabraded and unbroken monaxons, randomly
oriented within SEM stubs. Minor rigid-bodied
lithistids and rare hexactinellid spicules are also
present. Apart from the ubiquitous monaxons
(monaxons are found within diverse demosponge
taxa), spicule composition is very heterogeneous,
with adjacent eld samples containing quite
different spicule complements (especially of
unusually ornamented varieties). Microscleres
(spicules < 63 lm in length) or small diatoms
are rare. Sediment is generally oxidized, unlithied and friable, with negligible silica diagenesis.
Pure spiculites grade both laterally and vertically
into muddy spiculite (SP1) and spongolite (SP3)
facies (Fig. 7).

Interpretation
The pure spiculite facies is interpreted as a lower
shoreface deposit derived from the disintegration
of numerous soft-bodied demosponges. In comparison to muddy spiculite, decreasing water
depths increased wave energy and resulted in
increased mud bypassing (Fig. 7). Adjacent samples with heterogeneous spicule complements
suggest minimal transport and mixing. Spicules
were possibly derived from individual soft-bodied
sponges, remaining near situ post mortem. Random spicule orientation, plus lack of abrasion and
breakage also suggest minimal post mortem spicule transport and reworking, despite spicules
being the terrigenous sand substitute within lower
shoreface environments. Three interrelated reasons why spicules were probably not transported
in lower shoreface environments may be: (1)
current bafing by what must have been a prolic
living sponge population; (2) the rod-like shape of
loose monaxon spicules will dictate unusual
transport properties; and (3) the possible formation
of microbe-bound, interlocked-meshed spicule
mats, as found in modern, deep-water spiculite
mats (van Wagoner et al., 1989; Conway et al.,
1991; Henrich et al., 1992). The massive unlaminated nature of these deposits does not suggest

spiculite mats or microbial lms. However, bioturbation, silica diagenesis, weathering and the
poor carbonate preservation within these facies
may have obscured the evidence of such mats. The
paucity of mud, microscleres, or small diatoms
suggests that ner-grained material was probably
bypassed to muddy facies. It also highlights the
terrigenous sand-poor, ne-grained nature of the
Late Eocene silliciclastic sediment supply.

SP3 spongolite
These rigid-bodied sponge biostromes either contain (SP3a), or lack (SP3b and SP3c), a well-sorted
spiculite matrix (Tables 13; Figs 8 and 9). The
diverse lithistid demosponge assemblage (de
Laubenfels, 1953; Pickett, 1982, 1983; Gammon
et al., 2000b) contains mostly digitate and
branching-digitate growth forms, although platelike, globular, fan, vase and rosette morphologies
are also common. There is a consistent upward
gradation from SP3a to SP3b to SP3c (Fig. 7).

SP3a
These sediments, up to 3 m thick, are composed
of massive, poorly sorted, spiculitic spongolite
(Fig. 8C). Most sponges are whole and unabraded,
having only been broken from their basal attachment and randomly reoriented into a horizontal
plane. Basal attachment was mostly to other rigidbodied sponges. Plate-like growth morphologies
are dominantly right-way-up. Sediments grade
vertically upward from pure spiculite to wellsorted spongolite (Fig. 7).

Interpretation
These spiculitic spongolite deposits are essentially in situ mid-upper shoreface lithistid palaeocommunities (lithistid demosponges have
articulated spicules and hence a rigid-skeleton).
Like pure spiculites, hydrodynamic energy was
signicant, energetic enough to prevent deposition
of ne-grained terrigenous clastic sediment, and
increasingly winnow spicular material as SP3a
grades to SP3b. Current bafing by both living and
dead lithistid sponges and spicular mats were
again probably important in preventing wholesale
reworking of soft and rigid-bodied sponges.

SP3b
These moderately sorted lithistid spongolite
deposits are up to 2 m thick, and do not have a
spiculite matrix (Fig. 9B). Deposits grade vertic-

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation

573

Fig. 9. Biosiliceous and calcareous facies. (A) Large, generally unbroken and weakly abraded lithistid sponges
(lithistid spongolite subfacies SP3b). Similar to Fig. 8C, but with very few loose spicules preserved within cavities
between rigid-bodied sponges, suggesting that hydrodynamic energy was too high for spicule accumulation; Fitzgerald Member at Hamersley River. Gradations visible on scale bar (arrowed) are in inches. (B) Well-sorted spongolite
with lithistid sponge fragments both abraded and broken (well-sorted spongolite subfacies SP3c). Fitzgerald Member
at Hamersley River. (C) Large fenestrate bryozoan mould within a silicied muddy calcareous spiculite (facies SP4).
Black scale bar is 2 cm long; Fitzgerald Member at Bremer Bay. (D) Calcareous marl (LM1) containing silicied
lithistid sponges (arrowed). (E) Bryozoan grainstone (facies LM2) with common echinoid spines. (D) and (E) are from
Wilson Bluff Limestone exposure at Point Culver. Photograph localities are on Figs 2 and 5.

ally from spiculitic spongolite (SP3a) below, to


well-sorted spongolite (SP3c) above. Most lithistid
sponge fossils are whole, some exhibit minor abrasion/breakage, and are randomly oriented within a
horizontal plane. Attached sponges are generally
restricted to rare, large lithistids (Fig. 8B).

Interpretation
The gradation from SP3a to SP3b is interpreted to
reect increasing hydrodynamic energies, too
high to retain individual spicules in these ener-

getic upper shoreface environments. Only the


largest individual lithistid fossils were heavy
enough to remain in situ, and these indicate that
lithistid sponges did live in these environments.

SP3c
These sediments, up to 4 m thick, are massive
well-sorted spongolite composed of fragmented
and abraded lithistid sponges without a spiculite
matrix (Fig. 9A). SP3c deposits grade upward
from SP3b deposits below, and always form the

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

574

P. R. Gammon and N. P. James

top of the Upper Eocene stratigraphical section.


Sponge fragments average 25 cm in diameter,
and have specic gravities (SG's) between 13 and
19 g cm3, with a mode of 15 g cm3. These
deposits also contain: (1) clusters (up to 05 m2)
of digitate sponges with vertically to horizontally
aligned long-axes; and (2) plate-like sponges that
are mostly overturned, or locally subvertically
oriented. There are no attached or whole sponges
in this facies, together suggesting that sponges did
not live in these environments.

ods of higher hydrodynamic energy (?storms).


The most calcareous beds are marly limestone,
and are intermediate in composition between
Wilson Bluff marl and Pallinup spiculite. The
muddy calcareous spiculite (SP4) to marl (LM1)
transition is interpreted as a gradational facies
continuum (Fig. 7), although as yet the full range
of intermediate lithologies cannot be demonstrated due to poor outcrop between the Bremer and
Eucla Basins.

Interpretation

CALCAREOUS FACIES ASSOCIATION

Well-sorted, abraded and broken sponge fragments


imply repetitive reworking of these deposits, and
they are interpreted as beach surf and swash zone
accumulations (imagine a beach composed entirely of siliceous sponges!). These unusual deposits consistently cap the succession, suggesting that
their accumulation was in the shallowest marine
environment (beach) during regression, immediately prior to subaerial exposure. Reworked sponge
fragments were probably derived from former and/
or contemporaneous SP3a and SP3b lithistid
sponge communities. The low specic gravity of
these fragments suggests that vertically aligned
clusters of sponge fragments may be formed
through wave-worked packing, conceptually similar to pumice strandlines (Allen, 1982).

