100% found this document useful (1 vote)
354 views218 pages

(Oscar R. López-Gamundí, Luis A. Buatois) Late P PDF

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
354 views218 pages

(Oscar R. López-Gamundí, Luis A. Buatois) Late P PDF

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 218

B THE GEOLOGICAL SOCIETY

Special Paper468 - OF AMERICA®

Late Paleozoic GRaciall Event§


a n dl Po§tgRadiaR Tiran§gire§§fton §
lin G ondlwan a

India~

Africa Antarctica
Late Paleozoic Glacial Events and Postglacial
Transgressions in Gondwana

edited by

Oscar R. López-Gamundí
Hess Corporation
One Allen Center
500 Dallas Street
Houston, Texas 77002
USA

and

Luis A. Buatois
Department of Geological Sciences
University of Saskatchewan
114 Science Place
Saskatoon, SK S7N 5E2
Canada

Special Paper 468


3300 Penrose Place, P.O. Box 9140 Boulder, Colorado 80301-9140, USA

2010
Copyright © 2010, The Geological Society of America (GSA), Inc. All rights reserved. GSA grants
permission to individual scientists to make unlimited photocopies of one or more items from this volume
for noncommercial purposes advancing science or education, including classroom use. For permission
to make photocopies of any item in this volume for other noncommercial, nonprofit purposes, contact
The Geological Society of America. Written permission is required from GSA for all other forms of
capture or reproduction of any item in the volume including, but not limited to, all types of electronic
or digital scanning or other digital or manual transformation of articles or any portion thereof, such
as abstracts, into computer-readable and/or transmittable form for personal or corporate use, either
noncommercial or commercial, for-profit or otherwise. Send permission requests to GSA Copyright
Permissions, 3300 Penrose Place, P.O. Box 9140, Boulder, Colorado 80301-9140, USA. GSA provides
this and other forums for the presentation of diverse opinions and positions by scientists worldwide,
regardless of their race, citizenship, gender, religion, or political viewpoint. Opinions presented in this
publication do not reflect official positions of the Society.

Copyright is not claimed on any material prepared wholly by government employees within the scope of
their employment.

Published by The Geological Society of America, Inc.


3300 Penrose Place, P.O. Box 9140, Boulder, Colorado 80301-9140, USA
www.geosociety.org

Printed in U.S.A.

GSA Books Science Editors: Marion E. Bickford and Donald I. Siegel

Library of Congress Cataloging-in-Publication Data

Late Paleozoic glacial events and postglacial transgressions in Gondwana / edited by Oscar R. López-
Gamundí and Luis A. Buatois.
p. cm. — (Special paper ; 468)
Includes bibliographical references.
ISBN 978-0-8137-2468-3 (pbk.)
1. Geology, Stratigraphic—Paleozoic. 2. Glacial landforms—Gondwana (Continent) 3. Periglacial
processes—Gondwana (Continent) 4. Gondwana (Continent) I. López-Gamundí, Oscar R. II. Buatois,
Luis A.
QE654.L375 2010
551.7′2—dc22
2010019884

Cover, front: Map showing the distribution of glacial basins in Gondwana. From Isbell, J.L.,
“Environmental and paleogeographic implications of glaciotectonic deformation of glaciomarine
deposits within Permian strata of the Metschel Tillite, southern Victoria Land, Antarctica” (Chapter 3,
this volume). Back: Postglacial transgressive scenarios. (Upper left) Fjord environment influenced
by extreme freshwater discharge from retreating glaciers. A freshwater ichnofauna occurs in the
transgressive deposits (Guandacol Formation). (Lower right) Coastal environment without direct
influence of freshwater influx from melting ice masses. A freshwater ichnofauna is present in coastal
lakes and temporally inundated flood plains (Trace-Fossil Assemblage 1). A brackish-water ichnofauna
(Trace-Fossil Assemblages 2 and 3) occurs in distal-bay facies (Tupe Formation). From Desjardins, P.R.,
Buatois, L.A., Mángano, M.G., and Limarino, C.O., “Ichnology of the latest Carboniferous–earliest
Permian transgression in the Paganzo Basin of western Argentina: The interplay of ecology, sea-level rise,
and paleogeography during postglacial times in Gondwana” (Chapter 8, this volume).

10 9 8 7 6 5 4 3 2 1
Contents

Introduction: Late Paleozoic glacial events and postglacial transgressions in Gondwana . . . . . . . . . . v


Oscar R. López-Gamundí and Luis A. Buatois

1. Transgressions related to the demise of the Late Paleozoic Ice Age: Their sequence
stratigraphic context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Oscar R. López-Gamundí

2. From bergs to ergs: The late Paleozoic Gondwanan glaciation and its aftermath in
Saudi Arabia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
John Melvin, Ronald A. Sprague, and Christian J. Heine

3. Environmental and paleogeographic implications of glaciotectonic deformation


of glaciomarine deposits within Permian strata of the Metschel Tillite, southern
Victoria Land, Antarctica . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
John L. Isbell

4. Formation of euxinic lakes during the deglaciation phase in the Early Permian
of East Africa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Thomas Kreuser and Gebretinsae Woldu

5. Stratigraphic and paleofloristic record of the Lower Permian postglacial succession


in the southern Brazilian Paraná Basin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
Roberto Iannuzzi, Paulo A. Souza, and Michael Holz

6. “Levipustula Fauna” in central-western Argentina and its relationships with the


Carboniferous glacial event in the southwestern Gondwanan margin . . . . . . . . . . . . . . . . . . . . 133
Gabriela A. Cisterna and Andrea F. Sterren

7. Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana: Reconstructing


salinity conditions in coastal ecosystems affected by strong meltwater discharge . . . . . . . . . . . . 149
Luis A. Buatois, Renata G. Netto, and M. Gabriela Mángano

8. Ichnology of the latest Carboniferous–earliest Permian transgression in the Paganzo


Basin of western Argentina: The interplay of ecology, sea-level rise, and paleogeography
during postglacial times in Gondwana . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Patricio R. Desjardins, Luis A. Buatois, M. Gabriela Mángano, and Carlos O. Limarino

9. Reconstruction of a high-latitude, postglacial lake: Mackellar Formation (Permian),


Transantarctic Mountains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
Molly F. Miller and John L. Isbell

iii
The Geological Society of America
Special Paper 468
2010

Introduction: Late Paleozoic glacial events and


postglacial transgressions in Gondwana

Oscar R. López-Gamundí
Hess Corporation, 500 Dallas Street, Houston, Texas 77002, USA

Luis A. Buatois
Department of Geological Sciences, University of Saskatchewan, 114 Science Place, Saskatoon, SK S7N 5E2, Canada

The stratigraphic record suggests that glaciations have fied subsequently by the seminal work led by John Crowell and
occurred episodically at different time intervals in Earth’s his- Lawrence Frakes (Frakes and Crowell, 1967, 1969; Frakes et al.,
tory (Crowell, 1982, 1999). One of those glaciations affected 1969; Frakes and Crowell, 1970; Crowell and Frakes, 1971a,
the Gondwanan Supercontinent during the late Paleozoic and 1971b, 1972; Frakes et al., 1971; Crowell, 1978). With the addi-
constituted the longest period of continuous glaciation in the tional help of later contributions, the body of evidence about the
Phanerozoic (Eyles, 1993). Carboniferous to Early Permian gla- duration, areal extent, and influence of the Late Paleozoic Ice
ciogenic successions have been known on all the subcontinents Age (LPIA) on the biota has significantly grown.
of Gondwana, most notably South America, Africa, India, and However, uncertainty remains over the exact timing of onset
Australia, and later work expanded to Antarctica and the Middle and demise of each glacial episode of the LPIA, particularly
East (Fig. 1). This glacial age can be subdivided into three dis- when attempts are made to link these glacial episodes in Gond-
tinct episodes (López-Gamundí, 1997). Glacial episodes II and wana with cyclothems in the Northern Hemisphere (particularly
III occurred during the early Late Carboniferous and the Late the United States and Europe), following Wanless and Shepard’s
Carboniferous–Early Permian, respectively. An earlier, short- (1936) hypothesis. The picture becomes particularly blurred if,
lived glacial episode in the Late Devonian–earliest Carbonifer- as exemplified by Wright and Vanstone (2001) for the Viséan car-
ous (glacial episode I) identified in central and northern South bonate successions in the Northern Hemisphere (UK), glacioeu-
America (Fig. 1) extended even further the duration of this ice static sea-level oscillations invoked to account for high-frequency
age (Veevers and Powell, 1987). In general, the locus of ice cover, cyclicity had an approximate 100 ka periodicity, which may cor-
and its stratigraphic record, progressively moved across Gond- respond to Milankovitch eccentricity. Thus, far field studies can
wana from South America to Australia (Crowell, 1999), tracking sometimes be based on shaky grounds, particularly owing to the
the transpolar trajectory across Gondwana. This polar wander difficulty of estimating the magnitude and hierarchy of the near
across the Gondwanan Supercontinent controlled paleolatitudes field glacioeustatic fluctuations (Rygel et al., 2008) and the less
and accounts for the diachroneity of glacial episodes I, II, and than optimal chronostratigraphic resolution of the Gondwanan
III; however, the exact timing of waxing and waning of ice cen- faunas and floras associated with the LPIA.
ters during each glacial episode (particularly for the longest-lived Maximum expansion of Gondwanan continental ice sheets
episode III) seemed to have been influenced by basin dynamics, occurred during earliest Permian time (glacial episode IV) under
topographic barriers, glaciation styles, and other local factors. paleoatmospheric CO2 levels as low as the present ones to values
The recognition of ancient glacial deposits of similar late of up to 12 times higher by the late Early Permian (Montañez
Paleozoic age in South Africa and South America (Du Toit, 1927) et al., 2007). Widespread Early Permian (Sakmarian) collapse of
helped, in conjunction with other lines of evidence, to argue in ice sheets coincided with the onset of rising atmospheric CO2 lev-
favor of the principles of seafloor spreading and indirectly to els, after which time surface temperatures and atmospheric partial
build the theory of plate tectonics (Wegener, 1915). The pio- pressure of carbon dioxide (pCO2) rose. The more detailed and
neering work during the first half of the last century was solidi- deeper our knowledge about the LPIA gets, the better positioned
López-Gamundí, O.R., and Buatois, L.A., 2010, Introduction: Late Paleozoic glacial events and postglacial transgressions in Gondwana, in López-Gamundí, O.R.,
and Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana: Geological Society of America Special Paper 468, p. v–viii,
doi: 10.1130/2010.2468(00). For permission to copy, contact [email protected]. ©2010 The Geological Society of America. All rights reserved.

v
vi López-Gamundí and Buatois

Glacial Episode III

Arabia
Glacial Episode II
2
Glacial Episode I
A F R I C A
340 Ma
Polar Path
360 Ma
340 Ma
INDIA

1
360 Ma

4
S O U T H AUSTRALIA
320 Ma
5
7 A N T A R C T I C A
A M E R I C A
1
7 8 1 7 280 Ma
250 Ma
7 7
260 Ma 1
265 Ma

1 6 3 7 9
0 2000 km

Figure 1. Gondwana Supercontinent and the Late Paleozoic Ice Age (LPIA) basins with glacial evidence in their strati-
graphic record highlighted. Glacial episodes after López-Gamundí (1997); polar path after Powell and Li (1994). Loca-
tions of contributions in this volume indicated by numbers corresponding to chapters in the volume.

we will be to understand the relationship between shifts in pCO2, (3) the chronostratigraphic resolution of paleofaunas (Cisterna
temperature, and ice volume and greenhouse gas forcing of past and Sterren, Chapter 6) and paleofloras (Ianuzzi et al., Chapter
and future climates. Additionally, a better understanding of the 5) that coexisted with extreme glacial and relatively milder early
mechanisms of postglacial transgressions along basin margins postglacial conditions, and the presence of freshwater and brack-
will allow us to refine our search for fossil fuels through the iden- ish water ichnofaunas related to postglacial marine transgres-
tification of potential marine source rocks and coals. We hope sions (Buatois et al., Chapter 7; Desjardins et al., Chapter 8);
this volume contributes to both ends. (4) the characterization of high-latitude, postglacial lakes (Miller
This volume’s contributions constitute a wide range of top- and Isbell, Chapter 9); and (5) the search for a unifying sequence-
ics related to this extreme paleoclimatic episode in Earth’s his- stratigraphic model for the glacial-postglacial transition (López-
tory. Original presentations were part of the IGCP 471 project Gamundí, Chapter 1).
(“Evolution of the western Gondwana during the late Paleo- The contributions included in this volume cover a broad
zoic: Tectonosedimentary record, paleoclimate and biological geography across Gondwana, but they do not have the objective
changes”)–sponsored session Late Paleozoic glacial events and of giving a state-of-the-art review of the LPIA, a theme that has
postglacial transgressions in Gondwana during the 32nd Inter- been periodically dealt with since Hambrey and Harland’s (1981)
national Geological Congress (Florence, August 2004). It was volume. A recent update can be found in Fielding et al. (2008).
evident at that time that consensus had been reached on some Rather, the present volume is focused on key specific topics related
basic problems about the LPIA, but some new challenges had to the LPIA that, in our opinion, required further study. These top-
emerged. Some of the unanswered questions that this volume ics deal with the two main episodes identified during the LPIA:
attempts to address revolve around (1) relatively less known the early Late Carboniferous (Namurian–Westphalian) glacial
glacial deposits in some regions of Gondwana (Central Africa, episode II, mostly confined to the Paleopacific margin of south-
Kreuser and Woldu, Chapter 4; Arabia, Melvin et al., Chapter 2); ern South America, and a much more widespread early Permian
(2) the controversy between a single massive ice sheet versus glacial episode III, which affected the rest of the supercontinent
numerous glacial centers and alpine glaciers (Isbell, Chapter 3); (López-Gamundí, 1997; Isbell et al., 2003). Instead of providing
Introduction vii

brief summaries of specific areas, the authors were encouraged to outline the regional extent of these anoxic lakes in various late
expand their views, providing full documentation. Paleozoic basins of eastern and southern Africa.
The book consists of two main parts. The first half deals with The nature of continental and shallow-marine ecosystems
sedimentologic, paleoenvironmental, and paleoclimatic aspects of during glacial and postglacial times is still poorly understood. The
the glacial event. The second half explores paleobiologic aspects last five contributions of this book focus on this topic, touching
of glacial and glacially influenced ecosystems. The first contribu- also on biostratigraphic implications. In Chapter 5, Iannuzzi et al.
tion, which is by López-Gamundí, focuses on the sequence stra- provide for the first time a sequence-stratigraphic and paleoen-
tigraphy of the late Paleozoic glacial event and the subsequent vironmental framework for palynozones and plant zones in the
postglacial phase, setting the stage for the rest of the volume. He Rio Grande do Sul portion of the Paraná Basin. These authors
notes that, irrespective of the age, there is a common stratigraphic find that the boundaries of palynozones lie near the maximum
stacking pattern in each of the transgressive events. As a result flooding surfaces. In addition, they note that the plant zones pre-
of the combined effect of fast glacioeustatic sea-level rise and viously defined correspond to distinct ecofacies and are better
subsidence along basin margins, a drastic landward facies shift regarded as ecozones rather than as biozones. Based on this anal-
took place in the transition from glacially dominated to glacially ysis, a link is suggested between the increase in floral diversity of
influenced early postglacial environments. Available information the Glossopteris-Rhodeopteridium Zone and the appearance of
allows recognition of two basic types of transgressive systems more complex coastal ecosystems as recorded in the Rio Bonito
tracts (TSTs): (1) a complete TST, with a basal diamictite unit, Formation. Cisterna and Sterren (Chapter 6) evaluate taphonomic
followed by shale with ice-rafted debris (IRD) and IRD-free and paleoecologic aspects of the Levipustula Fauna. This fauna
shales, culminating in a maximum flooding surface; and (2) a is typical of lower Upper Carboniferous glacially related deposits
base-cut TST in which the TST is dominated by open-marine in the Andean basins of Argentina. Based on studies in differ-
shales, generally devoid of IRD. ent stratigraphic units of the Calingasta-Uspallata Basin, these
The other three contributions are case studies based on authors are able to distinguish two associations: intraglacial and
specific areas of Gondwana but bearing implications at a more postglacial. They also note that the postglacial association is
global scale. Melvin et al. (Chapter 2) provide a detailed charac- more diverse and displays more abundance than the intraglacial
terization of the Upper Carboniferous–Lower Permian Unayzah association. This increase in diversity and abundance is explained
Formation of subsurface Saudi Arabia. This unit (subdivided as a result of less stressful conditions resulting from climatic
into four members) is particularly relevant, because it provides amelioration.
a full picture of the paleoclimatic evolution in this region of The last three contributions deal with ichnology. In particu-
Gondwana, from glacial through postglacial to semiarid and arid lar, trace fossils are ideally suited for ecosystem studies because
conditions. The wide spectrum of facies documented includes they provide in situ evidence of organism-substrate interactions.
tillite, reworked diamictite, glaciolacustrine fines and turbi- In Chapter 7, Buatois et al. summarize ichnologic data from
dites, fluvial deposits, paleosoils, and eolianites. These authors eight different Gondwanan basins (Paganzo, San Rafael, Tarija,
differentiate between climatically and tectonically controlled Paraná, Karoo, Falkland, Transantarctic, and Sydney) and note
transgressions and provide correlations across the Arabian Pen- the presence of fresh-water ichnofaunas in direct association
insula. In Chapter 3, Isbell explores the paleoenvironmental with glacially influenced coasts affected by strong discharges of
and paleoclimatic implications of glaciomarine deposits in the meltwater as a recurrent theme. These authors suggest that fresh-
Permian Metschel Tillite of the Transantarctic Mountains. In water conditions prevailed in coastal areas during most of the
contrast to previous interpretations, he proposes that ice entered postglacial times because of a strong discharge of fresh water
the area from at least two different ice centers on opposite sides from melting of the ice masses. They conclude that the classic
of the basin. Abundant evidence of glaciotectonic structures is marine-nonmarine dichotomy used in ichnologic studies may be
documented, including sheared diamictites and thrust sheets. misleading in this type of setting. Desjardins et al. (Chapter 8)
The global significance of this study resides in that multiple gla- document the ichnofauna present in transgressive deposits of the
ciers contain less ice volume than a single massive ice sheet, uppermost Carboniferous and lowermost Permian Tupe Forma-
impacting on global climate and eustatic sea level in a different tion of the Paganzo Basin in western Argentina. These authors
way than would have a single massive ice sheet. Kreuser and discuss ichnologic aspects of the transition from postglacial flu-
Woldu (Chapter 4) provide a detailed characterization of Perm- vial deposits to bay deposits formed during a rise in sea level.
ian euxinic lake deposits preserved in the Idusi Formation of The recognized trace-fossil assemblages reflect the changing
the Ruhuhu Basin in Tanzania. The succession reflects a tran- environmental conditions that result from a base-level rise. This
sition from glacial to postglacial deposits, culminating in the study underscores the interplay of ecology, sea-level rise, and
development of distal alluvial fans during climatic amelioration. paleogeography as controlling factors for trace-fossil distribu-
These extensive anaerobic stratified lakes provided appropriate tion. In Chapter 9, Miller and Isbell focus on the paleoecologic
conditions for deposition of black shales with abundant organic implications of an ichnofauna formed in a large and deep turbid-
matter (up to 11% total organic carbon [TOC]). These authors itic Permian lake, recorded in the Mackellar Formation of the
underscore the importance of these deposits as source rocks and Transantarctic Mountains. The Mackellar ichnofauna is of low
viii López-Gamundí and Buatois