At Point Culver, basal bryozoan calcareous marls


grade upward to bryozoan limestones (wackestone, packstone, grainstone and nally rudstone;
Tables 13; Figs 2 and 9).

LM1 marl
Thin to thick-bedded marls are composed mainly
of carbonate mud (4060%), body fossils (20
30%) and terrigenous clay (2030%), with rare,
angular, monocrystalline quartz silt (Fig. 9D).
Robust branching and arborescent bryozoans
dominate (Bone & James, 1993 classication),
with moderately common silicied lithistid sponges (510%, locally 15%, commonly in lenses).
Fossils are unabraded and unbroken.

SP4 calcareous spiculite

Interpretation

These ne-grained sediments are poorly sorted,


muddy, calcareous spiculite with common to
abundant calcareous fossil moulds (Tables 13;
Fig. 9C). In those areas containing this facies,
calcareous faunal diversity and percentage
always increases stratigraphically upwards. Softbodied demosponge spicules dominate, particularly monaxons, with common lithistid sponges
and fenestrate bryozoans. All fossils are randomly
oriented within a horizontal plane and are generally unabraded and unbroken. This facies
grades vertically and laterally to muddy spiculite
(SP1), and laterally to muddy sandstone (TC2;
Fig. 7).

Common silicication (concretionary and disseminated) may indicate early remobilization of


siliceous sponge material resulting in a paucity of
spicule moulds. This interpretation is supported
by uncollapsed, silicied, Thalassinoides burrows that suggest early cementation (see Jones &
Fitzgerald, 1986). Marls are interpreted as lowenergy shelf deposits, formed below storm wave
base from suspension fallout and in situ biogenic
production (James & von der Borch, 1991; James
et al., 1992). The marls are lithologically and
biotically similar to many descriptions of the
deep-shelf `Sponge Megafacies' found in Mesozoic, Tethyan Europe (e.g. Trammer, 1991; Leinfelder et al., 1994; and references therein; Wiedenmayer, 1994; Rosales et al., 1995; Pisera,
1997a, 1999).

Interpretation
The depositional environment is interpreted as
sublittoral, probably inner to mid-shelf. Fossil
taphonomy indicates a relatively low-energy setting (Kidwell, 1991; Kidwell & Holland, 1991;
Meldahl, 1993; Kondo et al., 1994), whereas the
sandy terrigenous component may indicate peri-

LM2 limestone
Thick-bedded wackestones and packstones are
composed of 3050% bryozoa, increasing with

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation


decreasing terrigenous mud (Fig. 9E). Bryozoans
are dominantly robust branching, arborescent and
fenestrate growth forms (Bone & James, 1993
classication). Echinoids and foraminifera are
common, with minor infaunal bivalves, brachiopods, serpulids and rare siliceous sponges
(Lowry, 1970). Thick-bedded grainstones and
rudstones have similar biogenic components to
wackestone and packstone, but epifaunal bivalves
(mainly pectinids) and serpulids are more common, while fenestrate bryozoans are less abundant
and sponges absent. Breakage is ubiquitous and
abrasion is apparent in at least 25% of rudstone
grains. Terrigenous clastic material decreases
with decreasing mud per cent to less than 2%
quartz silt/sand in grainstones and rudstones.
Silicication (concretionary or disseminated) decreases with decreasing mud per cent, and is rare or
absent in grainstone or rudstone facies.

Interpretation
Taphonomy suggests rudstones accumulated in
shoreface environments (Kondo et al., 1994). The
marl to rudstone succession at Pt. Culver is
interpreted as a cleaning-upward succession due
to increasing hydrodynamic energy with decreasing water depths (Fig. 3; Gammon et al., 2000a).
Shallowing water depths were probably part of
the regressive phase of the Late Eocene transgressiveregressive cycle (McGowran et al.,
1997). Because the marls are interpreted to be
deep-water shelf deposits and the spiculites
shallow-water embayment deposits, it is likely
that the grainstones and rudstones were deposited after regressive subaerial exposure of
embayment areas, and that the marls are more
probably the time equivalents to Pallinup deposits (as schematically shown on Figs 3 and 7).
This is also suggested by the concentration of
sponges within marls. The calcareous biota
represents a typical heterozoan, cool-water carbonate assemblage (James, 1997), but with the
addition of siliceous sponges.

575

more open environments. Two consistent facies


trends occur within this palaeogeographical
framework, one east-west and one north-south.
Facies change eastward from spongolite (SP3),
spiculite (SP2) and muddy spiculite (SP1) deposits
in the most protected embayments of sector 1,
through the muddy spiculites and calcareous
spiculite deposits of island archipelagos between
Hopeton and Israelite Bay (sectors 23), to open
marine calcareous marls and limestones in sector 4
(Fig. 2).
A north-south transect through the most
sponge-rich sector 1 (well exposed in Fitzgerald
River National Park), illustrates pure spiculites
and spongolites at the palaeo-escarpment, grading
offshore through muddy spiculites in the central
protected embaymentarchipelago area, to calcareous spiculites in outer areas of the archipelago
(Fig. 10). Within sectors S2 and S3, a Cowan
palaeovalley-Munglinup High-Esperance transect
(nominally north-south due to limited outcrop,
Fig. 2), gives a spiculitemuddy spiculitecalcareous spiculite facies trend, an identical palaeogeographyfacies relationship to sector 1.
The north-south transect is also interpreted as a
deepening trend, from shoreface to inner and
mid-shelf depositional environments. Individual
sections that contain both shelf and shoreface
environments (e.g. Hamersley River, Fig. 5), have
constant ratios of biosiliceous to calcareous fauna
within all facies. This suggests that although
embayments had a range of water depths, the
change from biosiliceous to calcareous biota is
not facies or water depth-dependent. Instead, the
biota mirrors the degree of palaeogeographical
protection from the Late Eocene embryonic
Southern Ocean. Sponge-rich communities
enjoyed the most protected inboard localities
and calcareous organisms preferred the open
shelf, with a gradation between the two centred
on the outer, lesser protected, archipelago
islands.
DISCUSSION

PALAEOGEOGRAPHICAL CONTROL ON
SEDIMENTARY FACIES DISTRIBUTION
The Upper Eocene succession accumulated in an
area of complex inboard topography undergoing
low subsidence. West of Hopeton (S1) the Upper
Eocene succession was deposited in archipelagoprotected embayments, while east of Hopeton
(S2-3) archipelagos were still present, but with

Facies mosaic and sedimentation history


The Upper Eocene succession is a single major
transgressiveregressive sequence (Fig. 3). The
preceding palaeogeographical, hydrodynamic,
facies distribution and facies interpretation can
be combined into facies mosaics at different
stages during the Late Eocene relative sea-level
cycle (Fig. 11).

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

576

P. R. Gammon and N. P. James

Fig. 10. Schematic plan view, based upon Fig. 5, of facies distribution during Fitzgerald Member times. Spongolites
and pure spiculites are conned to the landward embayment margin, while the majority of the embayment accumulated muddy spiculites. Outer Archipelago areas accumulated calcareous spiculites that grade seawards to shelf
carbonate facies. All facies transitions are gradational.

Transgression
Initial transgression ooded embaymentarchipelago areas that had last been inundated during
mid-Upper Eocene Werrilup Formation time.
Reworking of Werrilup Formation sediment plus
relic and palimpsest deposits led to the accumulation of unit 1 cross-bedded sandstone facies
(TC1). This facies is widespread, occurring as a
blanket across the basin. Deposition of unit 2
(TC2) started once relative sea-level had risen
high enough to erode deeply weathered regolith,
and escarpment retreat began. Shoreface escarpment erosion provided a source of rst-cycle
sand for unit 2 muddy sandstone facies (Figs 3
and 11A; see Jones, 1990). The late transgression
facies mosaic is relatively simple, with abundant
terrigenous material derived from the eroding
escarpment blanketing the embayments/archipelago. The start of unit 2 marks the abrupt
transition from a carbonate-dominated to a
siliceous sponge-dominated benthos throughout
embayment and archipelago environments. Siliceous sponges dominated embayment benthic
assemblages for the remainder of the relative
sea-level cycle.

Early regression
Regression ended escarpment erosion and, therefore, neritic embayments lost their source of

terrigenous sand. Early regression is therefore


approximated by the unit 23 boundary (Fig. 3;
see Gammon et al., 2000a for the detailed
sequence stratigraphy). From this time on, lowgradient rivers delivered only mud-dominated,
suspended-load material to the coast (clay in
particular). Despite continued topographical
complexity, the early unit 3 (early regressive)
facies mosaic is again simple, with mudstone
deposition blanketing most of the embayment
(Fig. 11B). Combined unit 1 plus 2 thickness is
commonly <10 m, and the initially steep-sided
inselberg topography probably remained, limiting the aerial extent of shoreface environments
(Twidale, 1982). Where present, embayment
shoreface environments lacked a terrigenous
sand source, and their substitute [pure spiculite
(SP2) and spongolite (SP3) facies] probably
accumulated from this time on. Condensed
green clay (TC4) deposition covered isolated,
submerged topographical highs. Marl deposition probably continued in open-shelf environments.

Late regression
A more complex facies mosaic developed as
regression continued (Fitzgerald Member deposition). Sedimentation had preferentially inlled
low-lying areas and reduced initial embayment
archipelago topographical complexity (Fig. 11C).

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation

577

Fig. 11. Schematic facies mosaics at the late transgressive (unit 2), early regressive (unit 3) and middle to late
regressive (Fitzgerald Member) stages of the relative sea-level cycle. The mosaics demonstrate the evolving relationship between depositional facies and the embayment-archipelago-open marine palaeogeography. Initially the
embayments were terrigenous clastic dominated, with carbonates and marls in open marine settings. By Fitzgerald
Member times siliceous sponges dominated embayments, and calcareous epibenthos, principally bryozoans, dominated open marine areas. Localities are schematically plotted from the mapped palaeogeography, and are abbreviated as for Figs 2 and 5 (except for M, Munglinup).
2001 International Association of Sedimentologists, Sedimentology, 48, 559584

578

P. R. Gammon and N. P. James

Shoreface environments of spongolite and spiculite facies (SP2, SP3) reached their greatest aerial
extent as low-angle, sediment-oored shoreface
deposits accumulated at the base of steep-sided
inselbergs. These biosiliceous deposits graded
laterally to muddy deposits (SP1, TC3) in deeper
water central embayment areas. Embayment
muddy spiculites graded to calcareous spiculites
(SP4) in outer archipelago areas. Sediment inll
reconnected submerged topographical highs that
were previously isolated from terrigenous sedimentation, and muddy spiculite (SP1) replaced
green clay (TC4) deposition. Sponge-rich marl
(LM1) deposition is interpreted to have continued
in open-shelf localities.

Hydraulic regime
The Late Eocene was the last Palaeogene climatic
optimum (Shackleton & Kennett, 1975; McGowran, 1987; 1989b; Prothero, 1994), with a
reduced poleequator temperature gradient and
low zonal wind strengths (Janecek & Rea, 1983).
The overall oceanic wave regime impinging on
the southern Australian Late Eocene coast was
probably substantially lower that that of the
modern high-energy Southern Ocean. The southern Australian coast faced a large east-west
oceanic gulf between Australia and Antarctica
(Fig. 1; Lawver et al., 1992). The gulf was open to
deep and shallow ow to the west. To the east,
deep-water ow was impeded by a connection
between the South Tasman Rise and Antarctica,
but probably open to minor surface water ow
(Kennett & Warnke, 1992). Theoretically, Coriolis
forcing would circulate gulf waters in an anticlockwise gyre, which would be countered by the west
wind drift along the entire Late Eocene southern
Australian margin, resulting in diminished longshore currents (Fig. 1; Kemp, 1978). Atmospheric
models suggest that such opposing forces resulted
in offshore winds along the southern Australian
coast during the Late Eocene (Barron & Peterson,
1991; Kennett & Warnke, 1992; Golonka et al.,
1994). Offshore winds would have lessened any
oceanic waves impinging on the Late Eocene
Australian coast (Wright, 1995).
The modern coast is microtidal (Hydrographic
Service, 1995), which is consistent with the
general principle of narrower shelf-smaller tides
(Wright, 1995). The Late Eocene shelf was of
similar narrow width (Fig. 2; 3050 km wide),
and was probably also microtidal. Palaeovalleys,
with their >200 km length, may have had somewhat higher tidal currents.

Current energy would have been a combination


of wave, longshore and tidal forces in the protected embayments. Offshore winds, opposing west
wind drift and Coriolis currents, and a microtidal
regime would most probably have produced low
hydrodynamic energy conditions along the Late
Eocene Australian coast. Hydrodynamic energy
would have been further attenuated by the
embaymentarchipelago palaeogeography (Fig. 5),
resulting in tranquil embayment conditions, and
slow rates of exchange between embayment and
open-shelf waters. Occasional storm beds attest to
stronger currents episodically sweeping across
these shallow substrates, possibly preventing
benthic water stagnation. Storm beds are only
observed within mudstone (TC3) beds, and their
lag products are unbroken, unabraded sponge
skeletons and ovoid verdine pellets. This suggests
that higher energy storm currents within embayments were only capable of moving mud particles, and thus of generally low energy. The
overall hydrodynamic conditions within the protected embaymentarchipelago environments
was probably very calm, an energy regime equivalent to that of many modern sponge habitats
(Reid, 1968; van Wagoner et al., 1989; Conway
et al., 1991; Tabachnick, 1991).

Sponges as sedimentary particles


Sponge body fossils and spicules cannot be
interpreted as hydrodynamically equivalent to
similar sized terrigenous grains or carbonate
particles because of their very different specic
gravity. From three replicate analyses, sponge
Opal-A ranged in value from 179 to 186 g cm3,
with the average as 182 01 g cm3 (which is
within the 173213 g cm3 range quoted for opal
in Weast et al., 1972). The same method yielded
an SG of 263 01 g cm3 for pure quartz sand.
Measurement of 30 lithistid fragment SGs (pores
plus spicules) gave a range of 1216 g cm3 (plus
one outlier at 19 g cm3), with an average specic
gravity of 15 01 g cm3.
Wright (1995) gave theoretical formulae for the
critical shear stress (scr) for sediment particle
entrainment scr Cqwv2 YDg(qs qw) [Shields
equation; C (j ln1(z/z0))2 coefcient of drag
between the bed and current; j 0408 von
Karmens coefcient; (z/z0) measure of bed roughness; qw, qs density of water and sediment particle
respectively; v velocity of current; D diameter of
sediment particle; g acceleration due to gravity;
Y shields parameter 005 for coarse particles in
turbulent ow; equations are for planar, non-