Frakes, L.A., and Crowell, J.C., 1967, Facies and paleogeography of late Paleo-
diversity, and the degree of bioturbation is generally low, sug- zoic Lafonian diamictite, Falkland Islands: Geological Society of Amer-
gesting oxic conditions and restriction of the benthos to areas ica Bulletin, v. 78, p. 37–58.
with low rates of sedimentation. They compare this Permian Frakes, L.A., and Crowell, J.C., 1969, Late Paleozoic glaciation: I, South Amer-
ica: Geological Society of America Bulletin, v. 80, p. 1007–1042, doi: 10
lake with modern Lake Agassiz, and conclude that in spite of its .1130/0016-7606(1969)80[1007:LPGISA]2.0.CO;2.
high paleolatitude (~80°S), the lake was dynamic and biologi- Frakes, L.A., and Crowell, J.C., 1970, Late Paleozoic glaciation: II, Africa
cally active. exclusive of the Karroo basin: Geological Society of America Bulletin,
v. 81, p. 2261–2286, doi: 10.1130/0016-7606(1970)81[2261:LPGIAE
]2.0.CO;2.
ACKNOWLEDGMENTS Frakes, L.A., Amos, A.J., and Crowell, J.C., 1969, Origin and stratigraphy of
Late Paleozoic diamictites in Argentina and Bolivia, in Amos, A.J., ed.,
Gondwana Stratigraphy, IUGS Symposium (Buenos Aires, 1967): Earth
Finally the editors want to thank the following colleagues Sciences, v. 2, p. 821–843.
for dedicating their time to reviewing this volume’s contribu- Frakes, L.A., Matthews, J.L., and Crowell, J.C., 1971, Late Paleozoic gla-
tions: Lucia Angiolini (Università degli Studi di Milano, Italy), ciation: Part III: Antarctica: Geological Society of America Bulletin,
v. 82, p. 1581–1604, doi: 10.1130/0016-7606(1971)82[1581:LPGPIA
Christoph Breitkreuz (Institut für Geologie, Freiberg, Germany), ]2.0.CO;2.
Roberto d’Avila (Petrobras, Brazil), Jim Collinson (USA), Alme- Hambrey, M., and Harland, W., 1981, Earth’s Pre-Pleistocene Glacial Record:
rio Barros França (Petrobras, Brazil), Dirk Knaust (StatoilHydro, Cambridge, UK, Cambridge University Press, 1044 p.
Isbell, J.L., Miller, M.L., Wolfe, K.L., and Lenaker, P.A., 2003, Timing of
Norway), Ricardo Melchor (Universidad de La Pampa, Argen- late Paleozoic glaciation in Gondwana: Was glaciation responsible for
tina), Marcello Guimarães Simões (Sao Paulo State University the development of northern hemisphere cyclothems?, in Chan, M.A.,
at Botucatu, Brazil), Lynn Soreghan (University of Oklahoma, and Archer, A.W., eds., Extreme Depositional Environments: Mega End
Members in Geologic Time: Geological Society of America Special Paper
USA), Luis A. Spalletti (Universidad de La Plata, Argentina), 370, p. 5–24.
Antonio Rocha-Campos (University of Sao Paulo, Brazil), Alfred López-Gamundí, O.R., 1997, Glacial-postglacial transition in the late Paleozoic
Uchman (Jagiellonian University, Poland), John Veevers (Mac- basins of southern South America, in Martini, I.P., ed., Late Glacial and
Postglacial Environmental Changes, Quaternary, Carboniferous–Permian
Quarie University, Australia), and Joonas Virtasalo (University and Proterozoic: Oxford, UK, Oxford University Press, p. 147–168.
of Turku, Finland). Montañez, I., Tabor, N.J., Niemeier, D., DiMichele, W.A., Frank, T.D., Fielding,
C.R., Isbell, J.L., Birgenheier, L.P., and Rygel, M.C., 2007, CO2-forced
climate and vegetation instability during late Paleozoic deglaciation: Sci-
REFERENCES CITED ence, v. 315, p. 87–91, doi: 10.1126/science.1134207.
Powell, C.McA., and Li, Z.X., 1994, Reconstruction of the Panthalassan margin
Crowell, J.C., 1978, Gondwana glaciation, cyclothems, continental positioning of Gondwanaland, in Veevers, J.J., and Powell, C.McA., eds., Permian–
and climate change: American Journal of Science, v. 278, p. 1345–1372. Triassic Pangean Basins and Foldbelts along the Panthalassan Margin of
Crowell, J.C., 1982, Continental glaciation through geologic time, in Climate in Gondwanaland: Geological Society of America Memoir 184, p. 5–9.
Earth History: Studies in Geophysics, Washington, D.C., National Acad- Rygel, M.C., Fielding, C.R., Frank, T., and Birgeinheier, L.R., 2008, The mag-
emy Press, p. 77–82. nitude of late Paleozoic glacioeustatic fluctuations: A synthesis: Journal
Crowell, J.C., 1999, Pre-Mesozoic Ice Ages: Their Bearing on Understanding of Sedimentary Research, v. 78, p. 500–511, doi: 10.2110/jsr.2008.058.
the Climate System: Geological Society of America Memoir 192, 106 p. Veevers, J.J., and Powell, C.M., 1987, Late Paleozoic glacial episodes in Gond-
Crowell, J.C., and Frakes, L.A., 1971a, Late Palaeozoic glaciation of Australia: wanaland reflected in transgressive-regressive depositional sequences in
Journal of the Geological Society of Australia, v. 17, p. 115–155. Euramerica: Geological Society of America Bulletin, v. 98, p. 475–487,
Crowell, J.C., and Frakes, L.A., 1971b, Late Paleozoic glaciation: Part IV, Aus- doi: 10.1130/0016-7606(1987)98<475:LPGEIG>2.0.CO;2.
tralia: Geological Society of America Bulletin, v. 82, p. 2515–2540, doi: Wanless, H.R., and Shepard, F.P., 1936, Sea level and climatic changes related
10.1130/0016-7606(1971)82[2515:LPGPIA]2.0.CO;2. to late Paleozoic cycles: Geological Society of America Bulletin, v. 47,
Crowell, J.C., and Frakes, L.A., 1972, Late Paleozoic glaciation: Part V, p. 1177–1206.
Karoo Basin, South Africa: Geological Society of America Bulletin, Wegener, A., 1915, Die Entsehung der Kontinente und Ozeane: Braunchweig,
v. 83, p. 2887–2912, doi: 10.1130/0016-7606(1972)83[2887:LPGPVK Germany, Vieweg, 367 p.
]2.0.CO;2. Wright, V.P., and Vanstone, S.D., 2001, Onset of Late Paleozoic glacio-eustasy
Du Toit, A.L., 1927, A Geological Comparison of South America with South and the evolving climates of low latitude areas: A synthesis of current
Africa: Washington, D.C., Carnegie Institute, 157 p. understanding: Journal of the Geological Society [London], v. 158,
Eyles, N., 1993, Earth’s glacial records and its tectonic setting: Earth-Science p. 579–582.
Reviews, v. 35, p. 1–248, doi: 10.1016/0012-8252(93)90002-O.
Fielding, C.R., Frank, T.D., and Isbell, J.L., eds., 2008, Resolving the Late
Paleozoic Ice Age in Time and Space: Geological Society of America
Special Paper 441, 354 p. MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

Transgressions related to the demise of the Late Paleozoic Ice Age:


Their sequence stratigraphic context

Oscar R. López-Gamundí
Hess Corporation, 500 Dallas Street, Houston, Texas 77002, USA

ABSTRACT

The Gondwanan Icehouse Period spanned between the mid-Carboniferous and


Early Permian waning by the early Late Permian. Early postglacial sea-level rise
related to the final stage of the Late Paleozoic Ice Age favored creation of accommo-
dation space with preservation potential for productive anoxia events in the newly
inundated shelves and peat-forming conditions favored by rapid water table rise in
updip positions in the basin. The combined effect of fast glacioeustatic sea-level rise
and subsidence along basin margins led to a drastic landward facies shift; the newly
created space was sufficient to accommodate a transgressive systems tract (TST) that,
irrespective of the age of the glacial episode, exhibits common characteristics across
Gondwana. High fresh-water discharges related to the retreat of glaciers resulted
in associated reduction in coastal salinity. Therefore, fjord-like settings as part of
early postglacial inland seas seem to be a valid analogue for many of these TSTs. The
examples of glacial-postglacial transitions analyzed in this contribution are present in
a variety of basin types, namely, those ranging from backarc foreland basins to rifts.
In all of them a clear retrogradational stacking pattern is detectable in the transition
from glacially dominated settings to glacially influenced early postglacial environ-
ments. Examples from South America (Calingasta-Uspallata and Paganzo Basins),
South Africa (Karoo Basin), Peninsular India (several Gondwana basins), and eastern
Australia (Tasmania Basin) help define two basic types of TSTs: (1) complete TSTs,
with a basal part of clast-poor, massive to poorly stratified diamictites, thinly bedded
diamictites, shales with ice-rafted debris (IRD) and IRD-free shales, and an upper
part dominated by open-marine shales representing the maximum flooding of the
shelf; and (2) base-cut TSTs in which the basal transgressive portion is mostly omit-
ted, and the TST is exclusively represented by open-marine shales generally devoid of
IRD. Whereas the complete TSTs are common in cases in which high sediment supply
rates via rain-out, ice rafting, and settling of fines prevail during the early phase of
deglaciation, the base-cut TSTs, on the other end, reflect the dominance of drastic sea-
level rises related to fast glacier retreats.

INTRODUCTION the Earth (Fig. 1). Evidence of this glaciation is widespread in the
Gondwana Supercontinent (Hambrey and Harland, 1981; Crow-
The Late Paleozoic Ice Age (LPIA) is one of the best ell, 1983, 1999; Eyles, 1993; Fielding et al., 2008a). This glacial
recorded episodes of extreme climatic conditions that affected age coincides with a high paleolatitude for Gondwana and the
López-Gamundí, O.R., 2010, Transgressions related to the demise of the Late Paleozoic Ice Age: Their sequence stratigraphic context, in López-Gamundí, O.R.,
and Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana: Geological Society of America Special Paper 468, p. 1–35,
doi: 10.1130/2010.2468(01). For permission to copy, contact [email protected]. ©2010 The Geological Society of America. All rights reserved.

1
2 López-Gamundí

Arabia
20
19

T e
A F R I C A
18

t h
INDIA y
17 s
13 15
14 31
32 33
16
S O U T H 34
12
30
A M E R I C A 7 8 AUSTRALIA
9 29
2 A N T A R C T I C A
6 10 11 27
28
5 24
1
21 26
22
3
P 4 23 25
a n t h 0
a l a s s a 2000 km

Figure 1. Reconstruction of Gondwana supercontinent with simplified outlines of principal basins during the
Late Carboniferous–Early Permian (glacial episodes II and III). Positions of Arabia and Madagascar after
De Wit et al. (1988); South American basins after López-Gamundí et al. (1994); African basins from Veevers
et al. (1994) and Visser (1997a); Falkand-Malvinas Islands in their pre-breakup position east of the coast of
South Africa as originally proposed by Adie (1952); Indian basins after Wopfner and Casshyap (1997); Aus-
tralian basins modified from Struckmeyer and Totterdell (1992) and Lindsay (1997). 1—Calingasta-Uspallata
and Paganzo; 2—Tarija; 3—San Rafael; 4—Tepuel-Genoa; 5—Sauce Grande; 6—Chaco-Paraná; 7—Paraná;
8—Huab; 9—Kalahari; 10—Karoo; 11—Falkland-Malvinas Islands; 12—Zambesi; 13—Congo; 14—Tan-
zania; 15—Malagasy; 16—Peninsular India; 17—Extra-Peninsular India (Himalayan); 18—Salt Range;
19—Yemen–South Arabia; 20—Oman; 21—Pensacola Mountains; 22—Transantarctic Mountains; 23—Tasma-
nia; 24—Murray; 25—Sydney; 26—Bowen; 27—Galilee; 28—Cooper; 29—Pedirka-Arckaringa; 30—Perth;
31—Carnarvon; 32—Canning; 33—Browse; 34—Bonaparte.

growth of high standing topography when Gondwana collided Ghienne et al., 2007). The objective of this contribution is to
with Laurasia to create Pangea (Eyles, 2008). The stratigraphic illustrate the sequence stratigraphic context of the transgressions
record derived from this glaciation is abundant and diverse: dif- in relation to the demise of the LPIA. The emphasis of this study
ferent types of glacioterrestrial and glaciomarine sediments asso- is on the section temporally constrained to the initial deglaciation
ciated with glacially abraded surfaces are present in a wide variety when glacioisostatic rebound lags behind a rapid eustatic rise in
of basin types and tectonic settings. Despite this variability of sea level.
facies, basin types, and tectonic regimes, most of these glaciated
margins were affected during the early deglaciation phase by a CHRONOSTRATIGRAPHIC FRAMEWORK
sea-level rise that imparted a series of conspicuously diagnostic
stacking patterns that can be interpreted in sequence stratigraphic The record of the LPIA seems to cluster around three dis-
terms. Examples of such glacial-postglacial transition have been tinct episodes (Veevers and Powell, 1987; López-Gamundí,
documented not only for the LPIA but also for the Pleistocene 1997; Isbell et al., 2003). The Late Devonian is the oldest of
glaciation. Owing to its relation with large hydrocarbon reser- these episodes (episode I of López-Gamundí, 1997); it was origi-
voirs, evidence of such transition from surface and subsurface nally described by Caputo (1985) and is apparently confined to a
studies has been abundantly documented for the Late Ordovician broad SW-NE belt across the central and northern parts of South
glaciation that affected most of the North African and Arabian America (Caputo el al., 2008; López-Gamundí and Buatois, this
platforms (Beuf et al., 1971; Deynoux et al., 1985; Vaslet, 1990; volume) from Peru (Carlotto et al., 2004; Cerpa et al., 2004) and
Transgressions related to the demise of the Late Paleozoic Ice Age 3

Bolivia (Díaz Martínez and Isaacson, 1994) to Brazil (Solimões sisted after the waning of the Gondwanan Ice Sheet Complex
Basin, Eiras et al., 1994; Amazonas basin, Cunha et al., 1994; (Eyles, 1993) until the end of the Early Permian (Artinskian)
Parnaiba, Goes and Feijo, 1994) and recently extended to the along some margins and upland regions of western Gondwanan
Central African Republic and Niger in Africa (Isaacson et al., basins (northern margins of the Karoo Basin and the Kalahari
2008). A late Fammenian age has been suggested for this glacial Basin, Fig. 1) and particularly in eastern Gondwana, where small
event on the basis of palynological evidence mostly from Bolivia alpine ice caps provided glacial debris during three more minor
(Isaacson et al., 1999). A northward extension of this short-lived, glacial episodes in eastern Australia (Jones and Fielding, 2004).
areally confined glaciation has been proposed for North America The proof of these waning glacial conditions elsewhere in Aus-
(Cecil et al., 2004; Isaacson and Díaz Martínez, 2005; Brezinski tralia is mostly derived from the presence of ice-rafted debris
et al., 2008), but its evidence is tenuous so far and requires fur- (IRD) in the form of isolated dropstones in the Tasmania Basin
ther work. (Fig. 1, Clarke and Banks, 1975) and the Sydney Basin (Fig. 1,
The postglacial transgressive sections analyzed herein corre- Eyles et al., 1998). Indirect evidence for the presence of these
spond to glacial episodes II and III as defined by López-Gamundí waning, post-Sakmarian cold periods in eastern Australia is pro-
(1997). The former is associated with the mid-Carboniferous vided by studies that show lowered atmospheric pCO2 before the
Levispustula fauna, and the latter is characterized by the Early permanent transition to an ice-free Earth ca. 260 Ma (Montañez
Permian (Sakmarian) Eurydesma fauna. These two cold-water et al., 2007).
faunas have been identified in the marine fine-grained sediments The glacial-postglacial transition exhibits a similar retrogra-
that rest directly on, or in similar facies interbedded with, gla- dational stacking of facies irrespective of their ages; therefore
ciogenic deposits across the Gondwana Supercontinent. The the analysis attempted in this contribution will be focused on the
age of the Levipustula fauna has been traditionally considered sedimentological characteristics of these transgressive deposits
Namurian–Westphalian (Roberts et al., 1976; see Cisterna and and their sequence stratigraphic context rather than on their chro-
Sterren, this volume, for a recent review on the age of this fauna) nostratigraphic significance. Analogies for vertical stacking pat-
but more recently has been confined to the early Namurian on the terns, for example, will be highlighted irrespective of the ages
basis of absolute ages from SHRIMP (sensitive high-resolution of the successions. Furthermore, probably the best documented
ion microprobe) zircon-based dating in interbedded tuffs in Aus- pre-Pleistocene glacial-postglacial transition is the one that cor-
tralia (Roberts et al., 1995; Fielding et al., 2008b). Glacial episode responds to the Late Ordovician (Hirnantian, late Ashgill) glacia-
II is restricted to the westernmost part of the Gondwanan Super- tion in North Africa and the Arabian Platform (Beuf et al., 1971;
continent in southern South America (Amos and López-Gamundí, McClure, 1978; Hambrey, 1985; Vaslet, 1990; Le Heron et al.,
1981a; González, 1990; López-Gamundí, 1984, 1989, 1997). 2007; Ghienne et al., 2007), which shares sedimentological and
Glacial episode III is widespread across the Gondwanan basins sequence stratigraphic similarities with the late Paleozoic exam-
from South America (Frakes and Crowell, 1969; Rocha-Campos ples analyzed herein. Both glacial ages share a similar pattern
and dos Santos, 1981; López-Gamundí, 1997; Rocha-Campos of deglaciation characterized by a drastic landward facies shift
et al., 2008) across Africa (Crowell and Frakes, 1972; Theron during a fast transgression. However, unlike the short-lived Late
and Blignault, 1975; Visser, 1983, 1987a, 1987b, 1989, 1997a, Ordovician glaciation restricted between ~1 m.y. (Brenchley
1997b; Wopfner and Kreuser, 1986; Von Brunn, 1994, 1996), et al., 1994) and 10 m.y. (Ghienne, 2003), the LPIA is the longest
Arabia (Helal, 1964; Kruck and Thiele, 1983; McClure, 1980; lived glaciation in Earth history, perhaps one to two orders of
Braakman et al., 1982; Levell et al., 1988; Melvin and Sprague, magnitude longer than the Late Ordovician event (Eyles, 1993).
2006; Melvin et al., this volume) to India (Niyogi, 1961; Smith, Thus, eustatic fluctuations caused by numerous ice advances and
1963; Casshyap and Qidwai, 1974; Ghosh and Mitra, 1975; retreats are commonly recorded across Gondwana for the late
Casshyap and Srivastava, 1987), and Antarctica (Lindsay, 1970; Paleozoic. The emphasis in this contribution is on the sedimen-
Miller, 1989; Collinson et al., 1994; Isbell et al., 2008) to Austra- tological and sequence stratigraphic characteristics of the well-
lia (Crowell and Frakes, 1971; Struckmeyer and Totterdell, 1992; preserved glacial-postglacial transition rather than on the shorter
Lindsay, 1997). SHRIMP ages from the lower Dwyka Formation lived, often cannibalized, glacial-interglacial fluctuations. Four
(Bangert et al., 1999) in the Karoo Basin (Fig. 1) and the Itararé examples have been selected; they cover a wide spatial and tem-
Group (Rocha-Campos, 2006, in Rocha-Campos et al., 2008) in poral range from west to east: Calingasta-Uspallata and Paganzo
the Paraná Basin (Fig. 1) indicate a Late Carboniferous (Stepha- Basins (glacial episode II), Karoo and associated basins (glacial
nian) age for the onset of glacial episode III in South Africa and epsiode III), basins in Peninsular India (glacial episode III), and
South America. There seems to be consensus on a Sakmarian age the Tasmania Basin in Australia (glacial episode III) (Fig. 1).
for the widespread Early Permian collapse of ice sheets (Dick-
ins, 1996; Isbell et al., 2003). Tuff zones at the base of the post- SEQUENCE STRATIGRAPHIC CONTEXT FOR
glacial Prince Albert Formation in the southwest Karoo Basin POSTGLACIAL TRANSGRESSIONS
provided SHRIMP ages of ca. 290 Ma (Bangert et al., 1999), a
mid-Sakmarian age based on the numerical time scale of Grad- The term glacioeustasy is referred in this contribution to the
stein et al. (2004). Exceptionally, local glacial conditions per- process that generates changes in sea level that can be related to
4 López-Gamundí

changes in ice volume. A first approximation to the magnitude of offset by eustatic effects resulting in minimal sea-level fluctua-
eustatic sea-level rises related to the rather abrupt demise of any tions (Barrie and Conway, 2002). Nevertheless, the Pleistocene
glaciation is provided by estimates of volumes of ice released. and pre-Pleistocene glacial-postglacial stratigraphic record pro-
Table 1 is a compilation of the estimated present-day area and vides abundant examples of deepening-upward sections that indi-
volume of glaciers and the maximum sea-level rise potential. cate a drastic increase in accommodation space during the early
It is worth noting that minor melting episodes related to small, stages of deglaciation. This record is confined to a portion of the
local ice caps have negligible effects on sea-level fluctuations, as basin where the postglacial sea-level rise and the isostatic subsid-
opposed to major waxing of large areas (i.e., East Antarctica). In ence caused by water loading might have exceeded the combined
the latter scenario, significant (~65 m) maximum sea-level rises effect of glacial erosion and glacio-isostatic rebound (Bjorlykke,
are expected. This effect could have been attenuated if several 1985; Nystuen, 1985), resulting in a subsequent glacioeustatic
small ice sheets, rather than a large single one, were present as transgression (Crowell, 1978). Areas close to the ice margin
suggested for at least part of the LPIA by Isbell et al. (2003). A are commonly dominated by isostatic depression and rebound,
similar approach has been used by Le Heron and Dowdeswell whereas more distal areas are dominated by eustatic changes like
(2009) for their calculation of the postglacial sea-level rise related submergence during deglaciation (Miller, 1996).
to the Late Ordovician glaciation. These authors estimated a The main controls on stratal architecture in any given
postglacial eustatic sea-level rise of ~75 m on the basis of a small sequence can be narrowed down to (1) accommodation space
ice-sheet hypothesis and concluded that this approach adequately (the summation of subsidence rates and sea-level fluctuations),
accounts for the estimated magnitude of postglacial transgression and (2) sediment supply. Figure 2 illustrates some of the possible
(~45–80 m) associated with ice-mass decay. Their estimate of conditions under which accommodation space is created during
50 m of postglacial sea-level rise only from the contribution of an a stage of sea-level rise in a subsiding basin (Jervey, 1988). Case
ice sheet in the North African Platform (the largest of the three A illustrates a classic example in sequence stratigraphy where the
ice sheets modeled) is in agreement with independent estimates summation of a constant (in this case slow to moderate) subsid-
derived from studies on depth-related benthic fossil communities ence rate and a symmetrical sinusoidal eustatic sea-level curve
(Ross and Ross, 1996). creates a relative sea-level curve (accommodation space). Cases
The effects of isostasy and its lateral variations along a con- B, C, and D illustrate three alternative outcomes of the most
tinental margin (i.e., differential cross-shelf isostatic response of likely scenario for postglacial transgressions when the eustatic
a thin crust) may potentially mask the otherwise dominant effect sea-level cycle shows a clear asymmetry from a faster (glacioeu-
of glacioeustatically induced sea-level rises on the stratal archi- static) sea-level rise (expressed by the steep part of the eustatic
tecture and sequence development of the early postglacial basin sea-level curve). The subsidence rates vary and the eustatic sea-
fill. This was frequent in tectonically active regions affected by level cycle remains constant in these three last cases (Figs. 2B,
glaciation, in which the individual factors related to the glaciation 2C, and 2D). Case C differs from case B only in that case C
(eustasy and isostasy) and those of local origin (i.e., tectonically includes a faster subsidence rate; in this situation the postglacial
induced subsidence) are difficult to discern. Studies of Pleisto- sea-level rise could be partially masked because very fast subsid-
cene glaciation along the western Canadian continental shelf ence rates exceed the sea-level rate with accommodation space
show this extreme variability (Clague et al., 1982). Whereas a not significantly destroyed, and consequently, even when eus-
marine transgression owing primarily to eustatic rise occurred tatic fall is taking place, deep-water conditions similar to those
in the southern regions of the Strait of Georgia adjacent to the dominant during the transgressive stage may persist in the high-
central and northern Strait of Georgia, it appears that 100 m of sea-level stage. Finally, case D illustrates an extreme case with
isostatic adjustment associated with local tectonic changes were variable subsidence rates (from initially moderate to fast) and

TABLE 1. ESTIMATED PRESENT-DAY AREA AND VOLUME OF GLACIERS AND MAXIMUM SEA-LEVEL-RISE POTENTIAL
Geographic region Area Percent Volume Percent Maximum sea-level-rise
2 3
(km ) (%) (km ) (%) potential
(m)
Ice caps, ice fields, valley glaciers, etc. 680,000 4.24 180,000 0.55 0.45
Greenland (inland ice) 1,736,095 10 .82 2,600,000 7 . 90 6. 5 0
Local ice caps and other glaciers 48,599 0.30 20,000 0.06 0.05
Antarctica 13,586,400 84.64 30,109,800 91.49 73.44
East Antarctica 10,153,170 26,039,200 64.80
West Antarctica 1 ,9 1 8 , 1 7 0 3,262,000 8.06
Antarctic Peninsula 4 46,690 227, 100 0. 46
Ross Ice Shelf 53 6 , 0 70 2 2 9, 6 0 0 0 .01
Ronne-Filchner ice shelves 532 ,20 0 351 ,90 0 0.11
Totals 16 , 0 5 1 , 0 9 4 100 .0 0 32,909,800 1 00 .0 0 8 0. 4 4
Notes: From Williams and Ferrigno (2008). Values in italicized rows are a subset of the “Antarctica” row values.
Transgressions related to the demise of the Late Paleozoic Ice Age 5

A B
Depth/Thickness

Depth/Thickness
RS Figure 2. Accommodation space as a
L function of subsidence rates and sea-
RSL
level changes. Relative sea level (RSL)
is equivalent in this case to accommo-
dation because both curves begin at
zero-depth water. Modified from Jervey
nce ce (1988) and Emery and Myers (1996).