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation


rippled beds]. For an ideal transfer of uid force to
the sediment particle (i.e. C 1), and a 25-cm
diameter, 16 g cm3 sponge fragment (i.e. about
the upper limit), a current velocity of 85 cm s1
is required for entrainment. Wright (1995) in
studies of the low-energy Louisiana coast suggested coefcients of drag (C) of around 0003. Where
z/z0 is a measure of bed roughness for ne silt and
sand-sized sediment, z is usually taken as 1,
representing a depth of 1 m above the bed,
whereas z0, the height unevenness at the bed
surface, is taken here as 25 cm, the diameter of
one sponge fragment. This gives C 001, substantially greater than for ner-grained sediments,
as expected with the large size of these fragments.
Substituting for C in the above equation gives a
velocity of 60 cm s1 for entrainment of the
25-cm-diameter sponge fragment.
Sponge fragments are interpreted as part of
spongolite beach deposits, and 60 cm s1 currents
therefore represent swash-zone currents. Modern
beach wave regimes with such currents are
regarded as low-energy coasts (e.g. the Louisiana
coast; Wright, 1995). An equivalent-sized quartz
pebble would require high-energy 200 cm s1
currents, 25 times that of sponge particles. The
coarse-grained, well-sorted spongolite (SP3c)
deposits do not, therefore, represent high-energy
currents, conrming the notion that the mean
current strength during Pallinup Formation
deposition was low. Storm currents would, therefore, also have been of generally low hydrodynamic energy, but obviously strong enough to
resuspend the muddy substrate.
Soft demosponges in lower shoreface spiculites
(SP2) are separated from, but grade into, rigidbodied lithistid demosponge spongolites (SP3) in
upper shoreface conditions. As terrigenous
material is absent, the biotic separation is at least
partly due to the interpreted change in hydrodynamic regime. If rigid lithistid skeletons are more
resistant to wave-breakage than soft demosponge
tissue, then the more energetic upper shoreface
environments probably conferred an unusual
competitive advantage to lithistid sponges,
although the lack of life position specimens in
well-sorted spongolites (SP3c) suggests that they
in turn were limited by the surf-zone waveenergy. In lower shoreface environments, hydrodynamic energies were low enough to make
relative skeletal strength irrelevant, and soft
demosponges out-competed lithistids, as is found
in many present-day sponge-rich environments
(e.g. Rutzler, 1990; Reitner & Keupp, 1991).

579

Palaeoenvironments and palaeogeography


Two local analogues to the Pallinup Formation
geography are the modern Recherche Archipelago
(Fig. 2) and the underlying Werrilup Formation
(Cockbain, 1968a; Clarke, 1993). However, calcareous benthic marine organisms dominate the
biota in both the modern Recherche Archipelago
and ancient Werrilup Formation. Thus, despite
the close relationship between palaeogeography
and Western Australian Upper Eocene facies, the
physical palaeogeography alone was insufcient
to explain these large sponge deposits. The same
conclusion can be drawn from considering highstands in general, because ooded embayments
and archipelagos must have occurred during
numerous Phanerozoic highstands. Yet literature
reports of shallow-water sponge-dominated
deposits are few, and generally describe thin
horizons within predominantly calcareous or
siliciclasticcalcareous successions (e.g. Cavaroc
& Ferm, 1968; Lane, 1981).
While Upper Eocene muddy substrates within
embayments promoted sponge growth, open
marine muddy localities were more conducive to
calcareous epibenthos. These similar substrates
infer similar physical parameters (sediment texture, hydrodynamics), suggesting that water chemistry (salinity, nutrients, oxygenation) was the
most signicant factor in epibenthic segregation.
The overall embayment hydrodynamic regime was
low, and the palaeogeography probably led to
somewhat limited water exchange between
embayments and the open shelf. If embayment
water chemistry promoted sponge dominance,
sluggish water exchange would have helped
develop and/or maintain the ecological advantage
for sponges within embayments. The facies distribution supports such an interpretation. Greater
inhibition of water exchange (sector 1 and palaeovalleys) led to greater sponge-richness (water
exchange inhibition obviously has an upper limit,
e.g. hyper- or hyposalinity, dysaerobic stagnation
would make benthic conditions uninhabitable, a
situation that has not been recorded in any
southern Australian Upper Eocene deposit). As
water exchange inhibition decreased (sector 3 and
outer archipelago areas of sector 1), calcareous
organisms could compete and a hybrid biota
developed that is essentially a record of embayment and open-marine water mixing, an inboard
situation equivalent to an oceanic mixing front
(Fig. 10). Calcareous epibenthos dominated in
Late Eocene open-shelf areas with no water
exchange inhibition (sector 4). Although embay-

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

580

P. R. Gammon and N. P. James

ment waters entering the open marine realm


became highly diluted, their effect can still be
seen in the unusually high siliceous sponge
numbers within Wilson Bluff calcareous marl.
The development and distribution of sponge-rich
facies within the Upper Eocene of southern Western Australia mirrors the hypothesized palaeogeographical inuence on the mixing of embayment
and open-shelf waters of differing chemistry.
The chemistry of these waters can at present be
only surmised, but lithistid and hexactinellid
sponges, as well as rare calcareous fossils, all
indicate normal marine salinities. The Late Eocene
hinterland was a low gradient plain (Sircombe &
Freeman, 1999), and run-off would have been
sluggish, resulting in chemical rather than physical weathering, low erosion rates and a negrained sediment supply (Twidale, 1982, 1994;
Ollier et al., 1988). Chemical weathering liberates
many nutrients, especially silica (this weathering
process is often termed desilicication; Bland &
Rolls, 1998). Siliceous sponges can only utilize
dissolved silica, and the low dissolved silica
contents of modern oceanic waters is interpreted
as a major nutrient limitation for sponges (Maldonado et al., 1999). This circumstantial evidence
suggests that the sponge-rich fauna may have been
strongly inuenced by nutrient-rich, especially
silica rich, terrestrial water (probably groundwater) entering the valleyembaymentarchipelago
palaeoenvironments. Marine areas receiving these
terrestrial waters became the sites of prolic
siliceous sponge growth, and from which calcareous biota were mostly excluded, presumably by
being unable to compete. As these terrestrial
waters were diluted by open-marine waters hybrid
siliceous sponge carbonate fossil assemblages
formed, with carbonate-dominated assemblages in
areas of predominantly open marine waters. The
siliceous sponge to calcareous biota ratio was
probably determined by the degree of terrestrial
open-marine water mixing.
Physical palaeogeography in siliceous sponge
proliferation was mainly important as a means to
inhibit water mixing between embayment and
open-marine water masses, but it also maintained
a physical hydrodynamic regime that is generally
similar to modern calm, low turbidity deep-sea
conditions, e.g. (a) oceanic swells were ameliorated to produce an overall relatively calm environment; (b) a low gradient hinterland delivered
only ne-grained siliciclastic material, at probably fairly low sedimentation rates (Gammon
et al., 2000a). Lithistid sponges within the Late
Eocene of southern Western Australia were able

to tolerate shoreface current regimes of 60 cm s1


velocity, and storm beds attest to periods of
enhanced turbidity. This suggests that the modern deep-water distribution of siliceous spongedominated habitats (van Soest & Stentoft, 1988;
Levi, 1991; Wiedenmayer, 1994) is probably not
primarily due to the low-energy, calm hydrodynamic conditions (see Rutzler & Macintyre, 1978;
Rutzler, 1990). Other factors such as nutrient
limitation and competition may be more important in determining the modern distribution of
siliceous sponges (see Reid, 1967; Reid, 1968;
Tabachnick, 1991; Wiedenmayer, 1994). Nevertheless, the Late Eocene palaeogeography produced embayment habitats that mimicked
the calm hydrodynamic conditions of modern
siliceous sponge-dominated environments, a
situation that could only have enhanced the
competitive advantage conferred on siliceous
sponges by favourable water chemistry.