ES
side iden
ES
Sub L

L
s (A) Accommodation space curve as the
Sub
summation of a sinusoidal eustatic sea-
Time level curve (ESL) and a slow to moder-
Time ate subsidence rate. B, C, D: Resulting
accommodation space from an asym-
metric sinusoidal sea-level curve (ESL)
C D RSL
and variable subsidence rates. (B) Slow
to moderate subsidence (same as in A).
RSL
Depth/Thickness

Depth/Thickness
(C) Moderate to high subsidence rate;
e note higher resulting accommodation
nc
side space than in B; note that with faster sub-
Sub sidence, maximum accommodation is
progressively later. (D) Increasing sub-
ce ES sidence rate, which creates the most ac-
iden L ES commodation space. See discussion in
Subs L
text.

Time Time

creation of deep-water conditions during transgression and even tectonics, and/or delta-lobe switching; when recognizable
deeper water conditions during the high-sea-level stage. In cases globally, a flooding surface can be assigned to a eustatic
C, and particularly D, subsidence rates are so high that the basin sea-level rise.
undergoes no decrease in accommodation even though eustatic 3. Maximum flooding surface (MFS): the surface that cor-
fall may be occurring (Emery and Myers, 1996). Cases C and D responds to the most transgressive stratigraphic architec-
correspond to the concept of forced transgression of Chough and ture. Maximum flooding surfaces can be identified by the
Hwang (1997). presence of condensed sections that reflect distant sedi-
The classic definition of sequence is adopted in this contri- ment sources at the peak of transgression (Loutit et al.,
bution. Sequences are composed of depositional packages or sys- 1988; Posamentier and Allen, 1999). In shelfal and basi-
tems tracts (Brown and Fisher, 1977) deposited during specific nal settings, maximum flooding surfaces can be identi-
phases of the relative sea-level cycle (Posamentier and Allen, fied by evidence of condensation (i.e., firmgrounds),
1999). A systems tract is a linkage of contemporaneous depo- fine-grained (silt-clay) deposits, the presence of high
sitional systems and is defined by the nature of its boundaries organic content (expressed as high gamma-ray values in
(see below). Three principal tracts are used in this contribution: well logs), phosphate levels, fossiliferous beds, and outer
lowstand, transgressive, and highstand systems tracts (LST, TST, shelf carbonates. In paralic successions the MFS is coeval
and HST, respectively). Boundaries between system tracts are with the most landward position of the shoreline (Emery
defined by key sequence-stratigraphic surfaces: and Myers, 1996). In areas where it is difficult to identify
1. Sequence boundary (SB): an unconformity and its cor- and trace laterally a maximum flooding interval (MFI) or
relative conformity corresponding to the most regressive zone is recognized for such a surface. The MFS should lie
configuration of stratigraphic architecture. A sequence within the MFI, but no sedimentological evidence could
boundary is expressed as a facies dislocation, a surface be found to define such a surface.
where rocks of a shallower facies rest directly on rocks of 4. Transgressive ravinement surface (TRS): the erosional
a significantly deeper facies (Emery and Myers, 1996). surface cut by wave action during transgression. This
This facies dislocation (basinward facies shift) implies a transgressive surface is the first significant marine flood-
fall in relative sea level. ing surface across the shelf within a sequence (Van Wag-
2. Flooding surface (FS): a surface that separates abruptly oner et al., 1988). It defines the base of the TST. As pointed
deeper deposits from underlying shallower water sedi- out by Posamentier and Allen (1999), some potential con-
ments (Van Wagoner et al., 1990). This surface implies a fusion arises when terms like transgressive surface and
rise in relative sea level and can be generated by eustasy, flooding surface are used because they are commonly
6 López-Gamundí

used interchangeably. The former is reserved in this con- to crudely stratified diamictite (interpreted as subglacial till)
tribution for a surface “marking the onset of significant beds partly associated with lacustrine shales with IRD. The
and extended period of transgression within a succession” glacial beds grade upward to fluvial deposits, coals, and carbo-
(Posamentier and Allen, 1999, p. 5), restricting flooding naceous shales, which represent the postglacial stage. The typi-
only to the process of subsequent inundation. cal fluvial facies are predominantly coarse-grained sediments
made up of pebble to cobble conglomerates, pebbly sandstones
FACIES ASSOCIATIONS RELATED TO with trough cross-bedding, and medium- to fine-grained rippled
POSTGLACIAL TRANSGRESSIONS sandstones deposited in a gravelly braided channel complex or
as part of a subaqueous proximal outwash fan during a phase of
With few exceptions in which terrestrial subglacial tills and glacier retreat.
associated facies are the dominant deposits, the stratigraphic In coastal environments, marine influence can be expressed
record of the LPIA and its subsequent final deglaciation record as marine fossiliferous shales, considered as maximum flooding
across Gondwana have been preserved mostly in subaqueous surfaces, whereas the underlying coals and carbonaceous rocks
settings. The main reason for the skewed record is possibly the are interpreted as initial flooding surfaces within a TST. In exclu-
significantly different preservation potential of terrestrial and sively nonmarine settings the coals may represent the correlatives
subaqueous (marine or lacustrine) glacial environments. Preser- of maximum flooding surfaces developed basinward.
vation potential depends on a delicate balance of long- and short- Coals and carbonaceous shales present in this facies asso-
term components. The long-term component is largely controlled ciation have several characteristics that can be related to the
by the geotectonic setting (i.e., forearc basin, cratonic basin, etc.) accumulation of peat on initial rapid transgressions and increas-
and the related tectonic activity of the basin after deposition (i.e., ing accommodation space. They are relatively thin and areally
postdepositional erosion); the short-term component is con- restricted as a result of increasing accommodation (cf. Bohacs
nected to the erosion coeval with sedimentation (Nystuen, 1985). and Sutter, 1997). Also, they are characterized by high sulfur (S)
This short-term component of the preservation potential is sig- content since the original mires near the coast may have been sub-
nificantly low for terrestrial settings. merged in brackish or salt water. In peat forming environments
Two main facies associations were connected with glacier with brackish and marine influence (i.e., rheothropic mires),
retreat by López-Gamundí (1997) based on either nonmarine or groundwater and seawater are important sources of primary S. In
marine conditions that prevailed in any specific part of a basin. these brackish- to marine-influenced peats, the S content can be
Those are, respectively, a valley-glacier-retreat facies association high (>10% in extreme cases), whereas in fresh-water peats tend
and a submarine-retreat facies association. The Eyles et al. (1983) to be low (<1%) (Casagrande et al., 1977; Cohen et al., 1989;
facies code for diamictite sequences is used in this contribution. Phillips and Bustin, 1996). Therefore, high S contents in coals
have been traditionally used to infer marine influence (Williams
Valley-Glacier-Retreat Facies Association and Keith, 1963; Raymond and Davies, 1979).

This facies association developed along basin margins under Submarine-Retreat Facies Association
predominantly subaerial or shallow subaqueous conditions dur-
ing the retreat of a glacier front. Owing to the relatively low ini- This facies association is present in basin margins under
tial subsidence along the basin margins, the fill of these valleys is waning glacial marine influence; it is characterized by the domi-
modest in thickness when compared with the submarine-retreat nance of fine-grained (clay-silt) sedimentation in open-marine
facies association developed basinward. Any sea-level rise has areas predominantly below wave base. Underlying glacial facies
significant effects inland by raising the water table. This influ- are characterized by laterally extensive beds of massive to poorly
ence of a sea-level rise on mires was summarized by Bohacs and stratified diamictites (Dmm and Dms, Eyles et al.,1985) inter-
Sutter (1997), who estimated that a 5 m rise in sea level would preted as rain-out tills, thick to thin-bedded diamictites inter-
propagate inland 50 km from the shore across a low-topography preted as the product of remobilization of glacial debris via
coastal area, producing a 3.5 m rise in the groundwater table. gravity flows (mostly discrete debris flows), and IRD-bearing
Terrigenous organic matter can be preserved to form coals only mudstones and thin-bedded diamictites with outsize clasts (drop-
when and where the overall increase in accommodation approx- stones), interpreted as the combined product of cohesive debris
imately equals the production rate of peat (Bohacs and Sutter, flows and ice rafting (López-Gamundí, 1991). The presence of
1997). For mires, base level is the groundwater table. Subsidence intra-till striated boulder pavements within mostly massive dia-
varies much more slowly than does the groundwater table and mictites (interpreted as subglacial tills) is common and reflects
thus is commonly the major long-term control of accommodation high-frequency advance-retreat fluctuations of the glacier (cf.
in nonmarine settings. Eyles, 1988). These underlying glacial facies are similar to those
The postglacial facies in the valley-glacier-retreat setting described by Evans and Pudsey (2002) as proximal glaciomarine
rests on basement or, more commonly, on glaciogenic facies sediments under the Antarctic Peninsula ice shelf. They consist of
represented by thin (5–10 m), laterally discontinuous, massive coarse-grained facies with stratification and laminae, gradational
Transgressions related to the demise of the Late Paleozoic Ice Age 7

contacts, dropstone structures (cf. Ovenshine, 1970; Thomas and postglacial shales in the Calingasta-Uspallata Basin (Amos and
Connell, 1985) coarse-tail grading, and/or sorted sand and silt. Rolleri, 1965; González, 1985). An increasing body of evidence
Facies associations indicate the interaction of three sedimentary suggests that sedimentation in the western and parts of the east-
processes following decoupling of grounded ice during deglacia- ern Paganzo Basin took place in an inland sea at least during the
tion: meltout, gravity flow, and current transport. This proximal transgression that followed the glacier retreat, creating conditions
zone is an area of high sedimentation owing to ice melting (Drewry for fjord-like sedimentation with variable salinity (Kneller et al.,
and Cooper, 1981; Powell, 1984); therefore, rain-out tills prevail 2004; Buatois et al., 2006; Desjardins et al., this volume; Lima-
(Powell et al., 1986). Deposits from gravity flows include bedded rino et al., 2010).
diamicton, internally massive with normal coarse-tail grading Glacial sedimentation in the Calingasta-Uspallata Basin and
and with sharp contacts. These characteristics indicate deposition the western Paganzo Basin was dominated by massive (Dmm)
from cohesive debris flows. Similar thin-bedded (<50 cm) debris to poorly stratified (Dms) sandy or muddy diamictites, inter-
flow deposits have been described in Pleistocene glacially influ- preted as subglacial and rain-out tills, resting on metasedimen-
enced environments such as those around the continental margin tary basement, associated with well-stratified muddy diamictites
of Antarctica (Wright et al., 1983), the Nova Scotian Slope (Hill (Dms[r]), pebbly (dropstone) shales (Fld), and high- and low-
et al., 1982), and Baffin Bay (Aksu, 1984), and they constitute a density turbidites (López-Gamundí, 1987). Evidence of glacial
conspicuous facies in the early postglacial TSTs. abrasion in both basins is widespread (López-Gamundí and
The fine clastics of this facies association can rest directly Martínez, 2000). The basal fill of the Calingasta-Uspallata Basin
over the glaciomarine deposits or grade upward from an inter- and the western Paganzo Basin exhibits significant lateral facies
vening facies characterized by mudstones with IRD and thin- changes caused by progressive sedimentary fill across an irregu-
bedded diamictites. This transitional facies is interpreted as the lar pre-Carboniferous topography.
final stage of sedimentation under glacial influence as evidenced In the Calingasta-Uspallata Basin the basal deposits range
by glacially derived gravity (debris) flow deposits and ice raft- from proximal glaciomarine deposits with intra- and inter-till
ing. The glacial-postglacial section exhibits a clear fining-upward striated boulder pavements to glacially influenced sediments
tendency in concert with a drastic facies shift typical of condi- deposited by sediment gravity flow, mostly cohesive debris flows,
tions when the rate of accommodation (in this case dominated with high- and low-density turbidity currents. These facies are
by the rate of sea-level rise) exceeds the sediment supply rate widespread along the eastern margin of the Calingasta-Uspallata
(submarine retreat model, Edwards, 1986). A modern analogue Basin (Fig. 3). The Hoyada Verde section on the eastern basin
of this depositional setting is present in the ice shelves of the margin (Fig. 4) is an example of the glacial-postglacial transi-
Antarctic Peninsula (cf. Evans and Pudsey, 2002), where mas- tion. The lower glacial section is composed of clast-rich, massive
sive, bioturbated terrigenous silty clay dominates sedimentation to crudely stratified diamictites (Dmm) interpreted as subglacial
as the grounding line recedes toward the coast with progressive tills (Fig. 5A), stratified, clast-poor diamictites (Dms) considered
deglaciation. Decreasing quantities of IRD may also indicate sig- rain-out tills, and subordinate thinly bedded diamictites with IRD,
nificant distance from the retreating glacier front. interpreted as the combined product of debris flows and ice raft-
ing. This glacial section is capped by a single-layer boulder pave-
EXAMPLES OF POSTGLACIAL TSTS IN GONDWANA ment (Fig. 5B) made up of clasts up to boulder size with parallel
to subparallel striations (González, 1981; López-Gamundí, 1983).
Calingasta-Uspallata and Paganzo Basins The boulder pavement caps a thick, mostly massive diamictite
section with isolated paleochannels containing reworked fossils
The Calingasta-Uspallata and Paganzo Basins lie along the and glendonite nodules, as reported by González (1981), who
western margin of Gondwana (Fig. 1). They are part of a series also identified glendonite concretions in underlying shales within
of backarc foreland basins developed along the active Paleopa- the glaciogenic section. Paleo–ice flow directions are from NNE
cific margin (López-Gamundí et al., 1994). A series of basement to SSW, coincident with paleocurrents obtained from ripples in
highs fragmented sedimentation in this basin with a depocenter the postglacial fine-grained sandstones (Fig. 4, López-Gamundí,
close to the arc on the west (Calingasta-Uspallata) and a seg- 1983). The pavement is overlain by a laterally discontinuous, thin
mented basin cratonward (Paganzo) (Fig. 3). Sedimentation (>20 cm) bed of massive muddy diamictite (López-Gamundí and
began in about the Early Carboniferous (Sessarego and Cesari, Martínez, 2000) that passes upward into thin-bedded diamictites
1989) and expanded during the mid-Carboniferous (Namurian) (pebbly mudstones) with outsize clasts (Fig. 5C) and shales with
with predominantly glacially influenced marine sedimentation in dropstones that are considered part of the early postglacial TST.
the Calingasta-Uspallata Basin and the western Paganzo Basin This facies association grades upward by gradual disappearance
(Frakes and Crowell, 1969) as well as transitional environments of the thin-bedded diamictites with IRD into an interval domi-
in the eastern part of the Paganzo Basin. Glacial deposits are nated by IRD-free shales (Fig. 5D) with marine fossils of the
widespread in both basins; they correspond to glacial episode II Levipustula fauna (Sterren and Cisterna, 2006; Cisterna and Ster-
(López-Gamundí, 1997), which is biostratigraphically confined ren, this volume).These fossiliferous, IRD-free shales are consid-
by the presence of the Levipustula fauna mostly in overlying ered the maximum flooding interval (MFI) in the late postglacial
G SG
Si S Si SEDIMENTARY DOMAINS

Open marine

Transitional
(inland sea)
Continental

SG
Si

H
C Pa

M Pa

C
Je
HV
SG
Si

SG
Si
AJ
100 km

Coarse-grained
sandstones
Fine-grained
sandstones
with ripples (Sr)
Fine-grained
sandstones

TST
Mudstones
Late
Shales (Fl)

Dropstone Early
shales (Fld)

Thinly bedded 20 m
diamictites (Dms)
Massive and
HV crudely stratified
AJ diamictites (Dmm) 0

Figure 3. Calingasta-Uspallata and Paganzo Basins during glacial episode II (López-Gamundí, 1997)
with sections illustrating the glacial-postglacial transition in the three main sedimentary domains. Agua
de Jagüel (AJ) section modified from López-Gamundí (1984); Hoyada Verde (HV) section modified from
López-Gamundí (1983); Malanzán (M) section from Andreis et al. (1986); Cortaderas (C), Huaco (H),
and Paganzo (Pa) sections after Limarino et al. (2002). Other location cited in the text: Jejenes (Je). Only
upper-most glacial section illustrated in logs. Grain-size scale in logs: Si—silt; S—sand; G—gravel. TST—
transgressive systems tract.
TS

HST
Figure 5D

Maximum Flooding
Interval (MFI)

Early TST
Figure 5C TST
H o y a d a Ver d e F o r m a t i o n

Figure 5B

Figure 5A

Shales (Fl) Trough cross-


stratification

Mudstones Starved ripples

Dropstone shales (Fld) Burrows

Marine fauna
Fine-grained sandstones (Levipustula)
Deformed
Pebbly sandstone sandstone bodies

Dropstones (IRD)
30 m Conglomerates

Stratified diamictites (Dms), Paleo ice flow


thinly bedded directions

Stratified diamictites (Dms) Striated boulder


pavement (SBP)

0 Massive diamictites (Dmm) Paleocurrents

Si S
G

Figure 4. Hoyada Verde Formation section in Calingasta-Uspallata Basin. Adapted from López-Gamundí (1983,
1984). Grain-size scale in log: Si—silt; S—sand; G—gravel. HST—highland systems tract; TS—overlying
Tres Saltos Formation; TST—transgressive systems tract.
10 López-Gamundí

A C

B D
Sr+Fl

Fl

Figure 5. Hoyada Verde Formation outcrops; see log in Figure 4 for stratigraphic locations of photos. (A) Clast-rich massive diamictite (Dmm);
circled hammer as scale in center. (B) Inter-till striated boulder pavement. (C) Thin-bedded diamictites (Dms) with dropstone, early TST.
(D) Postglacial fossiliferous shales (Fl) of the late TST and overlying fine-grained sandstones and mudstones (Sr+Fl), part of the HST; cliff-
forming sandstone beds on the left (west) correspond to the overlying Tres Saltos Formation.