Regional and global considerations


A restricted marine foraminiferal assemblage
indicates a palaeogeographical inuence on
muddy, calcareous spiculite deposition in the
Upper Eocene Blanche Point Formation of South
Australia's St. Vincent Basin (Fig. 1; McGowran &
Beecroft, 1986; James & Bone, 2000). Late Eocene
palaeovalleys to the north of and draining into the
Eucla Basin also contain spiculitic facies (Jones,
1990; Benbow et al., 1995). The Late Eocene
relationship between sponge-rich facies and protected palaeogeographical position is regional,
extending across 2000 km of the southern Australian Late Eocene coast. In each case barriers
inhibiting water exchange existed, supporting
the notion that palaeogeography, by inhibiting
water exchange with the open ocean, maintained sponge-promoting, nutrient-rich, nearshore waters.
Sponge-dominated deposition occurred throughout Mesozoic, Tethyan Europe, especially in
Late Jurassic and mid-Cretaceous time (Krautter,
1987; Reitner & Keupp, 1991; Pisera, 1997a,b).
Wilson Bluff Limestone sponge-rich calcareous
marls are similar to many of these Mesozoic
Tethyan deposits (see Gygi, 1986; Gygi & Persoz,
1987; Wiedenmayer, 1994), and both are interpreted as low-energy, open-shelf deposits. Late
Jurassic deposits in France accumulated in small,
active half-grabens situated between the Armorican and Cornubian Massifs, again suggestive of a
protected, complex palaeogeography (Rioult
et al., 1991). Rosales et al. (1995) documented

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation


an Albian siliceous sponge-dominated section
within an overall carbonate environment. They
speculated that the limited geographical extent of
the sponges was due to their close proximity to
the axial rift within an active extensional tectonic
system. They inferred that rotated, fault-bounded
blocks probably produced a complex palaeogeography and a resultant complex facies distribution.
The corollary is that the tectonic inuence may
have produced a palaeogeographical situation
that promoted water chemistries suitable for
sponge colonization.
Palaeoenvironments rich in siliceous sponges
have occurred sporadically through the Phanerozoic (e.g. Reid, 1967, 1968; Rigby & Stearn, 1983;
Narbonne & Dixon, 1984; Lang, 1991; Rigby, 1991;
Wiedenmayer, 1994; Pisera, 1999). These have
generally been interpreted as representing deep
and/or polar water environments (Wiedenmayer,
1994; Beauchamp & Desrochers, 1997), an interpretation that primarily reects the modern distribution of sponge-dominated environments (van
Soest & Stentoft, 1988; van Wagoner et al., 1989;
Conway et al., 1991; Levi, 1991). The Pallinup
Formation in Western Australia demonstrates
that sponges are not necessarily conned to
deeper waters. With a suitable substrate, hydrodynamic and nutrient regime, siliceous sponges
can be the dominant epibenthic organism within
shallow marine environments.
SUMMARY AND CONCLUSIONS
In southern Western Australia, during a single
Late Eocene transgressiveregressive cycle, a
variety of unusual sponge-rich facies were deposited in shallow, nearshore areas of the Bremer and
westernmost Eucla basins. The transgression
ooded a pre-existing embaymentarchipelago
that abutted a deeply weathered continental
regolith to the north. In the most landward
embayment shoreface environments, spongolites
and pure spiculites developed in response to
gentle hydrodynamic agitation, and the absence
of coarse terrigenous clastics. These grade laterally offshore to low-energy, subshoreface, muddy
spiculites and spiculitic mudstones that accumulated in central embayment areas. Calcareous
fossils are rare in these facies, but their number
increase markedly outboard as central embayment spiculite facies grade seaward to outer
archipelago calcareous spiculite. Open-shelf settings were the sites of locally sponge-rich, coolwater carbonate accumulation.

581

The sponge-rich facies distribution shows a


consistent relationship to the embaymentarchipelago palaeogeography, with increasing palaeogeographical protection resulting in increasingly
sponge-rich facies. The role of palaeogeography
was as a facilitator to other palaeoecological
controls such as sediment supply and water
chemistry. The palaeogeographical setting supplied three principal features: (1) a low gradient,
etchplain landscape with low-energy run-off that
delivered a nutrient-rich, ne-grained, suspended-load sediment supply that was within the
turbidity tolerance of sponges; (2) protected
nearshore environments that maintained a generally low-energy hydrodynamic regime; and (3)
limited water exchange with open-shelf waters,
thus preserving an embayment water chemistry
that promoted sponges. The transition from
embayment siliceous to open ocean calcareous
benthos is centred on the outermost protected
archipelago areas. This zone of mixing between
embayment and open-shelf waters supported a
hybrid faunal assemblage rich in both biosiliceous and calcareous epibenthos.
Similar Upper Eocene sponge-rich facies are
present across southern Australia in variably
protected nearshore palaeoenvironments. Other
older Mesozoic deposits elsewhere, although not
common, were deposited in comparable settings,
suggesting that palaeogeography was a critical
element in their formation.
Most interpretations of ancient sponge-rich
sediments point to a deep and/or cold water
origin. The Eocene sponge deposits of southern
Western Australia suggest that unrecognized
palaeoecological factors, such as palaeogeography, water chemistry, or shallow waters, may
have played important roles in at least some of
these older biosiliceous sediments.
ACKNOWLEDGEMENTS
The authors wish to thank the many southern
Western Australians who helped in a variety of
ways to make the summer 1998 eldwork a great
success. Without their interest and assistance a
lot of remote outcrops would probably have
remained unknown to the authors. We thank
our colleagues Drs Y. Bone and J. Clarke for their
support in both eldwork and the many discussions over the relative signicance of different
features of these deposits. Spiculite SEM studies
were undertaken in collaboration with Dr Andrjez
Pisera at the Polish Academy of Sciences in