TST. The upward transition to the HST (Fig. 4) is manifested Kneller et al., 2004; Limarino et al., 2004, 2010) after a marine
by increasing participation of interbedded mudstones and fine- microfauna was reported from various localities (i.e., Las Lajas
grained sandstones with slightly asymmetric to symmetric (wave) Formation, Je in Fig. 3, Cesari and Bercowski, 1997; Guandacol
ripples (Figs. 4 and 5D; “flagstones” of Mésigos, 1953) with a Formation, Ottone, 1991). The microfauna found in the post-
clear coarsening- and thickening-upward stacking pattern. Traces glacial deposits of the Guandacol Formation is associated with
common in marginal-marine settings were recognized by Mán- low-salinity linguliformean brachiopods (Martínez, 1993) and a
gano et al. (2003). This interval has been interpreted as deposited brackish water ichnofauna (Buatois et al., 2006; Desjardins et al.,
in the transition from offshore to lower shoreface (Cisterna and this volume). Two main ichnofacies separated by a nonbiotur-
Sterren, this volume); it is considered in this contribution as part bated black shale were identified. A lower depauperate Cruziana
of the HST (Fig. 4). assemblage is characterized by brackish-water, low-diversity,
The predominant glacial association in the contiguous simple forms in fine-grained sandstones with current and
Paganzo Basin consists of massive and crudely bedded dia- combined-flow ripples and IRD, and a granule conglomerate and
mictites (Dmm), thin-bedded diamictites (Dms), and shales with very coarse grained sandstones (Mángano and Buatois, 2004).
ice-rafted material (dropstone facies, Fld) deposited in a brackish The upper assemblage (fresh-water Mermia ichnofacies, Buatois
seaway (Guandacol embayment) partially open to the Calingasta- and Mangano, 2003) is present above the nonbioturbated black
Uspallata Basin (Fig. 3). The western part of the Paganzo Basin shales (MFI) and is comparatively more diverse. It is dominated
was more influenced by marine conditions, and several examples by grazing trails in prodelta, fine-grained sandstones with current
of previously interpreted lacustrine mudstones resting on glacial ripple cross-lamination (Fig. 3).
deposits are now considered part of the postglacial transgres- Sections traditionally considered of lacustrine origin, par-
sion in coastal fjord-like environments (Limarino et al., 2002; ticularly in the eastern Paganzo Basin such as the Malanzán
Transgressions related to the demise of the Late Paleozoic Ice Age 11

Formation (Fig. 1, M in Fig. 3), have been recently reinterpreted as Karoo and Associated Basins (Paraná Basin, Falkland-
having been connected to an inland sea. The mudstone-dominated Malvinas Islands, and Sauce Grande–Ventana Fold Belt)
section (Member 2 of the Malanzán Formation, Andreis et al.,
1986) overlying the basal glacial diamictites and associated Karoo Basin
breccias and conglomerates (Member 1, Andreis et al., 1986) are The Karoo Basin (Figs. 1 and 6) developed as part of a
now considered part of the postglacial transgression of glacial series of linked foreland basins behind the Panthalassan mar-
episode II owing to the discovery of brackish-water acritarchs, gin (Veevers et al., 1994). Sedimentation was initially confined
green algae that could tolerate a wide range of salinities like to the southern Karoo Basin adjacent to the Cape fold belt, and
Botryococcus, and marine acritarchs (Navifusa and Greinervil- then spreading northward. The basin can be subdivided into three
lites) (Gutiérrez and Limarino, 2001). This section represents distinct segments: (1) a foredeep, close to the rising Cape fold
the easternmost extent of the postglacial transgression in the belt, transgressed by an interior seaway; (2) a forebulge uplifted
Paganzo Basin (Limarino et al., 2002, 2006); the association of above base level; and (3) a shallow backbulge depozone to the
the acritarchs and green algae with abundant spores, log frag- north (Catuneanu, 2004). This subdivision broadly corresponds
ments, and coalified particles suggests proximity to the coast. to the Dwyka glaciogenic facies distribution proposed by Vis-
Also the coexistence of specimens of marine salinity with other ser (1986, 1989), with a platform facies association in the south
elements of brackish- to fresh-water habitats indicates drastic and a valley facies association in the north. Only in the forebulge
salinity fluctuations probably related to extreme fresh-water area restricted to the northernmost part of the basin, nonmarine
release from ice melting into an inland sea (Buatois et al., 2006, deposits prevail with colluvial and glacial-outwash alluvial and
this volume) connected with an open sea to the west (Calingasta- fluvial facies (Faure et al., 1996; Bordy and Catuneanu, 2002).
Uspallata Basin). Lacustrine dropstone facies have been reported to the north,

+ 16° E + + + +
20° E + 22° E 24° E
+ 26° E
+
28° E

+ + + CZ
+ + + +
+ + + +
+ + + + + DP1
28° S + 28° S

+ +
+ + + + + A
+ +
+ + +
+ + + + +
+
30° S 30° S
+ +
+
At

+ +
la

an
+
nt

ce
ic

~ ~ ~

32° S
~ ~

32° S
~
O
West
Oc

Falkland n
ia
Ind
CZ B
ean

~ ~ ~ ~ ~ ~ ~~ ~ ~~ ~
~ ~ ~ ~ ~
34° S ~ ~ ~ ~ East
Falkland
18° E
~ CZ 22° E 24° E
CZ
26° E 28° E
0 100 200 300 km

Drakensberg Group Beaufort Group Dwyka Fm. + Cape Supergroup


CZ Cenozoic (Jurassic) (latest Permian-Triassic) Ecca Gr. ~ (Ordovician-Devonian)

Lebombo Group Stormberg Group Precambrian


(Jurassic) (Triassic - Early Jurassic)
Ecca Group Dwyka Formation ++ Basement

Location of cross section Distribution of Location of borehole DP1


A B (modified from Visser, 1987a) Vryheid Formation DP1 (from Faure and Cole, 1999)

Figure 6. Generalized map of the Karoo Basin with location of A–B cross section (Fig. 7), DP1 borehole, and areal extent of Vryheid (Lower
Ecca Formation). Pre-breakup Falkland Islands position based on Marshall (1994).
12 López-Gamundí

suggesting deposition in glacial or periglacial lakes. Marine fos- 1989). Seawater is an abundant source of S in sulfate; in contrast
sils have been reported from the base of the Prince Albert Forma- the concentration of sulfate is much lower in fresh water (Berner
tion (McLachlan and Anderson, 1973) in the central and western and Raiswell, 1984). Four samples from the postglacial shales
parts of the Karoo Basin. (Prince Albert Formation) indicate a C/S ratio of <2, which cor-
The Dwyka sediments exhibit a wide gamut of facies (mas- responds with marine mud deposition. Furthermore, the positive
sive to thin-bedded diamictites and dropstone (IRD) shales, “plat- S intercept in the C versus S plots suggests euxinic bottom con-
form facies” of Visser, 1986, 1989) arranged in four deglaciation ditions with H2S-laden water being present (Leventhal, 1987).
sequences (Fig. 7, Visser, 1997b); facies in the overlying Prince Mudstones within the glacial Dwyka section have a C/S ratio
Albert Formation are dominated by siliceous shales with phos- of 5 to 30, which suggests possible brackish conditions during
phatic chert and carbonate lenses and concretions. Glendonites, deposition. Geochemical analyses based on rare-earth-element
carbonate pseudomorphs after ikaite (CaCO3·6H2O), form at near (REE) patterns, Sr concentrations, and 87Sr/86Sr ratios from cal-
freezing temperatures and have been reported from just above the cite concretions found in the uppermost Dwyka sediments and
top of the Dwyka deposits (McLachlan et al., 2001). Pioneering overlying Prince Albert shales along the Cape fold belt suggest
work by Visser (1989) on geochemical studies on carbon (C) and that fresh-water conditions prevailed (Herbert and Compton,
sulfur (S) contents shed some light on the depositional condi- 2007). This conclusion led to the inference of a lacustrine origin
tions of the Dwyka Group and the overlying Prince Albert sedi- for this time interval in the entire basin, despite the most reliable
ments in the foredeep depozone (platform facies of Visser, 1986, paleogeographic reconstructions, which indicate the presence of

A B
North South
EDEN-BURG TROMPS-BURG GRAAFF-REINERT WOLWEFONTEIN

G
Si S G
SG G Si S
Si Si S Whitehill Formation

Prince Albert Formation

VS DS-3

200 m

Black shales (Fl) Clast-poor diamictite


DS-2
Shales (Fl) and fine-grained
Massive diamictite (Dmm)
sandstones (Sr)

Dropstone shale (Fld), thinly bedded


Diamictite with sste. bodies
diamictite (Dms) and shales (Fl)
DS-1
Stratified diamictite (Dms) Diamictite (Dmm) with boulder beds

0 100 km
Borehole section
HS

Figure 7. North-south cross section A–B (see location in Fig. 6) across the Karoo Basin; datum at the base of Whitehill Formation (modified after
Visser; 1997a). Deglaciation sequences (DS) and facies associations from Visser (1997b). Note that uppermost part of the DS-1 and DS-2 are
represented by the dropstone shale facies association, herein interpreted as the early part of the transgressive systems tract (TST). In contrast, the
TST of the DS-3 (Prince Albert Formation) shows both its early phase (dropstone shale facies association) and late phase (shales with subordinate
fine-grained sandstones). Grain-size scale in logs: Si—silt; S—sand; G—gravel. HS—horizontal scale; VS—vertical scale.
Transgressions related to the demise of the Late Paleozoic Ice Age 13

a large inland sea or an interior seaway in a foreland-backarc set- A 0 1


S total (wt %)
2 3
ting (Von Brunn and Gravenor, 1983; Visser, 1989, 1991; Veevers
et al., 1994; Johnson et al., 1996). The existence of a large inland -60
sea with fluctuating salinity during Dwyka and Prince Albert -65
Tieberg Formation
times is also consistent with the apparent discrepancy of some -70
geochemical results. -75
Deglaciation is also indicated by an increase of organic mat-
-80

Depth (m )
ter content during a transgression. TOC values have been used Whitehill Formation
to differentiate the onset of the deglaciation phase in the Karoo -85
basin; TOC values up to 7% were reported for the lowermost -90
part of the Prince Albert by Scheffler et al. (2006). The organic- -95
rich part of the postglacial transgression also was identified by -100 Prince Albert Formation
Faure and Cole (1999) in the DP1 borehole (see Fig. 6 for loca- -105
tion), with values up to 12% TOC (Fig. 8A). TOC versus S plots
-110
show a wide range of salinities ranging from marine to prevail-
ing nonmarine conditions (Fig. 8B). High V/Cr ratios in concert -115
0 2 4 6 8 10 12 14 16 18 20 22
with increasing TOC values indicate anoxic bottom-water condi-
tions at the onset of the postglacial transgression (Scheffler et al., TOC (wt %)
2003, 2006). The accumulation of phosphatic sediments points
to high bioproductivity resulting from elevated nutrient sup- B
ply. The final retreat of the glaciers is recorded in the southern 4 Tasmanites bed
(Tasmania)
and eastern Karoo Basin by a concomitant rise of the chemical Price Albert Formation
index of alteration (CIA) and Rb/K ratios (Scheffler et al., 2006). 3
(Karoo Basin)
Predominantly full marine conditions prevailed, and the estab- Black Rock Member
S total (wt %)
(Falkland Islands)
lishment of anoxic conditions is indicated by the occurrence of
TOC-rich sediments containing mineral phases such as apatite 2
and pyrite. Additional evidence in favor of marine conditions
during the deglaciation phase is provided by the Rb/K ratio, used 1
as a salinity proxy. Higher Rb/K ratios during the late phase of
each deglaciation sequence in the Dwyka Group and during the
0
lowermost Prince Albert suggest increasing salinity during those 6 10 12 14 16
0 2 4 8
time intervals (Scheffler et al., 2006). Furthermore, the CIA and
TOC (wt %)
Rb/K high values suggest increased salinity during the last phase
of Visser’s (1997b) deglaciation sequences (DSs) II, III, and IV Figure 8. Geochemical characteristics of postglacial transgressions.
(Scheffler et al., 2003). (A) Total organic carbon (TOC) content (wt%) and sulfur (S) content
The northeastern part (the shallow backbulge depozone) of (wt%) in borehole DP1 (Faure and Cole, 1999). See text for discussion.
(B) TOC content (wt%) vs. S content (wt%) plot; the field at the lower
the Karoo Basin became ice free as a result of the collapse of the right is considered fresh water, and the field at the upper left is consid-
marine ice sheet. Based on studies along the northeastern margin ered marine. Values for tasmanite beds from Revill et al. (1994).
of the basin and analogies with Pleistocene deglaciation models,
Haldorsen et al. (2001) identified three glacial phases: an initial
phase of fast (a few thousand years) marine ice retreat of 400 km
over the northeast with glaciers grounded in the shallower areas of the basin in Natal, Orange Free State, and Transvaal (Fig. 6),
around the shore of the basin, followed by a second phase char- where the Vryheid Formation (Artinskian, Cairncross, 2001)
acterized by smaller continental ice sheets that remained more forms a clastic wedge that pinches out basinward (Tankard et al.,
or less stationary for several tens of thousands of years; during 1982). Coals of the Vryheid Formation rest directly on basement,
this second phase, massive glaciomarine muds with dropstones Dwyka glaciogenic sediments, or marine postglacial shales (Piet-
accumulated in the open basin; the last phase corresponds to ermaritzburg Formation). Where present, the underlying Dwyka
final deglaciation, which might have taken 10–20 ka. According sediments contrast drastically with the basinward platform facies
to Haldorsen et al. (2001) the entire Early Permian deglaciation association of Visser (1986, 1989); the latter is characterized by
of the northeast margin of the basin might have been completed massive and homogeneous diamictites (Dmm); the former asso-
within thousands rather than millions of years. ciation (the valley/inlet facies association, Visser, 1986) is repre-
The valley-glacier-retreat facies association is represented sented by varied lithologies (with a higher participation of sandy
by the Vryheid Formation along the northern cratonic margins matrix-supported conglomerates and associated sandstones) rest-
of the Karoo Basin, particularly on the northeasternmost corner ing on an irregularly dissected paleosurface (Von Brunn, 1996).
14 López-Gamundí

The thickest (>5 m) coals in the Vryheid Formation are the lowest et al., 2000). Siliceous and phosphatic concretionary zones with
ones; they thin and pinch out against relatively steep-sided walls fish remains and invertebrate fossils (gastropods, brachiopods,
of paleovalleys scoured by glacial action (Cairncross, 1989). echinoids, crinoids, foraminifers, radiolarians), and fallout tuff
Cold-temperate Glossopteris-Gangamopteris flora is present in beds are present in the Ganigobis Shale Member, where TOC
the floral assemblages of the lowermost peat (coal) swamps. Pul- contents between 0.5 and 1.4% and an S content of up to 1%
sating glacial retreat produced meltwater discharge that reworked have been reported (Grill in Stollhofen et al., 2000). A similar
unconsolidated till and introduced coarse, gravelly detritus. Thus, vertical facies stacking is present in DS-III ( ~130 m thick, Stoll-
the lowermost coals are associated with conglomerates and sand- hofen et al., 2000) with rarely preserved massive diamictites at
stones deposited in braided plains by low sinuosity rivers (Le the base, followed by crudely bedded rain-out diamictites and
Blanc Smith and Eriksson, 1979; Cairncross, 1980). Some pale- massive to cross-bedded sandstones that grade up into IRD-rich
otopographic lows in the basement were sheltered from the coarse sandy mudstones, and finally offshore IRD-free mudstones with
glaciofluvial influx, and lacustrine shales (rhythmites) and del- bivalve shells of the Eurydesma fauna (Hardap Shale Member).
taic sediments capped by thin coals were deposited. These low- The DS-IV is thinner (maximum thickness of 70 m) and made up,
ermost coals are capped by an extensive transgressive phase of from base to top, of sediment gravity-flow sandstones followed
shales and siltstones. Trace fossils (i.e., Skolithos and Planolites) by IRD-rich, sandy mudstones that grade upward into IRD-poor
described from cores in bioturbated levels between coal seams mudstones with thin interbeds of turbidite sandstone.
of the Vryheid Formation in Transvaal have been interpreted as
part of abandonment phases in deltaic settings (Stanistreet et al., Paraná Basin
1980). Glauconite pellets, interpreted as glauconitized fecal pel- In the pre-drift neighboring Paraná Basin (Fig. 1) glaciogenic
lets, are associated with one of the bioturbated layers, suggesting sediments of glacial episode III correspond stratigraphically to
a marine influence during the abandonment phase. According to the Itararé Group. Gravenor and Rocha-Campos (1983) defined
Cairncross (1989), two factors related to the postglacial trans- a marine-freshwater facies in the Itararé deposits. It consists of
gression could have contributed to the preservation of relatively well-stratified thin-bedded diamictites (Dms) intercalated with
thick coals at the base of the Vryheid Formation. Bacterial and sandstones and shales with IRD, massive diamictites in beds up
microbial activity may have been suppressed by the cold tem- to 1 m thick interbedded with shales, rhythmites, and mudstones,
peratures, and the destruction of peat was not intense, allowing some with carbonate nodules. Marine fossils are present in the
thicker coal-seam formation. Also, meltwater discharge prob- shales and mudstones. This facies is considered by Gravenor and
ably promoted a high water table (cf. Bohacs and Sutter, 1997) Rocha-Campos (1983) as having been deposited in a basinal envi-
that would, in turn, have maintained the swamps and enhanced ronment during deglaciation, and is capped by marine shales. At
peat preservation. least four stratigraphically different basinwide marine “horizons”
Farther north in the Karoo-age rifts of Namibia, such as in have been recognized in the basin (Rocha-Campos and Rösler,
the southeastern Kalahari Basin (9 in Fig. 1), anoxic conditions 1978), and they are probably similar to the final stage of the
were established during the postglacial marine transgression, as deglaciation sequences described for the Karoo Basin (cf. Fig. 7)
evidenced in the lower Ecca shales (Tswane Formation). The by Visser (1997b). The basin margins show unequivocal evidence
organic-rich shales, which are carbonaceous in places, contain of glacier advances and retreats. During the late Early Permian,
up to 50% TOC (Scheffler et al., 2006). The shales are underlain marine conditions expanded in the basin, and a final postglacial
by fluvioglacial and glaciolacustrine deposits (Visser, 1997a). eustatic flooding is recognized by the widespread occurrence of
These deposits can be assigned to the valley-glacier-retreat facies relatively thin, fossiliferous marine beds in the uppermost part of
association that dominates the updip portions of the glaciated the Itararé sequence (dos Santos et al., 1996). Three settings for
basins. Farther basinward in the Huab and Kalahari Basins (8 and the glacial-postglacial transition in the Paraná basin were defined
9, respectively, in Fig. 1) the four deglaciation sequences (DSs) by dos Santos et al. (1996). The terrestrial setting is character-
defined by Visser (1997b) for the main Karoo Basin have been ized by subglacial diamictites with evidence of subglacial-style
identified (Bangert et al., 1999; Stollhofen et al., 2000). Maxi- soft sediment deformation, massive or cross-bedded fluvial sand-
mum flooding intervals have been identified as IRD-free dark stones, and thin coal layers. The valley setting consists of dia-
mudstones, locally known as the Ganigobis and Hardap Shale mictites as part of the basal fill of valleys controlled by basement
Members for DS-II and DS-III, respectively. The DS-II (with a fault zones and probably scoured and deepened by the ice, fluvial
maximum thickness of ~200 m) consists of a fining-upward suc- sandstones, massive or with cross bedding, and lacustrine mud-
cession of basal massive diamictites and outwash conglomerates stones and rythmites with IRD. In places this glacial terrestrial
that pass upward to rain-out diamictites and associated sandy facies association is overlain by mudstones and siltstones with
gravity-flow deposits that grade into IRD-bearing mudstones. marine acritarchs and Tasmanites. Botryococcus is present in this
The succession is capped by the Ganigobis Shale Member. This postglacial section. The terrestrial and valley settings of dos San-
unit is made up of dark, IRD-free mudstones deposited by off- tos et al. (1996) broadly correspond to the valley glacier retreat
shore suspension settling, interbedded with scarce, thin beds of facies association described in this contribution. The proximal
fine-grained sandstones interpreted as distal turbidites (Stollhofen glaciomarine setting consists of thin, discontinuous, subglacial,
Transgressions related to the demise of the Late Paleozoic Ice Age 15

massive diamictites (Dmm), subglacial meltout-fluvioglacial Formation (HST). A similar sequence stratigraphic framework
coarse-grained conglomerates, debris flow, stratified diamictites was proposed by Vesely and Assine (2004) in their study that
(Dms[r]) and rhythmites with dropstones (França et al., 1996; incorporated both outcrop and well-log data. They identified
Rocha-Campos et al., 2008). This facies association grades up five depositional sequences bounded by disconformities trace-
into the overlying postglacial, transgressive marine shales of the able over 400 km across the basin in an E-W depositional strike
Rio do Sul Formation. Thin, transitional sequences between the section. Three successive facies associations were defined within
rhythmites and marine shales consist of shallow-marine, het- each sequence, but the lower and upper ones are absent in places.
erolithic sequences of siltstones and sandstones with lenticular The lower facies association occurs only in the two lowermost
bedding and intersecting wavy laminae indicative of wave action sequences and is made up of subglacial facies (mostly Dmm);
(Canuto, 1993, in dos Santos et al., 1996). Dropstones are com- it corresponds to a glacial-maximum systems tract as defined by
mon in the lower few meters of marine shales, but they gradu- Vesely and Assine (2004). This tract grades upward into con-
ally disappear upward. This transgressive facies association glomerates and sandstones, which in turn are overlain by dia-
culminates the retrogradational pattern punctuated by short-term, mictites, turbidites, and shales with dropstones, labeled by the
transgressive-regressive episodes that characterize the deglacia- authors as a deglaciation systems tract (equivalent to the trans-
tion process (dos Santos et al., 1996). gressive systems tract of Canuto et al., 2001). The deglaciation
A glacial terrestrial setting is characterized by subglacial facies rests directly on the bounding erosional surfaces where
diamictites with evidence of subglacial-style soft sediment defor- the subglacial facies of the glacial maximum system is absent
mation, massive or cross-bedded fluvial sandstones, and thin coal basinward. The fine-grained laminated facies of the upper part of
layers. Along the basin margins the glacial facies are followed the deglaciation systems tract represents the record of maximum
by littoral and deltaic sequences interpreted as basinward progra- glacial retreat during interglacial periods. Highest gamma-ray
dational wedges (dos Santos et al., 1996), where eventually par- readings were used to determine the stratigraphic position of the
alic peat swamps developed. These progradational, commonly MFSs within the deglaciation systems tracts.
coal-bearing sequences have been identified along the northern
(Martini and Rocha-Campos, 1991), western, and eastern mar- Falkland-Malvinas Islands
gins (Buatois et al., 2006). Dos Santos et al. (1996) speculated Similarities between the late Paleozoic glacial-postglacial
about the correlation between these coal-bearing zones and the sections in the Karoo Basin and the Falkland Islands (Fig. 9)
postglacial marine “horizons.” were first suggested by Adie (1952). Later paleomagnetic
Lateral variability of sediment supply related to glacial activ- (Mitchell et al., 1986; Taylor and Shaw, 1989), radiometric
ity along the basin margins during glacial and early postglacial (Mussett and Taylor, 1994), and paleocurrent (Storey et al., 1999;
times could be the principal factor accounting for the different Trewin et al., 2002) data further confirmed Adie’s original cor-
number of sequences identified. However, the vertical arrange- relation. The pre-breakup reconstructed position of the Falkland
ment of each sequence seems to be strikingly similar. Canuto microplate east of the coast of South Africa (Figs. 1 and 6) is
et al. (2001) described seven fining- to coarsening-upward the result of an eastward translation and a 180° rotation from
sequences for the Itararé Subgroup in their outcrop-based study its present location (Marshall, 1994). The glaciogenic deposits
along the eastern margin of the basin. These sequences range are grouped under the Lafonia Formation (Frakes and Crowell,
between 50 and 180 m in thickness; their base is characterized 1967; Scasso and Mendía, 1985; Bellosi and Jalfin, 1987), and
by an erosional surface (striated bedrock or intraformational stri- the overlying postglacial sediments are part of the Port Sussex
ated pavements). Thin (>1 m), laterally discontinuous beds of Formation (Frakes and Crowell, 1967). The contact between the
compacted massive diamictites (Dmm), interpreted as lodgment uppermost Lafonia diamictites and the lowermost member of
tills, rest on the striated surfaces. This basal facies is followed the Port Sussex Formation (Hells Kitchen Member) was char-
by tabular beds of massive diamictites (Dmm), locally bedded acterized as an “abrupt undulating surface with a local relief of
(Dms) and interpreted as subglacial ablation tills associated 0.25 m” by Trewin et al. (2002), who also note that the member
with cohesive debris flow beds represented by massive to bed- thickness ranges from 3.5 to a maximum of 10 m and is made up
ded diamictites (Dms[r]) with soft sediment deformation. This of a coarsening-up cycle grading from “black fissile mudstone to
facies association is grouped in an LST by Canuto et al. (2001). fine- and medium-grained sandstones with granules and pebbles
The LST is followed by sandstones, siltstones, and claystones deposited as dropstones” (p. 9). Trewin et al. (2002) interpret the
with dropstones (Fld) of a TST, which, in turn, is overlain by contact between the Lafonian diamictite and the base of the Hells
deltaic sandstones stacked in a progradational pattern (highstand Kitchen Member as an erosional event that probably developed
and regressive glacio-isostatic systems tracts). The uppermost under shallow-water conditions during a transgression related
of these sequences (S7) represents the period of final degla- to deglaciation. This contact is considered herein as a transgres-
ciation in the Paraná Basin. This glacial-postglacial transition sive ravinement surface (TRS) that initiated the postglacial TST.
lies stratigraphically between the uppermost part of the Itararé Trewin et al. (2002, p. 9) suggest that the Hells Kitchen Member
Group (Rio do Sul Formation), containing the LST and TST, sediments were deposited “probably during a transgression asso-
and the basal Triunfo Member of the superjacent Rio Bonito ciated with deglaciation” and compared the section with a similar
16 López-Gamundí

SG
Si
SG
Si

Ripon
Bonete Fm.
G
Si S
Fm.
Highstand
systems tract
Collingham Fm.
Piedra Whitehill Fm.
Azul Fm. Port Transgressive
Prince Sussex Fm. systems tract
Albert Fm.

Lowstand
Dwyka Lafonia
Sauce Fm. Fm. systems tract
Grande
Fm.

100 m

Black shales (Fl) Massive diamictite (Dmm)

Shales (Fl) and fine-grained


sandstones (Sr) Coarse sandstone

Dropstone shale (Fld), thinly bedded


diamictite (Dms) and shales (Fl) Fine to medium sandstone

Stratified diamictite (Dms) Basement

Figure 9. Correlation of the postglacial mudstones (glacial epsiode III) across the Ventana fold belt (V), Karoo Basin (K), and Falkland
Islands (FI). Grain-size scale in logs: Si—silt; S—sand; G—gravel.

facies arrangement described by Visser (1993) for the base of the glaciogenic deposits by the Late Carboniferous (Keidel, 1916;
Prince Albert Formation in the Karoo Basin. Harrington, 1947; Coates, 1969; Amos and López-Gamundí,
The Hells Kitchen Member is overlain by interbedded black, 1981b). This is approximately the same time interval when sedi-
pyritic, partly carbonaceous shales and siliceous mudstones mentation, dominated by glacial deposits as well, began also in
(Black Rock Member, with a minimum thickness of 120 m); the Chaco-Paraná and Paraná Basins (Fig. 1). The basal fill of the
TOC content in the carbonaceous shales ranges ~2%–3% in the Gondwana cycle in both basins is characterized by glacial depos-
east to exceptionally up to 40% toward the west (Trewin et al., its, mostly nonmarine in the Chaco-Paraná Basin (Russo et al.,
2002). TOC versus S plots for two samples of the Black Rock 1987; Fernández Garrasino, 1996; Winn and Steinmetz, 1998)
Member indicate fresh-water salinities (Fig. 8B). and predominantly marine in the Paraná Basin (Rocha-Campos
and dos Santos, 1981; Zalán et al., 1990; França and Potter, 1991;
Sauce Grande Basin–Ventana Fold Belt França, 1994; Rocha-Campos et al., 2008).
Keidel (1916) first recognized the similarities between the Late Paleozoic glacial deposits exposed in the Ventana fold
Cape fold belt and the adjacent Karoo Basin in South Africa, belt are grouped under the Sauce Grande Formation. Equivalent
and the Ventana fold belt and contiguous Sauce Grande Basin diamictites and overlying finer grained units (Piedra Azul and
in Argentina (Fig. 9). The similarities are particularly striking Bonete Formations) also have been reported from offshore wells
for late Paleozoic times, when the overall basin evolution and (Lesta et al., 1980; Amos and López-Gamundí, 1981b; Fryklund
subsidence histories in both regions are considered (Du Toit, et al., 1996; Juan et al., 1996). The age of the Sauce Grande For-
1927; López-Gamundí and Rossello, 1998). Sedimentation in the mation is poorly constrained as pre–Early Permian by the pres-
Sauce Grande Basin adjacent to the Ventana fold belt began with ence of Eurydesma fauna and Glossopteris flora in the Bonete
Transgressions related to the demise of the Late Paleozoic Ice Age 17

Formation and post–mid-Carboniferous on the basis of palyno- a feature typical of extensional rift basins. A significant strike-
logical studies in equivalent sediments found in offshore wells slip component has been proposed for these basins (Chakraborty
(Archangelsky, 1996). The base of the Sauce Grande Formation et al., 2003; Chakraborty and Ghosh, 2005).
lies with regional unconformity on Devonian metasedimentary The Talchir Formation in the Indian peninsular basins
basement. The unit is mostly made up of massive to crudely strat- (Fig. 10) shows evidence of sedimentation under glacial influ-
ified diamictites with subordinate rhythmites, sandstones with ence (Ghosh and Mitra, 1975; Casshyap and Srivastava, 1987;
ripples, and scarce conglomerates toward the top (Andreis et al., Veevers and Tewari, 1995). Key evidence of glacial influence
1987). The massive (Dmm) and stratified (Dms) diamictites have during Talchir sedimentation is provided by striated pavements
been interpreted as glaciomarine (rain-out tills) partially remo- on basement rocks and boulder pavements, striated and bullet-
bilized downslope by gravity flows (Dms[r]) (Coates, 1969; shaped boulders in diamictites, and dropstone shales (Baner-
Andreis, 1984; Andreis and Torres Ribeiro, 2003). The marine jee, 1966; Ahmad, 1975) Marine faunas have been reported at
setting is confirmed by the presence of a solitary marine bivalve or near the top of the glaciogenic Talchir Formation in several
in the diamictites (Harrington, 1955). Gondwanan basins of India (see Veevers and Tewari, 1995, for
Detailed sedimentological studies by Andreis and Torres a review). These findings may be related to a marine transgres-
Ribeiro (2003) allowed subdivision of the Sauce Grande Forma- sion, interpreted by several authors as having been caused by a
tion in three megacycles. The lower megacycle (maximum thick- eustatic rise in sea level owing to deglaciation. Previous interpre-
ness, 700 m) is composed of abundant diamictites, sandstones, tations, however, originally considered that the glacial sediments
and scarce conglomerates. The middle megacycle (~50 m thick) and finer grained postglacial sediments were deposited at the
contains only sandstones and conglomerates. The upper cycle margin of lake basins, despite the rather gradational passage to
(~350 m thick) is made up of abundant fine- to coarse-grained marine deposits with abundant invertebrate fauna dominated by
sandstones; thick-bedded, massive, clast-poor, muddy diamictites bivalves in the Rewa Basin (Ghosh, 1954), Daltonganj-Rajhara
(Dmm); thin-bedded stratified diamictites (Dms[r]); and shales, Basin (Dutt, 1965, in Goswami, 2008), and Bokaro Basin (Sen-
with scattered dropstones. The Dmm facies is interpreted as rain- gupta et al., 1999) (Fig. 10A). The most distinctive element of
out tills, whereas the Dms(r) facies is considered glacial material these marine faunas is Eurydesma, a cold-water bivalve genus
remobilized downslope as gravity (mostly debris) flows. Andreis widespread in the Early Permian of the Gondwana Superconti-
and Torres Ribeiro (2003) interpret the finer grained nature (evi- nent (Harrington, 1955; Dickins, 1961; Dickins and Shah, 1977;
denced by increasing participation of shales and sandstones to the Runnegar, 1979). An inspection of the section described by Gosh
detriment of diamictites, particularly of the clast-rich subtype) (2003) in the Satpura Basin (Fig. 10B) suggests that the black
of the uppermost megacycle as a response to the transgression shales and subordinate, decimeter-scale limestone beds that rest
caused by glacier retreat that continued during Piedra Azul times. on the glaciogenic Talchir diamictites can be interpreted as part
Dickins (1984) also relates the Piedra Azul transgression to the of the postglacial marine transgression rather than as the result of
early Sakmarian glacioeustatic sea-level rise. The facies types sedimentation of predominately fine-grained (clay and silt) mate-
and their vertical stacking in the Sauce Grande upper megacycle rial and carbonate muds in a basinal lacustrine environment with
show similarities to the deglaciation sequences described by Vis- a sporadic marine incursion, containing a marine fauna charac-
ser (1997b) for the Karoo Basin (cf. Figs. 7 and 9). terized by Eurydesma and presence of other bivalves, ostracodes,
The Sauce Grande sediments grade upward to shales, biotur- and foraminifers. Marine acritachs have been identified in several
bated mudstones with gastropods, and subordinate fine-grained basins and, in a few cases, associated with Eurydesma (Venka-
sandstones with wave and current ripples (Piedra Azul Forma- tachala and Tiwari, 1988). Associated plant remains also have
tion), passing upward into bioturbated mudstones with fossils been reported in several localities (Gosh, 2003). The postglacial
of the Eurydesma fauna (Harrington, 1955; Rocha-Campos and section in the Satpura Basin is a marine, deep-water facies asso-
Carvalho, 1975; Amos, 1980), and fine-grained sandstones with ciation; isolated ripple trains can be interpreted as starved ripples
wave ripples and cross-bedding (Andreis et al., 1979). Coarsen- produced by weak bi- or unidirectional bottom currents. This find-
ing upward parasequences (up to 40 m thick each) are common ing extends also this marine embayment of the Indian Peninsula
in the Piedra Azul Formation; they consist, from bottom to top, of farther west from the areas where marine faunas were originally
mudstones, fine-grained sandstones with wave ripples, and chan- described (Rewa, Daltonganj-Rajhara, Bokaro Sub-basins of the
nelized medium- to coarse-grained sandstones with trough cross- Damodar Basin, Fig. 10A). The alternative interpretation should
stratification (Andreis and Japas, 1996). invoke a setting in which deep lacustrine facies are systematically
interrupted by open-marine shales with Eurydesma fauna depos-
Gondwana Basins of Peninsular India ited below wave base without any intervening facies.
The Gondwana basins of Peninsular India (Fig. 1) have been Bose at al. (1992) studied the sedimentary features of the Tal-
considered fault bounded troughs developed along preexisting chir sandstone facies in the West Bokaro Sub-basin (Fig. 10A) and
zones of weakness imparted by Precambrian structural fabrics interpreted them as the product of coarse-clastic sedimentation
(Naqvi et al., 1974). The basin fill is commonly asymmetric, with “in fjord-like glacier-fed coastal troughs” (p. 95). Supporting evi-
an overall increase in thickness toward one of the boundary faults, dence for a fjord-like setting with low salinity from a fresh-water
18 López-Gamundí

Micritic
limestones Hummocky
cross-stratification

78°E
Son
86°E West
Bokaro
Rahmajal B Shales /
mudstones
Trough
Daltonganj-Rajhara Fine-grained cross-stratification
N Rewa Basin sandstones
24°N ∗ ∗ ∗ Massive Starved ripples
∗ Jahria Raniganj diamictites
(Dmm)
Satpura Basin Damodar Basin
Burrows
Son Basin
Mahanadi
∗ Basin
Paleocurrents
Marine fauna
B (Eurydesma)
Raipur
Talchir
Bedasar
Irai
Basin Barakar
20°N
Fm.

Bay

Godavari of TST
Basin
Bengal

A
20 m
100 km
∗ Localities with marine fauna Talchir
Fm.
Si S Si S
G G

Figure 10. (A) Gondwana deposits in Peninsular India. Note presence of marine fossils (Eurydesma fauna) in postglacial mudstones (glacial
episode III). (B) Section of the glacial-postglacial transition in the Satpura Basin. Modified from Ghosh (2003). Grain-size scale in logs: Si—silt;
S—sand; G—gravel.

contribution from retreating glaciers comes also from ichno- in the late TST is represented by the mudstone facies. A similar
logical studies. Bhattacharya and Bhattacharya (2007) studied interpretation has been proposed for the upper Talchir interval
the ichnofacies associated with these postglacial intervals in the in the adjacent West Bokaro Sub-basin (Fig. 10A) by Bhatta-
Raniganj Sub-basin (Fig. 10A) and concluded, on the basis of charya et al. (2005). The basal Talchir glaciogenic section is
low ichnodiversity, sporadic distribution of the traces, small bur- made up of breccias, matrix- and clast-supported conglomerates,
row dimensions, absence of any body fossils, and dominance of and coarse-grained sandstones (their conglomerate-sandstone,
worms and annelids as trace-makers, that stressed environmental TCS, facies association) and is followed upward by fine-grained
conditions (cold climate and low marine salinity) prevailed. These sandstones with hummocky cross-stratification, sandy siltstones
conditions are assigned to the influx of glacier meltout fresh water with dropstones, and subordinate clast-supported conglomerates
in an ice-marginal sea during climatic amelioration and degla- (sandstone-siltstone, TSS, facies association). This fining-upward,
ciation. The postglacial TST exhibits a deepening-upward facies retrogradational section culminates with alternating thinly bed-
trend initiated with (1) mudstones with dropstones; (2) interbed- ded fine-grained sandstones and mudstones with marine fos-
ded laminated to massive siltstones, mudstones, and fine-grained sils (mollusks) and thick, multistoried, fine-grained sandstones
sandstones with current and combined-flow ripples and hum- with common hummocky and swaley cross-stratification (fine
mocky cross-stratification; and (3) overlying IRD-free, massive, sandstone–mudstone, TSM, facies association). Bhattacharya
locally bioturbated mudstones. Nereites and Zoophycos ichno- et al. (2005) interpret this retrogradational stacking pattern as the
assemblages are common in the two latter facies (Bhattacharya result of progressive deglaciation during a eustatic sea-level rise
and Bhattacharya, 2007). These facies with abundant trace fossils and deposition of shelf sediments under a transgressive phase for
are conspicuously IRD-free and indicative of a retreat of the ice- the TSS and TSM facies associations.
grounding line with a concomitant influx of glacier meltwater Geochemical evidence is also supportive of an environment
into the basin and subsequent lowering of the marine salinity. with significant fresh-water contribution. Bhattacharya et al.
The facies of siltstones, mudstones, and fine-grained sandstones (2002) studied the geochemistry of calcareous nodules in fine-
is interpreted as the product of sedimentation in an open shelf grained sediments (siltstones, rhythmites) that rest on the coarse
influenced by waves, whereas the overlying mudstone facies rep- glaciogenic basal part of the Talchir Formation in the Damodar,
resents offshore background sedimentation from suspension of Mahanadi, and Godavari Basins (Fig. 10A). Four nodules of sim-
clay-silt material. It is here proposed that the early TST is rep- ilar composition (micritic) from the uppermost Dwyka tillite of
resented by the facies of mudstones with dropstones; the MFI the Karroo Basin were also analyzed for comparative purposes.
Transgressions related to the demise of the Late Paleozoic Ice Age 19

The oxygen and carbon isotopic ratio (δ18O and δ13C) values bonate, CaCO3 · 6H2O (Suess et al., 1982), and seem to occur in
from both Talchir and Dwyka samples were similar and indicate those mudstone zones rich in organic matter derived from the
a fresh-water environment of formation. unicellular alga Tasmanites. This mineral is suggested to be an
authigenic precipitate, forming at low temperatures from inter-
Tasmania Basin stitial waters of organic-rich sediments, undergoing microbial
During the late Paleozoic, Tasmania (Fig. 1) was positioned degradation, accumulating rapidly in cold bottom waters (Suess
close to polar latitudes ~70° (Late Carboniferous) and ~80° et al., 1982; Shearman and Smith, 1985). High alkalinity from
(Early Permian) (Scotese and Langford, 1995; Li and Powell, decomposing organic matter also enhanced precipitation of ikaite
2001). The Tasmania Basin (Fig. 1) was glaciated throughout the (Bischoff et al., 1993; DeLurio and Frakes, 1999). Glendonites
Late Pennsylvanian (Clarke and Forsyth, 1989; Dickins, 1996). in the Sydney Basin (Fig. 1) are also commonly associated with
Ice-flow indicators suggest ice centers on the west and northwest organic-rich shelf facies with dropstones (Thomas et al., 2005;
and depocenters on the east. Local topographic highs created a Selleck et al., 2007).
fragmented shelf where glacially influenced sedimentation took The layers rich in Tasmanites (named tasmanite beds) are
place (Clarke and Forsyth, 1989; Hand, 1993). There is consen- laminated and occupy a consistent stratigraphic position ~10–
sus (Banks, 1981; Clarke and Forsyth, 1989; Hand, 1993) that the 30 m above the contact between the mudstone-prone postglacial
glaciomarine diamictite and associated rhythmites of the Tasma- interval and the underlying diamictic section (Figs. 11 and 12,
nia Basin were deposited in a fjord-like seaway characterized by Domack et al., 1993). Pyrite nodules and dropstones are abun-
rain-out and settling of fines with minor coarse debris deposited by dant in the tasmanite beds. TOC values for the tasmanite levels
rafting and turbidity currents originating from the glacier ground- in two boreholes (Douglas River and Ross 1) reach exceptionally
ing line (cf. Bartek and Anderson, 1991). The glaciomarine sedi- ~20%, with high hydrogen indexes (HI = 868 mg HC/g TOC);
ments pass upward to marine siltstones and mudstones (Woody even higher HI values >900 mg HC/g TOC have been reported
Inglis Siltstone, Quamby Mudstone, and equivalent units) with from other localities (Fig. 14, Revill et al., 1994). These high HI
abundant glendonite concretions and elements of the Eurydesma values indicate that the kerogen contains hydrogen-rich type I
fauna (Clarke and Banks, 1975). This interval has been assigned organic matter (Tissot and Welte, 1984). Sulfur values obtained
to a widespread marine transgression that covered most of Tas- rarely exceed 2%; when plotted against TOC content they fall
mania as glaciers retreated (Hand, 1993) during the early–middle on both fields (marine sediments and fresh-water sediments),
Sakmarian (Brakel and Totterdell, 1993). These fine-grained indicating conditions of variable salinity (Fig. 8B). As indicated
deposits are overlain by the Bundella Formation (late Sakmar- by Domack et al. (1993), the association of glendonite with the
ian), which includes the Darlington Limestone. Dropstones and TOC-rich tasmanite levels suggests a marine origin and also low
glendonite concretions are abundant within the limestone (Rogala temperatures (close to freezing), a condition required for the
et al., 2007); a high-abundance, low-diversity Eurydesma fauna formation of the precursor of glendonite (ikaite) (Shearman and
of calcareous invertebrates (mostly brachiopods, bryozoans, and Smith, 1985). Rao and Green (1982) estimated independently a
Eurydesma bivalves) has been identified (Clarke and Forsyth, sea surface temperature for Early Permian Tasmania of –1.8 °C,
1989; Rogala et al., 2007). The limestones consist of bioclastic close to the present average near the Antarctic ice shelf of –1.9 °C.
floatstones, rudstones, and grainstones deposited in neritic shelfal Cold water deposition is consistent with the low diversity of the
environments during sea-level highstands (Rogala et al., 2007). associated Eurydesma fauna (Clarke and Banks, 1975). The
Cold conditions persisted until the Late Permian, as evidenced organic-rich tasmanite beds in the Woody Island Formation and
by dropstones and cold-water Eurydesma fauna in Kungurian to equivalent units represent condensed sections within an MFI; a
Capitanian beds (Clarke and Forsyth, 1989). similar view was proposed by Rogala et al. (2007).
The glacial-postglacial transition in the Tasmania Basin has
been studied in outcrops and in subsurface (Figs. 11–13). Its con- DISCUSSION
tact is in general sharp (mudstones resting on mostly massive
diamictites), but in a few localities a distinctive facies between Framework of the Postglacial TSTs
the glaciogenic diamictites and the postglacial mudstones has
been identified. This facies consists of thin (0.5–1 m of bed Punctuation of long-term transgressions by repeated short-
thickness), stratified diamictites (Dms) with dropstone clusters term regressions is caused by the tendency of sediment supply
and pebble nests (Domack et al., 1993). These thin diamictite rates to outmatch, for short periods, accommodation increase
beds have flat, nonerosive lower contacts, and alternate with the rates (Cattaneo and Steel, 2003). This is a particularly common
isolated pebbly mudstones and dark gray to black mudstones; pattern in most transgressions during periods devoid of glacia-
eventually the mudstones predominate. The postglacial black tions. This punctuated shoreline movement has a complex zigzag
mudstones contain scattered dropstones; framboidal pyritic and shoreline trajectory (Helland-Hansen and Gjelberg, 1994) that
large (up to 10 cm) calcareous concretions (glendonites) also results in a retrogradational (or backstepping) stacking pattern
have been identified (Domack et al., 1993). Glendonites are cal- (retrogradational parasequence set, Van Wagoner et al., 1990). As
cite pseudomorphs after ikaite, a hydrated type of calcium car- part of the Pleistocene Ice Age, the late Quaternary transgressive
20 López-Gamundí