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

582

P. R. Gammon and N. P. James

Warszawa. Many thanks to Dr I. Rosales and the


anonymous reviewer for their thoughtful comments that helped improve the manuscript. We
thank the Canadian National Science and Engineering Reseach Council for research support
(grant no. OGP0002028).
REFERENCES
Allen, J.R.L. (1982) Sedimentary Structures. Their Character
and Physical Basis. Elsevier, Amsterdam.
Barron, E.J. and Peterson, W.H. (1991) The Cenozoic ocean
circulation based on ocean General Circulation Model
results. Palaeogeogr. Palaeoclimatol. Palaeoecol., 83, 128.
Beauchamp, B. and Desrochers, A. (1997) Permian warm-to
very cold-water carbonates and cherts in northwest Pangea.
In: Cool-Water Carbonates (Eds N.P. James and J.D.A.
Clarke). SEPM Spec. Publ., 56, 327347.
Benbow, M.C., Alley, N.F., Callen, R.A. and Greenwood, D.R.
(1995) Geology and Paleoclimate. In: The Geology of South
Australia, Vol. 2 the Phanerozoic (Eds J.F. Drexel and
W.V. Preiss). S. Aust. Geol. Surv. Bull., 54, 208217.
Bland, W. and Rolls, D. (1998) Weathering. An Introduction to
the Scientic Principles. Arnold, London.
BMR Palaeogeographic Group. (1990) Australia: Evolution of
a Continent. Australian Government Publishing Service,
Canberra, Australia.
Bone, Y. and James, N.P. (1993) Bryozoans as carbonate
sediment producers on the cool-water Lacepede Shelf,
southern Australia. Sed. Geol., 86, 247271.
Bunting, J.A., van de Graaff, W.J.E. and Jackson, M.J. (1973)
Palaeodrainages and Cainozoic palaeogeography of the
Eastern Goldelds, Gibson Desert and Great Victoria Desert.
Bur. Mineral Resour. Annu. Report, 1973, 4550.
Carpenter, R.J. and Pole, M. (1995) Eocene plant fossils from
the Lefroy and Cowan Paleodrainages, Western Australia.
Aust. Syst. Bot., 8, 11071154.
Cavaroc, V.V.J. and Ferm, J.C. (1968) Siliceous spiculites as
shoreline indicators in deltaic sequences. Geol. Soc. Am.
Bull., 79, 263272.
Clarke, J.D.A. (1993) Stratigraphy of the Lefroy and Cowan
palaeodrainages, Western Australia. J. Roy. Soc. W. Aust.,
76, 1323.
Clarke, J.D.A. (1994) Evolution of the Lefroy and Cowan palaeodrainages, Western Australia. Aust. J. Earth Sci., 41, 5568.
Clarke, E.D.C. and Phillipps, H.T. (1952) Physiographic and
other notes on a part of the south coast of Western Australia.
J. Roy. Soc. W. Aust., 37, 5990.
Clarke, E.D.C. and Phillips, H.T. (1955) The Plantagenet Beds
of Western Australia. J. Roy. Soc. W. Aust., 39, 1926.
Clifton, H.E., Hunter, R.E. and Phillipps, R.L. (1971) Depositional structures and processes in the non-barred highenergy nearshore. J. Sed. Petrol., 41, 651670.
Cockbain, A.E. (1968a) The stratigraphy of the Plantagenet
Group, Western Australia. Geol. Surv. W. Aust. Annu.
Report, 1967, 99101.
Cockbain, A.E. (1968b) Eocene foraminifera from the Norseman Limestone of Lake Cowan, Western Australia. Geol.
Surv. W. Aust. Annu. Report, 1967, 5960.
Collins, M.B. and Banner, F.T. (1980) Sediment transport by
waves and tides: problems exemplied by a study of
Swansea Bay, Bristol Channel. In: The Northwest European

Shelf Seas: the Sea Bed and the Sea in Motion (Eds F.T.
Banner, M.B. Collins and K.S. Massie), pp. 369388. Elsevier Oceanography Series 24B, Amsterdam.
Conway, K.W., Barrie, J.V., Austin, W.C. and Luternauer, J.L.
(1991) Holocene sponge bioherms on the western Canadian
continental shelf. Cont. Shelf Res., 11, 771790.
Cope, R.N. (1975) Tertiary epeirogeny in the southern part of
Western Australia. Geol. Surv. W. Aust. Annu. Report, 1974,
4046.
Darragh, T.A. and Kendrick, G.W. (1980) Eocene bivalves
from the Pallinup Siltstone near Walpole, Western Australia. J. Roy. Soc. W. Aust., 63, 520.
Davis, R.A., Klay, J. and Jewell, P. (1993) Sedimentology and
stratigraphy of tidal sand ridges southwest Florida inner
shelf. J. Sed. Petrol., 63, 91104.
Dayton, P.K., Robilliard, G.A., Paine, R.T. and Dayton, L.B.
(1974) Biological accommodation in the benthic community at McMurdo Sound, Antarctica. Ecol. Monogr., 44,
105128.
Droser, M.L. and Bottjer, D.J. (1986) A semiquantitative eld
classication of ichnofabric. J. Sed. Petrol., 56, 558559.
Feary, D.A. and James, N.P. (1998) Seismic stratigraphy and
geological evolution of the Cenozoic, cool-water Eucla
Platform, Great Australian Bight. AAPG Bull., 82, 792816.
Feary, D.A., James, N.P. and McGowran, G. (1994) Cenozoic
Cool-water Carbonates of the Great Australian Bight:
Reading the Record of Southern Ocean evolution, Sea-level,
Paleoclimate, and Biotic evolution. Australian Geological
Survey Organisation Record 1994/62, Australian Geological
Survey Organisation, Canberra, Australia.
Gammon, P.R., James, N.P., Bone, Y. and Clarke, J.D.A.
(2000a) Sedimentology and lithostratigraphy of a Late
Eocene sponge-dominated sequence, southern Western
Australia. Aust. J. Earth Sci., 47, 10871103.
Gammon, P.R., James, N.P. and Pisera, A. (2000b) Late Eocene
shallow-marine sponge deposits, Western Australia. Geology, 28, 855858.
Golonka, J., Ross, M.I. and Scotese, C.R. (1994) Phanerozoic
paleogeographic and paleoclimatic modelling maps. In:
Pangea: Global Environments and Resources (Eds A.F.
Embry, B. Beauchamp and D.J. Glass). Can. Soc. Petrol.
Geol. Mem., 17, 147.
van de Graaff, W.J.E. (1981) Paleogeographic evolution of a
rifted cratonic margin: S.W. Australia, a discussion. Palaeogeogr. Palaeoclimatol. Palaeoecol., 34, 163172.
van de Graaff, W.J.E., Crowe, R.W.A., Bunting, J.A. and
Jackson, M.J. (1977) Relict Early Cainozoic drainages in arid
Western Australia. Z. Geomorphol., 21, 379400.
Gygi, R.A. (1986) Eustatic sea level changes of the Oxfordian
(Late Jurassic) and their effect documented in sediments
and fossil assemblages of an epicontinental sea. Eclogae
Geol. Helv., 79, 455491.
Gygi, R.A. and Persoz, F. (1987) The epicontinental sea of
Swabia (southern Germany) in the Late Jurassic factors
controlling sedimentation. Neues Jb. Geol. Palaontol. Abh.,
176, 4965.
Henrich, R., Hartmann, M., Reitner, J., Schafer, P., Freiwald,
A., Steinmets, S. et al. (1992) Facies belts and communities
of the arctic Vesterisbanken Seamount (Central Greenland
Sea). Facies, 27, 71104.
Hinde, G.J. (1910) On fossil sponge spicules in a rock from
the Deep Lead (?) at Princess Royal Township, Norseman
District, Western Australia. Geol. Surv. W. Aust. Bull., 36,
724.

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Late Eocene sponge-rich sedimentation