Trunbridge Ross-2 Ross-1 Douglas River


G SG SG
Si S Si Si
0

60 G

G
120

G G

180
G
Depth (meters)

240

300

360

Ross-2 Douglas
River
Tunbridge
420 Ross-1
Shales (black mudstones) Diamictites

Silty mudstones Basement


Eaglehawk
480 Neck
Tasmanite beds Dropstones (IRD)
100 km
G Glendonites

Figure 11. Cross section showing stratigraphic position of tasmanite levels within the postglacial mudstone section in the Tasmania Basin. From
Domack et al. (1993).

deposits may represent an unusual record of high-frequency and variations within this theme are possible along a continental mar-
high-amplitude sea-level oscillations driven by glacioeustasy, gin subjected to deglaciation, as sediment supply may be later-
and therefore better analogues for the sections previously dis- ally variable in a basin, the rate of sea-level rise is significantly
cussed. Rather, the postglacial transgressions analyzed herein higher than any other process. In that sense the postglacial TST
seem to be related to abrupt upward-deepening of facies and identified in the late Paleozoic Gondwanan basins can be equated
the scarcity of surfaces of ravinement (erosion by wave action), to the late Quaternary transgressive deposits resulting from an
culminating in a level of deepest facies, commonly termed the unusual record of high-frequency, high-amplitude sea-level rise
maximum flooding interval (MFI) or surface (MFS). Direct driven by glacioeustasy. The response to this type of postglacial
juxtaposition of offshore mudstones over glacial deposits (most retreat is the creation of a continuous transgression triggered by
commonly rain-out tills or less abundant subglacial tills) is not steady (and fast) sea-level rise (Curray, 1964; Cattaneo and Steel,
unusual. In other cases a diagnostic intervening facies associa- 2003). Experimental sequence stratigraphy provides insights into
tion dominated by thin-bedded debris flow deposits with IRD and some of the salient features of fast transgressions. Heller et al.
dropstone shales is present. This facies association is the result of (2001) showed that a rapid base-level-rise rate leads to an abrupt
the complex interaction between glacier-derived sedimentation, shoreline retreat with maximum flooding at the end of the max-
gravity (debris flow) currents, and rain-out deposition. Although imum base-level rise. They also noticed that the magnitude of
G
Si S
TOTAL ORGANIC CARBON (%)
0 2 4 6 8

290

290

300

300

310

310

320

320 330

0 2 4 6 8

G
Shales (Fl) with Tasmanites Mudstones with granules
and pebbles
330 Mudstones Stratified diamictites (Dms),
thinly bedded

Dropstone shales (Fld) Stratified diamictites (Dms)

Fine-grained sandstones Basement

G Glendonites

Figure 12. Douglas River borehole section (modified from Domack et al., 1993), with TOC (total organic car-
bon) values for the Tasmanites-rich interval. Grain-size scale in log: Si—silt; S—sand; G—gravel. See location
in Figure 11.
22 López-Gamundí

Figure 13. Cores from the glacial-postglacial


transition in the Tasmania Basin; see Figure 11
A for location. (A) Massive diamictite (Dmm),
B probably of rain-out origin at base of Ross-1
well. (B) IRD-dominated section, Eaglehawk.
(C) Stratified diamictites (Dms) with soft sedi-
ment deformation, Eaglehawk. (D) Organic-rich
shale, Tasmanites-rich interval, Douglas River.
Bar scale: 10 cm. IRD—ice-rafted debris.

D
C

transgression is far greater for the rapid base-level rise than dur- Even in the cases where postglacial offshore mudstones are
ing the slow rise, and that the time of maximum flooding is nearly directly resting over the glacially derived, diamictitic section, the
synchronous with the end of the rapid base-level rise. In cases interval with the highest TOC values is invariably slightly higher
with slow rises, maximum flooding takes place far earlier than at in the interval dominated by offshore mudstones; thus most max-
the end of the rise. imum flooding shales analyzed herein do not seem to be basal
Transgressions related to the demise of the Late Paleozoic Ice Age 23

1,000
I
900
Oonah
Latrobe
800
Douglas
700 II River
Hydrogen Index

600

500 Figure 14. Geochemical characteristics


of the tasmanite beds. From Domack
400 et al. (1993).

300 100 km

200
Oonah
Latrobe
100 III Upper seam
Douglas River Lower seam
0
380 400 420 440 460 480 500
T max °C

transgressive black shales sensu stricto (BT model of Wignall, Basin in Australia seem to be present in other basins of the Gond-
1991, 1994). The organic-rich, condensed sections in the Falk- wana Supercontinent affected by the LPIA. In Oman (Fig. 1) the
land Islands and Tasmania, a few meters above the contact with glaciogenic sequence (Lower Permian Al Khlata Formation) is
the diamictitic succession, are examples of this type of maximum- developed in a rift setting. It rests disconformably over Proterozoic
flooding black shales. In other, leaner postglacial mudstone sec- basement and consists of diamictites, conglomerates, sandstones,
tions the maximum flooding interval is inferred by the abundance and mudrocks ranging from glaciofluvial, glaciolacustrine, and
of marine fossils and dominance of fine-grained sediments (cf. alluvial to paralic environments (Levell et al., 1988; Al-Belushi
most sections in the Calingasta-Uspallata and Paganzo Basins, et al., 1996; Martin et al., 2008). Evidence of glacial abrasion is
the Piedra Azul interval in the Ventana fold belt, and the Lower provided in outcrops by striated surfaces on the dolomitic base-
Barakar Formation in several basins of Peninsular India). ment of late Proterozoic age (Braakman et al., 1982). A paleo–
ice flow from northeast to southwest was inferred by Al-Belushi
Postglacial TSTs through Space and Time et al. (1996). Southward in the subsurface of the South Oman
Salt Basin the Al Khlata diamictites were entirely deposited by
As noted by Andrews (1997) the stratigraphic record may rain-out and debris flow with no evidence for the preservation of
not be able to furnish the nuances of an unstable ice sheet at the true tillites (Aitken et al., 2004). The upper part of the Al Khlata
time of its disintegration, especially without precise chrono- Formation consists of a sheetlike diamictite abruptly overlain
logical control such as in the case of the LPIA. Thus, although by the Sakmarian Rahab shale, which includes varve-like lami-
appealing, it is somewhat incorrect to infer direct correlations nated mudrocks (rhythmites) with dropstones deposited in a large
between glacial histories and ice volume record. Despite this fresh-water to brackish-water body, according to Hughes Clarke
potential impossibility, the stratigraphic record of the late Paleo- (1988) and Levell et al. (1988). The dropstones progressively dis-
zoic glacial-postglacial transition shows at hierarchies equivalent appear upward, leading to Levell et al. (1988) to interpret this
to third-order cycles a remarkable consistency in terms of facies uppermost section of the Rahab shale as a deglaciation sequence.
associations, facies stacking patterns, and sequence-stratigraphic This unit is followed by the Saiwan Formation. In subsurface
framework evolution. the Saiwan Formation unconformably overlies fine-grained
sandstones and siltstones of the Rahab shale or rests directly on
Other Occurrences of Postglacial TSTs Related to the LPIA diamictites of the Al Khlata Formation. Its lowest part (lower
The common characteristics of the glacial-postglacial transi- member of the Gharif Formation) includes a restricted marine
tion highlighted in this contribution for the Calingasta-Uspallata interval with the acritarchs, termed the maximum flooding shale
and Paganzo Basins in western Argentina, the Karoo Basin in by Guit et al. (1995) and interpreted as a possible postglacial eus-
South Africa, the Gondwana basins of India, and the Tasmania tatic flooding event (Stephenson and Osterloff, 2002). The base
24 López-Gamundí

of the Saiwan Formation, dated biostratigraphically as late Sak- previous glacial outwash valleys and structural depressions) that
marian (Angiolini et al., 2003), includes bioclastic sandstones enhanced restricted circulation; stratigraphically they correspond
with assemblages of filter-feeding brachiopods and bryozoans, to the early TST (Lüning et al., 2000).
which compare closely with the postglacial depositional sedi- The postglacial shale unit receives different lithostrati-
ments described from low-energy species of Antarctic shelves graphic denominations (Aïn Deliouine Formation in Morocco,
and indicate, according to Angiolini et al. (2003), the final stage Argiles à Graptolites in Algeria, basal Tanezuft Shale in Lybia,
of the Gondwanan deglaciation. Qusaiba Member in Saudi Arabia, Sahmah Formation in Oman,
Batra Mudstone and Mudawwara in Jordan, Abba Formation in
TSTs Associated with the Late Ordovician Glaciation Syria, Akkas in Iraq, Dadas Formation in southeastern Turkey,
The proposed correspondence between lowstand incision and Ghakum Formation in Iran), but it seems to correspond to
of paleovalleys filled with glacial sediments during a time of ice a single, fast contemporaneous episode of postglacial inunda-
buildup and subsequent glacioeustatically induced transgression tion across much of the North African and Arabian Platforms.
immediately after glacier retreat is not exclusive to the LPIA. The Detailed studies of the associated graptolite faunas indicate that
Late Ordovician glaciation, present mainly in intracratonic basins the deposition of the organic-rich hot shales was a synchronous
of Mauritania, Mali, Morocco, Algeria, Libya, Tunisia, Jordan, event across North Africa through Arabia of Llandovery (Rhud-
and Saudi Arabia (Fig. 15), is particularly abundant in examples danian) age during the Early Silurian (Lüning et al., 2000; Miller
of these mechanisms, particularly for the postglacial transgression and Melvin, 2005).
and its well-known product in North African and Arabian Plat- In Arabia (1 in Fig. 15) the hot shales are highly fossilifer-
forms: the world-class source rock informally known as “the hot ous and pyritic. They correspond to the basal part of the Qusaiba
black shales” (Keeley and Massoud, 1998; Jones and Stump, 1999; Member (Qalibah Formation), the principal source rock for
Lüning et al., 2000; Carr, 2002). The hot shales are defined by an Paleozoic hydrocarbons in Saudi Arabia (Mahmoud et al., 1992),
arbitrary cutoff (>200 API units) in the gamma-ray curves of well and were deposited as a condensed sequence on a sediment-
logs (Lüning et al., 2000). This value correlates approximately starved shelf; the high TOC values have been interpreted as the
with TOC values of 3% for maturities around the oil window. Ini- result of high productivity in high-latitude water masses (Jones
tial marine transgression is marked by an erosional ravinement and Stump, 1999).
surface locally overlain by thin, residual shallow-marine sands Sequence stratigraphic studies (Lüning et al., 2000; Dardour
(Boote et al., 1998). Recent paleoglaciological reconstructions of et al., 2004) frame the record of this glacial event and its associ-
the Late Ordovician Saharan ice sheet (Le Heron and Craig, 2008) ated subsequent transgression in the context of an LST of glacial
suggest a stepwise, southward recession from the ice maximum, sediments covered by transgressive marine shales (TST). Dardour
followed by a postglacial transgression. The hot shales (basal et al. (2004) proposed a Late Ordovician–Silurian (second-order)
Tanezzuft Shale) in the North African Platform were deposited super-sequence comprising a periglacial lowstand, the Tanezzuft
during the initial transgression in paleodepressions (formed by transgressive to early highstand shales, and the Acacus Formation

Basins with glacial deposits


0 km 500
N Mediterranean Sea Outcrops of glacial deposits
Morocco Tunisia
Shield areas
Atlantic 6
4 Jordan
ra

Ghadames
ha

Ocean Basin 5 Iraq


Sa
rn

3
te

2
es
W

Mauritania Murzuq Lybia Egypt Saudi


Algeria Illizi Basin Arabia
Basin

Taoudeni 1 UAE
Basin Mali
Sudan
Re

Niger Oman
d

Chad
Se
a

Yemen

Figure 15. Map of North Africa and Middle East regions, with record of Late Ordovician glacial deposits and associated Early Silurian post-
glacial transgressive deposits. Based on compilation from several sources by Le Heron et al. (2009). Numbers refer to outcrop and subsurface
records of postglacial transgressions discussed in the text.
Transgressions related to the demise of the Late Paleozoic Ice Age 25

late HSTs. The lower boundary for this super-sequence is well Llandovery to Pridoli age that prograded basinward (Mahmoud
preserved in the northern part of the Murzuk Basin (SW Lybia, et al., 1992; Konert et al., 2001).
2 in Fig. 15) and in the contiguous Illizi Basin (SE Algeria, 3 A similar sequence is described for the glacial-postglacial
in Fig. 15) as deep, incised valleys. This topography was gradu- transition by Le Heron et al. (2007) in the Anti-Atlas of Morocco
ally infilled by a heterogeneous glacial (mostly glaciofluvial) (6 in Fig. 15) where stratified, clast-poor, sandy diamictites, lying
lowstand clastic facies that grades laterally into a more uniform, directly above a striated surface, pass vertically into transgressive
distal facies in the Ghadames Basin (eastern Algeria, southern tidal deposits (sigmoidally cross-bedded sandstones), which in
Tunisia, and NW Libya) farther north (4 in Fig. 15, Dardour turn pass upward into offshore mudstones (Aïn Deliouine For-
et al., 2004). Several higher frequency glacial sequences are rec- mation). Rare outsize boulders in clast-poor diamictites deform
ognized that may reflect several cycles of glacial advance and underlying laminae, indicative of dropstones from icebergs. The
retreat. The North African Platform was subsequently flooded by overall fining-upward section has been interpreted as having been
a transgressive facies, the Tanezzuft Shale, which contains at its deposited during ice sheet retreat in the earliest Silurian.
base a short-lived Rhuddanian anoxic event represented by thin
but regionally extensive organic-rich hot shales (Lüning et al., Asymmetry in Sea-Level Rates: Its Impact on the
2000), interpreted here as the MFI of the sequence. The aggre- Stratigraphic Record
gate thickness of the hot shales rarely exceeds 20 m in a thin
(~40 m), hot-shale–bearing basal section of the Tanezzuft Shale The asymmetry (fast rates of sea level rise, cf. Fig. 2) in
in the Ghadames and Illizi Basins (Lüning et al., 2000). The post- the eustatic sea-level curve defines a clear stratigraphic signature,
glacial TST was followed by regressive highstand sedimentation especially in cases where the sediment supply rate is relatively
commencing in the late Llandoverian; the HST continued to the moderate to low with respect to the sea-level rise. The extent
end of the Silurian with sedimentation of a northerly prograding of erosion and creation of a TRS depends on the rate of rise of
shelfal to fluviodeltaic wedge. relative sea level. In cases such as the postglacial TST with a
Similar scenarios have been described elsewhere for the relatively rapid sea-level rise, erosion is minimized. Thus TRSs
same Late Ordovician–earliest Silurian glacial-postglacial transi- are infrequent, and transgressive sediments are preserved. The
tion. Turner et al. (2005) described lowstand channel incisions postglacial signature is characterized by retrogradational stack-
and fill of those channels with glaciofluvial and shoreface sand- ing patterns. Two subtypes of TSTs, which should be considered
stones in southern Jordan (5 in Fig. 15) related to a fall of relative end members within a continuum for the submarine-retreat facies
sea level during ice buildup in southern Arabia. Each incision has association (Fig. 16), are identified:
been correlated with the first glacial advance during a stage of 1. Complete TSTs: In cases where the sediment supply rate
reduced accommodation space. This stage was followed by gla- is fast enough to keep up with the sea-level rise rate, the
cial melting and marine transgressive filling of the incised valleys lower part of the TST is dominated by rain-out processes,
as accommodation space increased. This process was repeated gravity flows (mostly proximal and distal debris flows),
four times with glacier re-advances evidenced by glacially and background sedimentation of fines by settling from
scoured, striated surfaces. The final transgressive filling (equiva- suspension. The resulting association, called here early
lent to the postglacial TST here) is characterized by the hot black- TST, is made up of three main facies stacked in a retro-
shale interval locally known as the lower Batra mudstone. The gradational pattern: (i) clast-poor, massive to poorly
base of the black shale is coincident with the MFS (Armstrong stratified diamictites; (ii) thinly bedded diamictites; and
et al., 2005). Recently Armstrong et al. (2009) concluded that the (iii) shales with IRD. The upper part of this type of TSTs
Batra Formation black shale was deposited in a short-lived, per- (complete TSTs) is made up of open-marine, IRD-free
manently stratified marine basin where an influx of fresh water shales, unusually associated with fine-grained sandstones
from melting ice and nutrients resulted in enhanced photic-zone with bi- or unidirectional ripples indicative of tenuous
primary productivity and organic matter sedimentation. They wave reworking in shallow-marine environments. Clas-
proposed the stratified basins and fjords of east Antarctica as pos- tic dilution owing to high sediment contributions would
sible modern analogues. prevent accumulation of significant volumes of organic
Similar valleys incised to depths exceeding 600 m (McClure, matter in the shales.
1978; Vaslet, 1990), and subsequently filled by glacial diamic- 2. Base-cut TSTs: Where the basal transgressive portion
tites and pro-glacial sandstones, have been traced into the sub- is mostly omitted, the early TST is represented, from
surface of northern Saudi Arabia with seismic data (McGillivray base to top, by distinctive distal debris-flow deposits (in
and Husseini, 1992). After the retreat of glaciers the Arabian Plat- some cases associated with low-density Bouma-type tur-
form was flooded by a rapid, glacioeustatic sea-level rise (Kon- bidites) and shales with IRD. The upper part of the TST
ert et al., 2001) that set anoxic conditions for the deposition and is represented by basinal (below wave base) IRD-free
preservation of organic-rich hot shales (Qusaiba shale). The post- shales. Alternatively, in extreme settings with a very high
glacial transgression (TST) was followed by a thick (>1000 m) sea-level rise rate, the entire TST is made up exclusively
coarsening-upward sequence of shales and sandstones (HST) of of IRD-free shales (the basal shale model of Wignall,
26 López-Gamundí

with Australian palynomorph zones, is present in sediments rest-


Sediment supply rate ing on glaciogenic diamictites (Pagoda Formation). This palyno-
flora indicates that the glaciation in the Transantarctic Mountains
was restricted to the Asselian–Sakmarian (Isbell et al., 2005).
Sea-level rise rate These deposits above the glaciogenic diamictites are grouped
under the Mackellar Formation of the Central Transantarctic
Mountains (Fig. 1). This unit is mostly made up of black, finely
Si S
G
Si S
G
Si
SG laminated shales and subordinate fine-grained sandstones with
ripples, resting on the glaciogenic deposits of the Pagoda Forma-
TST TST
tion (Lindsay, 1970; Miller, 1989); the upper Pagoda Formation
TST is characterized by massive shales with IRD (Miller et al., 1991).
LST C/S ratios in these sediments are extremely high, indicative of
LST fresh to slightly brackish waters (Miller et al., 1991); TOC con-
LST
tent is low (<1%, exceptionally ~2%) and is dominated by ter-
restrial plants (type III). This last example, in addition to those
documented herein in more detail, indicates that early postglacial
Complete TST Base-cut TST times were characterized by inland seas with dominantly brack-
ish waters rather than lakes.
Alternative views sustain the presence of huge lakes rather
Black shales (Fl) Thin-bedded diamictite (Dms) than inland seaways. This presumption is not based on detailed
Shales (Fl) and fine-grained
paleogeographic studies that demonstrate that closed bodies of
Thick-bedded diamictite (Dms)
sandstones (Sr) water (lakes) without connection to the open sea prevailed during
Dropstone shale (Fld) Diamictite (Dmm) with boulder beds
postglacial times but rather on indirect evidence of brackish to
fresh-water salinities provided by geochemical studies and, con-
Figure 16. Postglacial transgressive systems tract (TST) spectrum, sequently, to the extreme scarcity to absence of marine fossils. A
based on relative influence of sediment supply and sea-level-rise rates. good example of that debate is the Karoo Basin during the glacial-
See text for further discussion. Grain-size scale in logs: Si—silt; S— postglacial transition. Owing to the presence of marine fossils at
sand; G—gravel. LST—lowstand systems tract. the top of the Dwyka Formation (McLachlan and Anderson, 1973)
along the western and northern margins of the basin, there seems
to be agreement for the existence of a marine transgression related
1994). Wave reworking is unlikely in these fast trans- to eustatic sea-level rise following rapid global deglaciation.
gressions. These base-cut TSTs are common in many However, nonmarine conditions are suggested for the overlying
glacial-postglacial transitions and basically reflect the Prince Albert, Whitehill, and Collingham Formations by Milani
drastic sea-level rise related to ice melting. Owing to a and De Wit (2008), who further speculate on nonmarine condi-
relatively low sediment supply rate with respect to the tions for the Dwyka diamictites by stating that the “water-lain dia-
sea-level-rise rate, organic-rich shales could be deposited mictites, previously modeled as deposited in a wide marine-shelf
and preserved. environment (Visser 1989, 1997), may represent glaciogenic lake
sediments deposited beneath and peripheral to the major conti-
Water Salinity during Deglaciation nental ice sheet that covered much of Gondwana at that time”
(p. 333). In this interpretation the postglacial marine mudstones
Widespread evidence indicates that fresh-water contribu- were part of “a short marine transgression related to eustatic sea-
tions into the inland seas developed in early postglacial times. level rise following rapid global de-glaciation” (p. 333). Fluctu-
This fresh-water incursion, as melting icebergs, is interpreted to ating salinity (from brackish to fresh-water conditions) actually
have had the combined effect of creating a layer of brackish water prevailed during early postglacial times, as evidenced mostly
and a relatively high suspended-sediment load, as described for by geochemical data, and similar fluctuating salinity conditions
fjords in Greenland by Syvitski et al. (1987, 1996). probably persisted throughout the sedimentation of the Prince
Three independent lines of evidence suggest drastic fluctua- Albert, Whitehill, and Collingham Formations. The geochemical
tions of salinity from normal marine through brackish to fresh data supporting events of algal blooms in brackish to fresh water
waters for these postglacial seas: (1) ichnofacies associations, for the Whitehill Formation (Faure and Cole, 1999) do not need
(2) associated fauna, and (3) geochemical characteristics. to be necessarily related to a landlocked inland body of standing
An inland sea paleogeographic model has been proposed water, the definition of a lake (cf. Bates and Jackson, 1984), but
for the Transantarctic Basin (Barrett et al., 1986; Collinson and rather to an inland sea whose salinity may have been still con-
Miller, 1991; Miller and Collinson, 1994; Collinson et al., 1994). trolled by fresh-water incursions related to retreating valley gla-
Early Permian palynoflora (fossil spores and pollen), recorrelated ciers along the northen margin of the Karoo Basin. The retreat of
Transgressions related to the demise of the Late Paleozoic Ice Age 27

marine conditions in the Karoo Basin was closely related to the then deposited in the paleovalleys, which became isolated basins
emerging Cape fold belt along its southern margin by Middle to where conditions for the generation of low-oxygen conditions
Late Permian time, when deep-water deposits were replaced by were more favorable.
deltaic sediments, followed by fluvial and lacustrine sediments Additionally, some paleolatitudinal constraints seem to
(Wickens, 1992; Veevers et al., 1994). have been common for the presence of tasmanite beds. As
Cataclysmic floods (jökulhlaups) related to the Pleistocene pointed out by Revill et al. (1994), sediments where Tasmanites
deglaciation also have been invoked to explain the widespread is abundant (tasmanite beds), such as in the Late Ordovician–
presence of areally extended fine-grained sediments deposited in Early Silurian of North Africa, the Devonian of Brazil, and the
fresh waters. After an ice sheet retreated in a terrestrial setting, Late Carboniferous–Early Permian tasmanites of Tasmania, seem
meltwater was ponded in proglacial lakes developed behind ter- to have had similar high (~70°–75°) paleolatitudes, inferred from
minal moraines. As lake level rose with deglaciation, breaching paleomagnetic work (Smith et al., 1981; Bachtadse and Briden,
of the moraines occurred, flooding vast areas. This process has 1990; Li and Powell, 2001) and continental reconstructions based
been identified in several areas that were affected by the Pleis- on climatically sensitive lithofacies (Scotese and Barrett, 1990).
tocene glaciation (for a review, see Baker, 1997, and references
therein). However, this model, when applied to many of the Evolution of Basin Margins after Postglacial
glacial-postglacial sections in the basins affected by the LPIA, Transgressions
fails to explain the scarcity of facies that could be assigned to
marginal lacustrine settings. On the contrary, open-marine fos- In general, the postglacial TSTs across Gondwana are
siliferous shales present in the postglacial TSTs analyzed in this succeeded by progradational deposits (HSTs) as the rate of
contribution seem to be interbedded with the allegedly lacustrine, sea-level rise and accommodation space decreased (Fig. 2).
deep-water (below wave base) deposits without any intervening Facies variability is the norm within this common prograda-
shallower water facies (i.e., Malanzán Formation in the Paganzo tional stacking pattern, ranging from forced regressions (sensu
Basin, Lower Barakar Formation in the Satpura Basin). Posamentier et al., 1992) dominated by the drastic intrusion
of coarse clastics (Paganzo Basin, western Argentina, Lima-
Postglacial Transgressions and Source Rocks rino et al., 2010) to gradual filling of the basin by prodelta to
delta-front fine- to medium-grained sandstones and mudstones
Early postglacial sea-level rise favored creation of accom- (Calingasta-Uspallata Basin, western Argentina), in some cases
modation space with preservation potential for productivity- with significant tidal influence and associated delta plain coals
anoxia events and peat-forming conditions by rapid water-table (Barakar Formation in India). This wide gamut of facies in the
rise. Total organic content is a function of the rate of diluting sed- overlying progradational association contrasts sharply with the
imentation, the rate of degradation, and the onset of low-oxygen, narrower spectrum of facies in the underlying postglacial TST,
anoxic conditions (Wignall, 1994). These conditions, particularly which seems to be the product of a relatively simpler interaction
low sediment input and anoxic bottom conditions, are produced between the rate of sediment supply and the rate of sea-level rise
by the increasing water depth during a transgression. Therefore, associated with glacier retreat.
the postglacial TST includes the condensed section that represents In exceptional cases, such as along the southern margin of the
conditions of offshore sediment starvation and consequently opti- Karoo Basin, where tectonic subsidence rates outstripped sedi-
mal conditions for the accumulation and preservation of types I ment supply rates along a foredeep next to an emerging mountain
and II organic matter (Loutit et al., 1988; Schlutter, 1998). Fine- range (Cape fold belt), the deepening of the basin continued, and
grained facies in a TST contain high values of TOC and high the postglacial TST was overlain by basinal organic-rich shales
HI (hydrogen index) values and may be more prone to anoxia (Whitehill Formation) and deep-water, low- and high-density
(Wignall, 1991; Pasley et al., 1991, 1993). Conditions of clastic turbidites (Collingham, Laingsburg, and Ripon Formations)
starvation are more likely during rapid transgressions triggered (Catuneanu et al., 2002). In this extreme context the rate of sedi-
by high rates of relative sea-level rise, such as in the postglacial ment supply was unable to keep up with the rate of new accom-
transgressions analyzed herein. The shales in most postglacial modation space created by a very high subsidence rate, creating
TSTs seem to fit the model of condensed-section shales, depos- conditions of basin underfilling (forced transgression of Chough
ited during a period of the most rapid rise in sea level when clas- and Hwang, 1997) after the effects of glacioeustatic sea-level rise
tic contributions are trapped updip. A similar context has been ceased (cf. Figs. 2C and 2D).
proposed for the earliest Silurian (Rhuddanian) postglacial hot
shales of North Africa (Lüning et al., 2000). The uneven lateral CONCLUDING REMARKS
extent of these deglacial, transgressive black shales has been
linked to their deposition within a glacially sculpted topogra- The examples of glacial-postglacial transitions reviewed
phy. During the early stages of this transgression, triggered by herein cover a wide range of tectonic settings and basin types,
a rapid ice sheet recession, the most deeply scoured areas of the from backarc foreland basins to rifts. In all of them a clear ret-
postglacial shelf were the first to be flooded. Black shales were rogradational stacking pattern is detectable in the transition from
28 López-Gamundí

glacially dominated settings to glacially influenced early post- los de Bariloche, Argentina, Annual Meeting Project IGCP-211, Abstracts,
glacial environments. Drastic landward facies shifts are indica- p. 28–29.
Andreis, R.R., and Japas, S., 1996, Cuencas Sauce Grande y Colorado, in Arch-
tive of considerable relative sea-level rises, whose main driving angelsky, S., ed., El Sistema Pérmico en la República Argentina y en la
mechanism in these cases is glacioeustasy. Although tectonically República Oriental del Uruguay: Córdoba, Argentina, Academia Nacional
induced subsidence has to be considered an additional element in de Ciencias, p. 45–64.
Andreis, R.R., and Torres Ribeiro, M., 2003, Estratigrafía, facies y evolu-
the relative sea-level rise inferred from these sections, these post- ción depositacional de la Formación Sauce Grande (Carbonífero Supe-
glacial transgressions cannot be explained primarily by the sole rior), Cuenca Sauce Grande, Sierras Australes, Buenos Aires, Argentina:
effect of thermal sag-like or tectonic subsidence and consequent Revista de la Asociación Geológica Argentina, v. 58, p. 137–165.
Andreis, R.R., Lluch, J.J., and Iñíguez Rodríguez, A.M., 1979, Paleocorrientes
flooding. The resulting stratigraphic record can be framed in y paleoambientes de las formaciones Bonete y Tunas, Sierras Australes,
sequence stratigraphic terms as postglacial transgressive systems provincia de Buenos Aires, Argentina: Congreso Geológico Argentino
tracts (TSTs) principally driven by glacioeustasy. The balance (Bahía Blanca), 6th, Actas 2, p. 207–224.
Andreis, R.R., Leguizamón, R.R., and Archangelsky, S., 1986, El paleovalle
between sediment supply and sea level is critical to the deposi- de Malanzán: Nuevos criterios para la estratigrafía del Neopaleozoico de
tion of organic-rich black shales in the postglacial TSTs. A rela- la sierra de Los Llanos, República Argentina: Córdoba, Argentina, Aca-
tively high sediment supply rate can produce significant clastic demia Nacional de Ciencias, Boletin (Instituto de Estudios de Poblacion
y Desarrollo [Dominican Republic]), v. 57, p. 3–119.
dilution and prevent starved conditions in the newly inundated Andreis, R.R., Amos, A.J., Archangelsky, S., and González, C.G., 1987, Cuen-
postglacial shelf; on the other hand, this negative influence could cas Sauce Grande (Sierras Australes)—Colorado, in Archangelsky, S.,
be counterbalanced by the rapid eustatic sea-level rise produced ed., El Sistema Carbonífero en la República Argentina: Córdoba, Argen-
tina, Academia Nacional de Ciencias, p. 213–223.
during deglaciation. The two types of TSTs defined in this contri- Andrews, J.T., 1997, Northern Hemisphere (Laurentide) deglaciation: Pro-
bution (complete TST and base-cut TST) can be considered end cesses and responses office sheet/ocean interactions, in Martini, I.P.,
members, reflecting these two possible scenarios. ed., Late Glacial and Postglacial Environmental Changes, Quaternary,
Carboniferous–Permian and Proterozoic: Oxford, UK, Oxford University
Press, p. 9–27.
ACKNOWLEDGMENTS Angiolini, L., Balini, M., Garzanti, E., Nicora, A., and Tintori, A., 2003, Gond-
wanan deglaciation and opening of Neotethys: The Al Khlata and Sai-
wan Formations of Interior Oman: Palaeogeography, Palaeoclimatology,
This work benefited greatly from helpful discussions through Palaeoecology, v. 196, p. 99–123.
the years with many colleagues, particularly Arturo Amos, Jim Archangelsky, S., 1996, Palinoestratigrafía de la Plataforma Continental, in
Collinson, John Crowell, and Johan Visser. The author thanks Ramos, V.A., and Turic, M.A., eds., Geología y Recursos Naturales de
la Plataforma Continental Argentina, Relatorio: Buenos Aires, Congreso
Antonio Rocha-Campos and John Veevers for their insightful Geológico Argentino, 13th, and Congreso de Exploración de Hidrocarbu-
reviews of the manuscript. ros, 3rd, p. 67–72.
Armstrong, H.A., Turner, B.R., Makhlouf, I.M., Weedon, G.P., Williams, A.,
Smadie, A., and Salahe, A.A., 2005, Origin, sequence stratigraphy and
REFERENCES CITED depositional environment of an upper Ordovician (Hirnantian) deglacial
Adie, R.J., 1952, The position of the Falkland Islands in a reconstruction of black shale, Jordan: Palaeogeography, Palaeoclimatology, Palaeoecology,
Gondwanaland: Geological Magazine, v. 89, p. 401–410, doi: 10.1017/ v. 220, p. 273–289.
S0016756800068102. Armstrong, H.A., Abbott, G.D., Turner, B.R., Makhlouf, I.M., Muhammad,
Ahmad, N., 1975, Son Valley Talchir glacial deposits, Madhya Pradesh, India: A.B., Pedentchouk, N., and Peters, H., 2009, Black shale deposition in
Journal of the Geological Society of India, v. 16, p. 475–484. an Upper Ordovician–Silurian permanently stratified, peri-glacial basin,
Aitken, J.F., Clark, N.D., Osterloff, P.L., Penney, R.A., and Mohiuddin, U., southern Jordan: Palaeogeography, Palaeoclimatology, Palaeoecology,
2004, Regional core-based sedimentological review of the glacially- v. 273, p. 368–377, doi: 10.1016/j.palaeo.2008.05.005.
influenced Permo-Carboniferous Al Khlata Formation, South Oman Salt Bachtadse, V., and Briden, J.C., 1990, Paleomagnetic constraints on the posi-
basin, Oman: GeoArabia, v. 9, p. 16. tion of Gondwana during Ordovician to Devonian times, in McKerrow,
Aksu, A.E., 1984, Subaqueous debris flow deposits in Baffin Bay: Geo-Marine W.S., and Scotese, C.R., eds., Palaeozoic Biogeography and Palaeogeog-
Letters, v. 4, p. 83–90, doi: 10.1007/BF02277077. raphy: Geological Society [London] Memoir 12, p. 43–48.
Al-Belushi, J., Glennie, K.W., and Williams, B.P.J., 1996, Permo-Carboniferous Baker, V.R., 1997, Megafloods and glaciation, in Martini, I.P., ed., Late Gla-
glaciogenic Al Khlata Formation, Oman: A new hypothesis for origin of cial and Postglacial Environmental Changes, Quaternary, Carboniferous–
its glaciation: GeoArabia, v. 1, p. 389–403. Permian and Proterozoic: Oxford, UK, Oxford University Press,
Amos, A.J., 1980, La fauna de invertebrados en la cronología del Carbónico p. 98–108.
y Pérmico de Argentina: Buenos Aires, Congreso Argentino de Paleon- Banerjee, I., 1966, Turbidites in a glacial sequence: A study of the Talchir for-
tología y Bioestratigrafía, 2nd, and Congreso Latinoamericano de Paleon- mation, Raniganj coalfield, India: Journal of Geology, v. 74, p. 593–606,
tología, 1st, Actas 4, p. 231–234. doi: 10.1086/627191.
Amos, A.J., and López-Gamundí, O.R., 1981a, Late Paleozoic tillites and dia- Bangert, B., Stollhofen, H., Lorenz, V., and Armstrong, R.L., 1999, The
mictites of the Calingasta-Uspallata and Paganzo basins, in Hambrey, M., geochronology and significance of ash-fall tuffs in the glaciogenic,
and Harland, W., eds., Earth’s Pre-Pleistocene Glacial Record: Cam- Carboniferous-Permian Dwyka Group of Namibia and South Africa:
bridge, UK, Cambridge University Press, p. 859–868. Journal of African Earth Sciences, v. 29, p. 33–49, doi: 10.1016/S0899
Amos, A.J., and López-Gamundí, O.R., 1981b, Late Paleozoic Sauce Grande -5362(99)00078-0.
Formation of eastern Argentina, in Hambrey, M., and Harland, W., eds., Banks, M.R., 1981, Late Paleozoic tillites of Tasmania, in Hambrey, M., and
Earth’s Pre-Pleistocene Glacial Record: Cambridge, UK, Cambridge Uni- Harland, W., eds., Earth’s Pre-Pleistocene Glacial Record: Cambridge,
versity Press, p. 872–877. UK, Cambridge University Press, p. 495–501.
Amos, A.J., and Rolleri, E., 1965, El Carbónico marino en el valle Calingasta- Barrett, P.J., Elliot, D.H., and Lindsay, J.F., 1986, The Beacon Supergroup
Uspallata, San Juan-Mendoza: Boletín de Informaciones Petroleras, (Devonian-Triassic) in the Beardmore Glacier area, Antarctica, in Turner,
v. 368, p. 50–71. M.D., and Splettstoesser, J.F., eds., Geology of the Central Transantarctic
Andreis, R.R., 1984, Análisis litofacial de la Formación Sauce Grande (Car- Mountains: American Geophysical Union, Antarctic Research Series,
bónico superior?) Sierras Australes, provincia de Buenos Aires: San Car- v. 36, p. 339–429.
Transgressions related to the demise of the Late Paleozoic Ice Age 29