Hydrographic Service, R.A.N. (1995) 1995 Australian
National Tide Tables Digital Edition. Flinders University,
Adelaide, Australia.
Ilan, M. and Abelson, A. (1995) The life of a sponge in a sandy
lagoon. Biological Bulletin, 189, 363369.
James, N.P. (1997) The cool-water carbonate depositional
realm. In: Cool-Water Carbonates (Eds N.P. James and J.D.A.
Clarke). SEPM Spec. Publ., 56, 120.
James, N.P. and von der Borch, C.C. (1991) Carbonate shelf
edge off southern Australia: a prograding open-platfrom
margin. Geology, 19, 10051008.
James, N.P. and Bone, Y. (1994) Paleoecology of cool-water,
subtidal cycles in mid-Cenozoic limestones, Eucla Platform,
Southern Australia. Palaios, 9, 457476.
James, N.P. and Bone, Y. (2000) Eocene cool-water carbonate
and biosiliceous sedimentation dynamics, St. Vincent
Basin, South Australia. Sedimentology, 47, 761786.
James, N.P., Bone, Y., Von Der Borch, C.C. and Gostin, V.A.
(1992) Modern carbonate and terrigenous clastic sediments
on a cool water, high energy, mid-latitude shelf: Lacepede,
southern Australia. Sedimentology, 39, 877903.
Janecek, T.R. and Rea, D.K. (1983) Eolian deposition in the
northeast Pacic Ocean: Cenozoic history of atmospheric
circulation. Geol. Soc. Am. Bull., 94, 730738.
Jones, B.G. (1990) Cretaceous and Tertiary sedimentation on
the western margin of the Eucla Basin. Aust. J. Earth Sci.,
37, 317329.
Jones, J.B. and Fitzgerald, M.J. (1986) Silica-rich layering at
Blanche Point, South Australia. Aust. J. Earth Sci., 33,
529551.
Kemp, E.M. (1978) Tertiary climatic evolution and vegetation
history in the southeastern Indian Ocean region. Palaeogeogr. Palaeoclimatol. Palaeoecol., 24, 169208.
Kennett, J.P. and Warnke, D.A. (1992) The Antarctic paleoenvironment: a perspective on global change. Part one. In:
Antarctic Research Series, Vol. 56, pp. 385. Am. Geophys.
Union, Washington, D.C.
Kidwell, S.M. (1991) Taphonomic feedback (live/dead
interactions) in the genesis of bioclastic beds: keys to
reconstructing sedimentary dynamics. In: Cycles and
Events in Stratigraphy (Eds G. Einsele, W. Ricken and
A. Seilacher), pp. 268282. Springer-Verlag, Berlin, Heidelberg.
Kidwell, S.M. and Holland, S.H. (1991) Field description of
coarse bioclastic fabrics. Palaios, 6, 426434.
Koltun, V.M. (1970) Sponges of the Arctic and Antarctic; A
faunistic review. In: The Biology of the Porifera (Ed. W.G.
Fry). Symposia of the Zoological Society of London, 25,
285297.
Kondo, Y., Matsui, S. and Chinzei, K. (1994) Taphonomy and
Paleoecology of the Pleistocene molluscs in the Boso Peninsula. 29th International Geological Conference Field Trip
Guide B22, Geol. Surv. Japan; Ibaraki, Japan.
Krautter, M. (1987) Aspekte aur Palaokologie postpalaozoische Kieselschwamme. Prol, 11, 133324.
Lane, N.G. (1981) A nearshore sponge spicule mat from the
Pennsylvanian of west-central Indiana. J. Sed. Petrol., 51,
197202.
Lang, B. (1991) Bafing, binding, or debris accumulation?
Ecology of Upper Jurassic sponge-bacterial buildups
(Oxfordian, Franconian Alb, Southern Germany). In: Fossil
and Recent Sponges (Eds J. Reitner and H. Keupp),
pp. 516521. Springer-Verlag, Berlin, Heidelberg.
de Laubenfels, M.W. (1953) Fossil sponges of Western Australia. J. Roy. Soc. W. Aust., 37, 105117.

583

Lawver, L.A., Gahagan, L.M. and Cofn, M.F. (1992) The


development of paleoseaways around Antarctica. In: The
Antarctic Paleoenvironment: a Perspective on Global
Change (Eds J.P. Kennett and D.A. Warnke). Antarctic
Research Series, 56, 730.
Leinfelder, R.R., Krautter, M., Laternser, R., Nose, M., Schmid, D.U., Schweigert, G. et al. (1994) The origin of Jurassic
reefs: current research developments and results. Facies, 31,
156.
Levi, C. (1991) Lithistid sponges from the Norfolk Rise. Recent
and Mesozoic genera. In: Fossil and Recent Sponges (Eds
J. Reitner and H. Keupp), pp. 7282. Springer-Verlag,
Berlin, Heidelberg.
Li, Q., James, N.P., Bone, Y. and McGowran, B. (1996) Foraminiferal biostratigraphy and depositional environments
of the mid-Cenozoic Abrakurrie Limestone, Eucla Basin,
southern Australia. Aust. J. Earth Sci., 43, 437450.
Lowry, D.C. (1970) Geology of the Western Australian part of
the Eucla Basin. Geol. Surv. W. Aust. Bull., 122, 201.
Maldonado, M., Carmona, M.C., Uriz, M. and Cruzado, A.
(1999) Decline in Mesozoic reef-building sponges explained
by silicon limitation. Nature, 401, 785788.
McBride, E.F. and Folk, R.L. (1977) The Caballos Novaculite
revisited. Part 2; Chert and shale members and synthesis.
J. Sed. Petrol., 47, 12611286.
McGowran, B. (1987) Late Eocene perturbations: foraminiferal
biofacies and evolutionary overturn, southern Australia.
Paleoceanography, 2, 715727.
McGowran, B. (1989a) Silica burp in the Eocene Ocean.
Geology, 17, 857860.
McGowran, B. (1989b) The later Eocene transgression in
southern Australia. Alcheringa, 13, 4568.
McGowran, B. and Beecroft, A. (1986) Foraminiferal biofacies in a silica-rich neritic sediment, Late Eocene, South
Australia. Palaeogeogr. Palaeoclimatol. Palaeoecol., 52,
321345.
McGowran, B., Li, Q. and Moss, G. (1997) The Cenozoic
neritic record in southern Australia: the biogeohistorical
framework. In: Cool-Water Carbonates (Eds N.P. James and
J.D.A. Clarke). SEPM Spec. Publ., 56, 185204.
Meldahl, K.H. (1993) Geographic gradients in the formation of
shell concentrations: Plio-Pleistocene marine deposits, Gulf
of California. Palaeogeogr. Palaeoclimatol. Palaeoecol., 101,
125.
Narbonne, G.M. and Dixon, O.A. (1984) Upper Silurian lithistid sponge reefs on Somerset Island, Arctic Canada. Sedimentology, 31, 2550.
Nittrouer, C.A. and Wright, L.D. (1994) Transport of particles
across continental shelves. Rev. Geophys., 32, 85113.
Nummedal, D. (1991) Shallow marine storm sedimentation
the oceanographic perspective. In: Cycles and Events in
Stratigraphy (Eds G. Einsele, W. Ricken and A. Seilacher),
pp. 227248. Springer-Verlag, Berlin, Heidelberg.
Odin, G.S. (1988) Green Marine Clays. Oolitic Ironstone Facies,
Verdine Facies, Glaucony Facies and Celadonite-Bearing
Facies a Comparative Study. Elsevier, Amsterdam.
Ollier, C.D., Chan, R.A., Craig, M.A. and Gibson, D.L. (1988)
Aspects of landscape history and regolith in the Kalgoorlie
region, Western Australia. Bur. Mineral Resour. J. Aust.
Geol. Geophys., 10, 309321.
Pemberton, S.G., MacEachern, J.A. and Frey, R.W. (1992)
Trace fossil facies models: Environmental and allostratigraphic signicance. In: Facies Models: Response to SeaLevel Change (Ed. R.G. Walker and N.P. James), pp. 4772.
Geol. Assoc. Can., St. Johns, Newfoundland.

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

584

P. R. Gammon and N. P. James

Pickett, J.W. (1982) Vacelatia progenitor, the rst Tertiary


sphinctozoan (Porifera). Alcheringa, 6, 241247.
Pickett, J.W. (1983) An annotated bibliography and review of
Australian fossil sponges. Assoc. Australas. Paleontol.
Mem., 1, 93120.
Pisera, A. (1991) Upper Jurassic sponge megafacies in Spain:
preliminary report. In: Fossil and Recent Sponges (Eds
J. Reitner and H. Keupp), pp. 486497. Springer-Verlag,
Berlin, Heidelberg.
Pisera, A. (1997a) Upper Jurassic Siliceous Sponges from the
Swabian Alb: Taxonomy and Paleoecology. Publishing
Department of the Institute of Paleobiology, Warszawa.
Pisera, A. (1997b) Hexactinellid sponges of the Admiralty Bay,
King George Island, Antarctica. In: Polish Polar Studies. 24th
Polar Symposium (Ed. P. Glowacki), pp. 203205. Institute of
Geophysics of the Polish Academy of Sciences, Warszawa.
Pisera, A. (1999) Post-Paleozoic history of the siliceous
sponges with rigid skeleton. (Ed. J.N.A. Hooper), Queensland Mus. Mem., 44, 463472, Brisbane.
Prothero, D.R. (1994) The Eocene-Oligocene Transition:
Paradise Lost. Columbia University Press, New York.
Quilty, P.G. (1969) Upper Eocene planktonic Foraminiferida
from Albany, Western Australia. J. Roy. Soc. W. Aust., 52,
4158.
Reid, R.E.H. (1967) Tethys and the zoogeography of some
modern and Mesozoic Porifera. In: Aspects of Tethyan
Biogeography (Eds D.V. Ager and C.G. Adams), pp. 171
181. The Systematics Association, London.
Reid, R.E.H. (1968) Bathymetric distributions of Calcarea and
Hexactinellida in the present and past. Geol. Mag., 105,
546559.
Reiswig, H.M. (1990) In situ feeding in two shallow-water
hexactinellid sponges. In: New Perspectives in Sponge
Biology (Ed. K. Rutzler), pp. 504510. Smithsonian Institution Press, Washington, London.
Reitner, J. and Keupp, H. (1991) Fossil and Recent Sponges,
pp. 578. Springer-Verlag, Berlin, Heidelberg.
Rigby, J.K. (1991) Evolution of Paleozoic heteractinid calcareous sponges and demosponges patterns and records. In:
Fossil and Recent Sponges (Eds J. Reitner and H. Keupp),
pp. 83101. Springer-Verlag, Berlin, Heidelberg.
Rigby, J.K. and Stearn, C.W. (1983) Sponges and Spongiomorphs. University of Tennessee, Indianapolis.
Rioult, M., Dugue, O., du Chene, J., Ponsot, C., Fily, G.,
Moron, J.-M. et al. (1991) Outcrop sequence stratigraphy of
the Anglo-Paris Basin, Middle to Upper Jurassic (Normandy, Maine, Dorset). Bull. Centres Rech. Explor.-Prod.
Elf-Aquitaine, 15, 101194.
Rosales, I., Mehl, D., Fernandez-Mendiola, P.A. and GarcaMondejar, J. (1995) An unusual poriferan community in the
Albian of Islares (north Spain): paleoenvironmental and
tectonic implications. Palaeogeogr. Palaeoclimatol. Palaeoecol., 119, 4761.
Rutzler, K. (1990) New perspectives in sponge biology. In:
Third International Conference on the Biology of Sponges,
Woods, Hole, 1985, pp. 533. Smithsonian Institution Press,
Washington, London.
Rutzler, K. and Macintyre, I.G. (1978) Siliceous sponge spicules in coral reef sediments. Mar Biol., 49, 147159.
Shackleton, N. and Kennett, J.P. (1975) Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation:
Oxygen and carbon isotope analyses in DSDP sites 277, 279,
281. Init. Rep. Deep Sea Drilling Proj., 29, 801807.
Sircombe, K.N. and Freeman. M.J. (1999) Provenance of
detrital zircons on the Western Australia coastline

Implications for the geologic history of the Perth basin and


denudation of the Yilgarn craton. Geology, 27, 827882.
van Soest, R.W.M. and Stentoft, N. (1988) Barbados DeepWater Sponges. Foundation for Scientic Research in
Surinam and the Netherlands Antilles, Amsterdam.
Swift, D.J.P., Hudelson, P.M., Brenner, R.L. and Thompson, P.
(1987) Shelf construction in a foreland basinstorm beds,
shelf sandbodies, and shelf-slope depositional sequences in
the Upper Cretaceous mesaverde Group, Book Cliffs, Utah.
Sedimentology, 34, 423457.
Swift, D.J.P., Oertel, G.F., Tillman, R.W. and Thorne, J.A.
(1991) Shelf Sand and Sandstone Bodies: Geometry, Facies
and Sequence Stratigraphy. Int. Assoc. Sedimentol. Spec.
Publ., 14, 532.
Tabachnick, K.R. (1991) Adaptation of the hexactinellid
sponges to deep-sea life. In: Fossil and Recent Sponges (Eds
J. Reitner and H. Keupp), pp. 378386. Springer-Verlag,
Berlin, Heidelberg.
Thom, R. and Chin, R.J. (1984) Explanatory notes on the
Bremer Bay geological sheet. In: Geol. Surv. W. Aust.
1:250,000 Geological Series, Sheet SI/5012. Geol. Surv. W.
Aust., Perth, Australia.
Trammer, J. (1991) Ecologic history of the Oxfordian sponge
assemblage in the Polish Jura. In: Fossil and Recent Sponges
(Eds J. Reitner and H. Keupp), pp. 506515. Springer-Verlag, Berlin, Heidelberg.
Twidale, C.R. (1982) Role of subterranean water in landform
development in tropical and subtropical regions. In:
Groundwater as a Geomorphic Agent (Ed. R.G. LaFleur),
pp. 91134. Allen & Unwin, Winchester.
Twidale, C.R. (1994) Gondwanan (Late Jurassic and Cretaceous) palaeosurfaces of the Australian craton. Palaeogeogr.
Palaeoclimatol. Palaeoecol., 112, 157186.
Twidale, C.R. and Bourne, J.A. (1998) The use of duricrusts
and topographic relationships in geomorphological correlation: conclusions based in Australian experience. Catena,
33, 105122.
Veevers, J.J. (1984) Phanerozoic Earth History of Australia.
Clarendon Press, Oxford.
van Wagoner, N.A., Mudie, P.J., Cole, F.E. and Daborn, G.
(1989) Siliceous sponge communities, biological zonation,
and recent sea-level change on the Arctic margin; Ice Island
results. Can. J. Earth Sci., 26, 23412355.
Walker, R.G. and Plint, A.G. (1992) Wave- and storm-dominated shallow marine systems. In: Facies Models: Response
to Sea-Level Change (Eds R.G. Walker and N.P. James),
pp. 219238. Geol. Assoc. Can., St. Johns, Newfoundland.
Weast, R.C., Tuve, G.L., Selby, S.M. and Sunshine, I. (1972)
Handbook of Chemistry and Physics. The Chemical Rubber
Co., Cleveland.
Wiedenmayer, F. (1994) Contributions to the Knowledge of
Post-Palaeozoic Neritic and Archibenthal Sponges (Porifera). Kommission der Schweizerischen Palaontologischen
Abhandlungen, Basel.
Wise, S.W.J. and Weaver, F.M. (1973) Origin of cristobaliterich Tertiary sediments in the Atlantic and Gulf Coastal
Plain. Trans. Gulf Coast Assoc. Geol. Soc., 23, 305323.
Wright, L.D. (1995) Morphodynamics of Inner Continental
Shelves. CRC Press, Boca Raton, Ann Arbor.

Manuscript received 19 January 2000;


revision accepted 25 September 2000.

2001 International Association of Sedimentologists, Sedimentology, 48, 559584

Vous aimerez peut-être aussi