Barrie, J.V., and Conway, K.W., 2002, Contrasting glacial sedimentation pro- Brown, L.F., and Fisher, W.L., 1977, Seismic stratigraphic interpretation of
cesses and sea-level changes in two adjacent basins on the Pacific margin of depositional systems: Examples from Brazil rift and pull-apart basins, in
Canada, in Dowdeswell, J.A., and O’Cofaioh, C., eds., Glacier-Influenced Payton, C.E., ed., Seismic Stratigraphy—Applications to Hydrocarbon
Sedimentation on High-Latitude Continental Margins: Geological Soci- Exploration: American Association of Petroleum Geologists Memoir 26,
ety [London] Special Publication 203, p. 181–194. p. 213–248.
Bartek, L.R., and Anderson, J.B., 1991, Facies distribution resulting from sedi- Buatois, L.A., and Mangano, M.G., 2003, Caracterización icnológica y
mentation under polar interglacial climatic conditions within a high-latitude paleoambiental de la localidad tipo de Orchesteropus atavus, Huerta de
marginal basin, McMurdo Sound, Antarctica, in Anderson, J.B., and Ash- Huachi, provincia de San Juan, Argentina: Ameghiniana, v. 40, p. 53–70.
ley, G.M., eds., Glacial Marine Sedimentation: Paleoclimatic Significance: Buatois, L.A., Netto, R., Mangano, M.G., and Balistieri, P.L., 2006, Extreme
Geological Society of America Special Paper 261, p. 27–49. freshwater release during the late Paleozoic Gondwana deglaciation and its
Bates, R.L., and Jackson, J.L., 1984, Dictionary of Geological Terms: Alexan- impact on coastal ecosystems: Geology, v. 34, p. 1021–1024, doi: 10.1130/
dria, Virginia, American Geological Institute, 571 p. G22994A.1.
Bellosi, E., and Jalfin, G.A., 1987, Area Islas Malvinas, in Archangelsky, S., Buatois, L.A., Netto, R.G., and Mángano, M.G., 2010, this volume, Ichnol-
ed., El Sistema Carbonifero en la Republica Argentina: Córdoba, Argen- ogy of late Paleozoic post-glacial transgressive deposits in Gondwana:
tina, Academia Nacional de Ciencias, p. 226–237. Reconstructing salinity conditions in coastal ecosystems affected by
Berner, R.A., and Raiswell, R., 1984, C/S method for distinguishing freshwater strong meltwater discharge, in López-Gamundí, O.R., and Buatois,
from marine rocks: Geology, v. 12, p. 365–368, doi: 10.1130/0091-7613 L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions
(1984)12<365:CMFDFF>2.0.CO;2. in Gondwana: Geological Society of America Special Paper 468, doi:
Beuf, S., Biju-Duval, B., de Charpal, O., Rognon, P., Gariel, O., and Benna- 10.1130/2010.2468(07).
cef, A., 1971, Les Grès du Palaeozoique Inferiéur au Sahara: Paris, Edi- Cairncross, B., 1980, Anastomosing river deposits: Palaeoenvironmental con-
tions Technip, 464 p. trol on coal quality and distribution, northern Karoo Basin: Transactions—
Bhattacharya, B., and Bhattacharya, N.H., 2007, Implications of trace fossil Geological Society of South Africa, v. 83, p. 327–332.
assemblages from Late Paleozoic glaciomarine Talchir Formation, Rani- Cairncross, B., 1989, Paleodepositional environments and tectonosedimentary
ganj Basin, India: Gondwana Research, v. 12, p. 509–524, doi: 10.1016/ controls of the postglacial Permian coals, Karoo Basin, South Africa:
j.gr.2006.12.002. International Journal of Coal Geology, v. 12, p. 365–380, doi: 10.1016/
Bhattacharya, H.N., Chakraborty, A., and Bhattacharya, B., 2005, Significance 0166-5162(89)90058-X.
of transition between Talchir Formation and Karharbari Formation in Cairncross, B., 2001, An overview of the Permian (Karoo) coal deposits of
lower Gondwana Basin evolution—A study in West Bokaro Coal Basin, southern Africa: Journal of African Earth Sciences, v. 33, p. 529–562, doi:
Jharkhand, India: Journal of Earth System Science, v. 114, p. 275–286, 10.1016/S0899-5362(01)00088-4.
doi: 10.1007/BF02702950. Canuto, J.R., dos Santos, P.R., and Rocha-Campos, A.C., 2001, Estratigrafia
Bhattacharya, S.K., Ghosh, P., and Chakrabarti, A., 2002, Isotopic analysis of de seqüências do Subgrupo Itararé (Neopaleozóico) no leste da Bacia
do Paraná, nas regiões sul do Paraná e norte de Santa Catarina, Brasil:
Permo-Carboniferous Talchir sediments from East-Central India: Signa-
Revista Brasileira de Geociencias, v. 31, p. 107–116.
ture of glacial melt-water lakes: Chemical Geology, v. 188, p. 261–274,
Caputo, M.V., 1985, Late Devonian glaciation in South America: Palaeogeog-
doi: 10.1016/S0009-2541(02)00140-7.
raphy, Palaeoclimatology, Palaeoecology, v. 51, p. 291–317, doi: 10.1016/
Bischoff, J.L., Fitzpatrick, J.A., and Rosenbauer, R.J., 1993, The solubility and
0031-0182(85)90090-2.
stabilization of ikaite (CaCO3·6H2O) from 0 degrees to 25 degrees C;
Caputo, M.V., Melo, J.H.G., Streel, M., and Isbell, J.L., 2008, Late Devonian
environmental and paleoclimatic implications for thinolite tufa: Journal
and Early Carboniferous glacial records of South America, in Fielding,
of Geology, v. 101, p. 21–33, doi: 10.1086/648194.
C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice
Bjorlykke, K., 1985, Glaciations, preservation of their sedimentary record and Age in Time and Space: Geological Society of America Special Paper
sea level changes: A discussion based on the Late Precambrian and Lower 441, p. 161–173.
Paleozoic sequence in Norway: Palaeogeography, Palaeoclimatology, Carlotto, V., Díaz Martinez, E., Cerpa, L., Arispe, O., and Cardenas, J., 2004,
Palaeoecology, v. 51, p. 197–207, doi: 10.1016/0031-0182(85)90085-9. Late Devonian glaciation in the northern Central Andes: New evidence
Bohacs, K., and Sutter, J., 1997, Sequence stratigraphic distribution of coaly from southeast Peru: Florence, International Geological Congress, 32nd,
rocks: Fundamental controls and paralic examples: American Association Abstracts Volume, pt. 2, abstract 205-12, p. 96.
of Petroleum Geologists Bulletin, v. 81, p. 1612–1639. Carr, I.D., 2002, Second order sequence stratigraphy of the Palaeozoic of North
Boote, D.R.D., Clark-Lowes, D.D., and Traut, M.W., 1998, Palaeozoic petro- Africa: Journal of Petroleum Geology, v. 25, p. 259–280, doi: 10.1111/
leum systems of North Africa, in MacGregor, D.S., Moody, R.T.J., and j.1747-5457.2002.tb00009.x.
Clark-Lowes, D.D., eds., Petroleum Geology of North Africa: Geological Casagrande, D.I., Sieffert, K., Berschinski, C., and Sutton, N., 1977, Sulfur in
Society [London] Special Publication 132, p. 7–68. peat forming systems of the Okefenokee Swamp and Florida Everglades:
Bordy, E.M., and Catuneanu, O., 2002, Sedimentology of the lower Karoo Super- Origins of sulfur in coal: Geochimica et Cosmochimica Acta, v. 41,
group fluvial strata in the Tuli Basin, South Africa: Journal of African Earth p. 161–167, doi: 10.1016/0016-7037(77)90196-X.
Sciences, v. 35, p. 503–521, doi: 10.1016/S0899-5362(02)00129-X. Casshyap, S.M., and Qidwai, H.A., 1974, Glacial sedimentation of late Palae-
Bose, P.K., Mukhopadhyay, G., and Bhattacharya, H.N., 1992, Glaciogenic ozoic Talchir diamict, Pench valley coalfield, central India: Geological
coarse clastics in a Permo-Carboniferous bedrock trough in India: A sedi- Society of America Bulletin, v. 85, p. 749–760, doi: 10.1130/0016-7606
mentary model: Sedimentary Geology, v. 76, p. 79–97, doi: 10.1016/0037 (1974)85<749:GSOLPT>2.0.CO;2.
-0738(92)90140-M. Casshyap, S.M., and Srivastava, V.K., 1987, Glacial and proglacial Talchir sedi-
Braakman, J.H., Martin, J.H., Potter, T.L., and van Vliet, A., 1982, Late mentation in Son-Mahanadi Gondwana basin: Palaeogeographic recon-
Paleozoic Gondwana glaciation in Oman: Nature, v. 299, p. 48–50, doi: struction, in McKenzie, G.D., ed., Gondwana Six: American Geophyscal
10.1038/299048a0. Union Geophysical Monograph 41, p. 167–182.
Brakel, A.T., and Totterdell, J.M., 1993,The Sakmarian-Kungurian of Australia, Cattaneo, A., and Steel, R., 2003, Transgressive deposits: A review of their vari-
in Findlay, R.H., Unrug, R., Banks, M.R., and Veevers, J.J., eds., Gond- ability: Earth-Science Reviews, v. 62, p. 187–228, doi: 10.1016/S0012-8252
wana Eight, Assembly Evolution and Dispersal: Brookfield, Vermont, (02)00134-4.
A.A. Balkema, p. 385–396. Catuneanu, O., 2004, Basement control on flexural profiles and the distribution
Brenchley, P.J., Marshall, J.D., Carden, G.A.F., Robertson, D.B.R., Long, of foreland facies: The Dwyka Group of the Karoo Basin, South Africa:
D.G.F., Meidla, T., Hints, L., and Anderson, T.F., 1994, Bathymetric and Geology, v. 32, p. 517–520, doi: 10.1130/G20526.1.
isotopic evidence for a short-lived Late Ordovician glaciation in a green- Catuneanu, O., Hancox, P.J., Cairncross, B., and Rubidge, B.S., 2002, Fore-
house period: Geology, v. 22, p. 295–298, doi: 10.1130/0091-7613(1994 deep submarine fans and forebulge deltas: Orogenic off-loading in the
)022<0295:BAIEFA>2.3.CO;2. underfilled Karoo Basin: Journal of South African Earth Sciences, v. 35,
Brezinski, D.K., Cecil, C.B., Skemac, V.W., and Stamm, R., 2008, Late Devo- p. 489–502, doi: 10.1016/S0899-5362(02)00154-9.
nian glacial deposits from the eastern United States signal an end of the Cecil, C.B., Brezinski, D.K., and Dulong, F., 2004, The Paleozoic Record of
mid-Paleozoic warm period: Palaeogeography, Palaeoclimatology, Palae- Changes in Global Climate and Sea Level: Central Appalachian Basin:
oecology, v. 268, p. 143–151, doi: 10.1016/j.palaeo.2008.03.042. U.S. Geological Survey Circular 1264, 34 p.
30 López-Gamundí

Cerpa, L., Carlotto, V., Arispe, O., Díaz Martínez, E., Cárdenas, J., Valder- Petroleum Geology, v. 27, p. 141–162, doi: 10.1111/j.1747-5457.2004
rama, P., and Bermúdez, O., 2004, Formación Ccatca (Devónico supe- .tb00050.x.
rior): Sedimentación glaciomarina en la Cordillera Oriental de la región DeLurio, J.L., and Frakes, L.A., 1999, Glendonites as a paleoenvironmental
de Cusco: Congreso Peruano de Geología, 12th, Resúmenes Extendidos, tool: Implications for Early Cretaceous high latitude climates in Australia:
Sociedad Geológica del Perú, Publicación Especial, v. 6, p. 424–427. Geochimica et Cosmochimica Acta, v. 63, p. 1039–1048, doi: 10.1016/
Cesari, S.N., and Bercowski, F., 1997, Palinología de la Formación Jejenes S0016-7037(99)00019-8.
(Carbonífero) en la quebrada de Las Lajas, provincia de San Juan. Nuevas Desjardins, P.R., Buatois, L.A., Mángano, M.G., and Limarino, C.O., 2010, this
inferencias paleoambientales: Ameghiniana, v. 34, p. 497–509. volume, Ichnology of the latest Carboniferous–earliest Permian transgres-
Chakraborty, C., and Ghosh, S.K., 2005, Pull-apart origin of the Satpura sion in the Paganzo Basin of western Argentina: The interplay of ecology,
Gondwana basin, central India: Journal of Earth System Science, v. 114, sea-level rise, and paleogeography during postglacial times in Gondwana,
p. 259–273, doi: 10.1007/BF02702949. in López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial
Chakraborty, C., Mandalb, N., and Ghosh, S.K., 2003, Kinematics of the Gondwana Events and Postglacial Transgressions in Gondwana: Geological Society
basins of peninsular India: Tectonophysics, v. 377, p. 299–324, doi: 10.1016/ of America Special Paper 468, doi: 10.1130/2010.2468(08).
j.tecto.2003.09.011. De Wit, M.J., Jeffery, M., Bergh, H., and Nicolaysen, L., 1988, Geological Map
Chough, S.K., and Hwang, I.G., 1997, The Duksung fandelta, SE Korea: of Sectors of Gondwana Reconstructed to Their Disposition—150 Ma:
Growth of delta lobes on a Gilbert-type topset in response to relative sea- American Association of Petroleum Geologists and University of Witwa-
level rise: Journal of Sedimentary Research, v. 67, p. 725–739. tersrand, scale: 10,000,000, 1 sheet.
Cisterna, G.A., and Sterren, A.F., 2010, this volume, “Levipustula Fauna” in Deynoux, M., Sougy, J., and Trompette, R., 1985, Lower Paleozoic rocks of
central-western Argentina and its relationships with the Carboniferous West Africa and the western part of central Africa, in Holland, C.H.,
glacial event in the southwestern Gondwanan margin, in López-Gamundí, ed., Lower Palaeozoic of Northwestern and West Central Africa, in
O.R., and Buatois, L.A., eds., Late Paleozoic Glacial Events and Postgla- Lower Palaeozoic Rocks of the World, vol. 4: New York, Wiley & Sons,
cial Transgressions in Gondwana: Geological Society of America Special p. 375–495.
Paper 468, doi: 10.1130/2010.2468(06). Díaz Martínez, E., and Isaacson, P.E., 1994, Late Devonian glacially-influenced
Clague, J.J., Harper, J.R., Hebda, R.J., and Howes, D.E., 1982, Late Quaternary marine sedimentation in western Gondwana: The Cumaná Formation,
sea levels and crustal movements, coastal British Columbia: Canadian Altiplano, Bolivia: Canadian Society of Petroleum Geologists Memoir 17,
Journal of Earth Sciences, v. 19, p. 597–618, doi: 10.1139/e82-048. p. 511–522.
Clarke, M.J., and Banks, M.R., 1975, The stratigraphy of the Lower (Permo- Dickins, J.M., 1961, Eurydesma and Peruvispira from the Dwyka beds of South
Carboniferous) parts of the Parmeener super-group, Tasmania, in Camp- Africa: Palaeontology, v. 4, p. 138–148.
bell, K.S.W., ed., Gondwana Geology: Canberra, Australian National Dickins, J.M., 1984, Late Paleozoic glaciation: Journal of Australian Geology
University Press, p. 453–467. and Geophysics, v. 9, p. 163–169.
Clarke, M.J., and Forsyth, S.M., 1989, Late Carboniferous–Triassic, in Burrett, Dickins, J.M., 1996, Problems of a late Paleozoic glaciation in Australia and
C.F., and Marten, E.L., eds., Geology and Mineral Resources of Tasma- subsequent climate in the Permian: Palaeogeography, Palaeoclimatol-
ogy, Palaeoecology, v. 125, p. 185–197, doi: 10.1016/S0031-0182(96
nia: Geological Association of Australia Special Paper 15, p. 209–293.
)00030-2.
Coates, D.A., 1969, Stratigraphy and sedimentation of the Sauce Grande Forma-
Dickins, J.M., and Shah, S.C., 1977, Correlation of the Permian marine
tion, Sierra de la Ventana, Southern Buenos Aires Province: International
sequences of India and western Australia: Calcutta, Hindusthan Publish-
Gondwana Symposium, 1st (Mar del Plata, 1967), Paris, 2, p. 799–816.
ing, Gondwana Symposium, 4th, v. 2, p. 387–408.
Cohen, A.D., Raymond, R., Ramirez, A., Morales, Z., and Ponce, F., 1989, The
Domack, E.W., Burkley, L.A., Domack, C.R., and Banks, M.R., 1993, Facies
Changuinola peat deposit of northwestern Panama: A tropical back·barrier,
analysis of glacial marine pebbly mudstones in the Tasmania Basin: Impli-
peat (coal)-forming environment: International Journal of Coal Geology,
cations for regional paleoclimates during the late Paleozoic, in Findlay,
v. 12, p. 157–192, doi: 10.1016/0166-5162(89)90050-5. R.H., Unrug, R., Banks, M.R., and Veevers, J.J., eds., Gondwana Eight,
Collinson, J.W., and Miller, M.F., 1991, Comparison of Lower Permian post- Assembly Evolution and Dispersal: Brookfield, Vermont, A.A. Balkema,
glacial black shale sequences in the Ellsworth and Transantarctic Moun- p. 471–484.
tains, Antarctica, in Ulbrich, H., and Rocha-Campos, A., eds., Gondwana Dos Santos, P.R., Rocha-Campos, A.C., and Canuto, J.R., 1996, Patterns of
Seven Proceedings: São Paulo, Instituto de Geociencias University of São late Palaeozoic deglaciation in the Paraná Basin, Brazil: Palaeogeography,
Paulo, p. 217–231. Palaeoclimatology, Palaeoecology, v. 125, p. 165–184.
Collinson, J.W., Isbell, J.L., Elliot, D.H., Miller, M.F., and Miller, J.M.G., Drewry, D.J., and Cooper, A.P.R., 1981, Processes and models of Antarctic gla-
1994, Permian-Triassic Transantarctic Basin, in Veevers, J.J., and Powell, ciomarine sedimentation: Annals of Glaciology, v. 2, p. 117–122.
C.McA., eds., Permian-Triassic Pangean Basins and Foldbelts along the Du Toit, A.L., 1927, A Geological Comparison of South America with South
Panthalassan Margin of Gondwanaland: Geological Society of America Africa: Washington, D.C., Carnegie Institution Publication 381, 157 p.
Memoir 184, p. 173–222. Edwards, M., 1986, Glacial environments, in Reading, H.G., ed., Sedimentary
Crowell, J.C., 1978, Gondwanan glaciation, cyclothems, continental positioning Environments and Facies: London, Blackwell Science, p. 445–469.
and climate change: American Journal of Science, v. 278, p. 1345–1372. Eiras, J.F., Becker, C.R., Souza, E.M., Gonzaga, F.G., Da Silva, F.G., Daniel,
Crowell, J.C., 1983, The recognition of ancient glaciation, in Medaris, L.G., J.G.F., Matsuda, N.S., and Feijo, F.J., 1994, Baçia de Solimões: Boletim
Jr., Byers, C.W., Mickelson, D.M., and Shanks, W.C., eds., Proterozoic de Geociencias da Petrobras, v. 1, p. 17–46.
Geology: Selected Papers from an International Proterozoic Symposium: Emery, D., and Myers, K.J., 1996, Sequence Stratigraphy: London, Blackwell
Geological Society of America Memoir 161, p. 289–297. Science, 297 p.
Crowell, J.C., 1999, Pre-Mesozoic Ice Ages: Their Bearing on Understanding Evans, J., and Pudsey, C.J., 2002, Sedimentation associated with Antarctic Pen-
the Climate System: Geological Society of America Memoir 192, 106 p. insula ice shelves: Implications for palaeoenvironmental reconstructions
Crowell, J.C., and Frakes, L.A., 1971, Late Paleozoic glaciation: Part IV, Aus- of glacimarine sediments: Journal of the Geological Society [London],
tralia: Geological Society of America Bulletin, v. 82, p. 2515–2540, doi: v. 159, p. 233–237, doi: 10.1144/0016-764901-125.
10.1130/0016-7606(1971)82[2515:LPGPIA]2.0.CO;2. Eyles, C.H., 1988, A model for striated boulder pavement formation on glaci-
Crowell, J.C., and Frakes, L.A., 1972, Late Paleozoic glaciation: Part V, ated, shallow-marine shelves: An example from the Yakataga Formation,
Karoo Basin, South Africa: Geological Society of America Bulletin, Alaska: Journal of Sedimentary Research, v. 58, p. 62–71.
v. 83, p. 2887–2912, doi: 10.1130/0016-7606(1972)83[2887:LPGPVK Eyles, C.H., Eyles, N., and Miall, A.D., 1985, Models of glaciomarine sedi-
]2.0.CO;2. mentation and their application to the interpretation of ancient glacial
Cunha, P.R.C., Gonzaga, F.G., Coutinho, L.F.C., and Feijo, F.J., 1994, Baçia do sequences: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 51,
Amazonas: Boletim de Geociencias da Petrobras, v. 1, p. 47–56. p. 15–84, doi: 10.1016/0031-0182(85)90080-X.
Curray, J.R., 1964, Transgressions and regressions, in Miller, R.L., ed., Papers Eyles, C.H., Eyles, N., and Gostin, V.A., 1998, Facies and allostratigraphy
in Marine Geology: New York, Macmillan, p. 175–203. of high-latitude, glacially influenced marine strata of the early Permian
Dardour, A.M., Boote, D.R.D., and Baird, A.W., 2004, Stratigraphic controls southern Sydney basin, Australia: Sedimentology, v. 45, p. 121–161, doi:
on Paleozoic petroleum systems, Ghadames basin, Lybia: Journal of 10.1046/j.1365-3091.1998.00138.x.
Transgressions related to the demise of the Late Paleozoic Ice Age 31

Eyles, N., 1993, Earth’s Glacial Records and Its Tectonic Setting: Earth-Science Ghosh, S.K., and Mitra, N.D., 1975, History of Talchir Sedimentation in Damo-
Reviews, v. 35, 248 p., doi: 10.1016/0012-8252(93)90002-O. dar Valley Basins: Memoirs of the Geological Survey of India, v. 105,
Eyles, N., 2008, Tectonic boundary conditions for glaciation: Glacio-epochs 117 p.
and the supercontinent cycle after ~3.0 Ga: Palaeogeography, Palaeocli- Goes, A.M.O., and Feijo, F.J., 1994, Baçia do Parnaiba: Boletim de Geocien-
matology, Palaeoecology, v. 258, p. 89–129, doi: 10.1016/j.palaeo.2007 cias da Petrobras, v. 1, p. 57–68.
.09.021. González, C.R., 1981, Pavimento glaciario en el Carbónico de la Precordillera:
Eyles, N., Eyles, C.H., and Miall, A.D., 1983, Lithofacies types and vertical Revista de la Asociación Geológica Argentina, v. 36, p. 262–266.
profile models: An alternative approach to the description and environ- González, C.R., 1985, Esquema bioestratigráfico del Paleozoico superior marino
mental interpretation of glacial diamict and diamictite sequences: Sedi- de la Cuenca Uspallata-Iglesia, República Argentina: Acta Geológica Lil-
mentology, v. 30, p. 393–410, doi: 10.1111/j.1365-3091.1983.tb00679.x. loana, v. 16, p. 231–272.
Faure, K., and Cole, D.I., 1999, Geochemical evidence for lacustrine microbial González, C.R., 1990, Development of the late Paleozoic glaciation of South
blooms in the vast Permian Main Karoo, Paraná, Falkland Islands and American Gondwana in western Argentina: Palaeogeography, Palaeo-
Huab basins of southwestern Gondwana: Palaeogeography, Palaeoclima- climatology, Palaeoecology, v. 79, p. 275–287, doi: 10.1016/0031-0182
tology, Palaeoecology, v. 152, p. 189–213, doi: 10.1016/S0031-0182(99 (90)90022-Y.
)00062-0. Goswami, S., 2008, Marine influence and incursion in the Gondwana basins of
Faure, K., Armstrong, R.A., Harris, C., and Willis, J.P., 1996, Provenance of Orissa, India: A review: Palaeoworld, v. 17, p. 21–32, doi: 10.1016/j.palwor
mudstones in the Karoo Supergroup of the Ellisras Basin, South Africa: .2007.08.001.
Geochemical evidence: Journal of African Earth Sciences, v. 23, p. 189– Gradstein, F.M., Ogg, J.G., and Smith, A.G., eds., 2004, A Geologic Time
204, doi: 10.1016/S0899-5362(96)00061-9. Scale: Cambridge, UK, Cambridge University Press, 610 p.
Fernández Garrasino, C., 1996, Cuenca Chacoparanaense, in Archangelsky, S., Gravenor, C.P., and Rocha-Campos, A.C., 1983, Patterns of Late Paleozoic
ed., El Sistema Pérmico en la República Argentina y en la República Ori- glacial sedimentation on the southeast side of the Paraná Basin, Brazil:
ental del Uruguay: Córdoba, Argentina, Academia Nacional de Ciencias, Palaeogeography, Palaeoclimatology, Palaeoecology, v. 43, p. 1–39, doi:
p. 27–38. 10.1016/0031-0182(83)90046-9.
Fielding, C.R., Frank, T.D., and Isbell, J.L., eds., 2008a, Resolving the Late Guit, F.A., Al-Lawati, M.H., and Nederlof, P.J.R., 1995, Seeking new potential
Paleozoic Ice Age in Time and Space: Geological Society of America in the Early–Late Permian Gharif play, west Central Oman, in Al-Husseini,
Special Paper 441, 354 p. M.I., ed., Middle East Petroleum Geosciences, GEO ’94: Bahrain, Gulf
Fielding, C.R., Frank, T.D., Birgenheier, L.P., Rygel, M.C., Jones, A.T., and Petrolink, p. 447–462.
Roberts, J., 2008b, Stratigraphic imprint of the Late Palaeozoic Ice Age Gutiérrez, P.R., and Limarino, C.O., 2001, Palinología de la Formación Malan-
in eastern Australia: A record of alternating glacial and nonglacial climate zán (Carbonífero Superior), La Rioja, Argentina: Nuevos datos y consid-
regime: Journal of the Geological Society [London], v. 165, p. 129–140, eraciones paleoambientales: Ameghiniana, v. 38, p. 99–118.
doi: 10.1144/0016-76492007-036. Haldorsen, S., Von Brunn, V., Maud, R., and Truter, E., 2001, A Weichselian
deglaciation model applied to the Early Permian glaciation in the north-
Frakes, L.A., and Crowell, J.C., 1967, Facies and paleogeography of late
east Karoo Basin, South Africa: Journal of Quaternary Science, v. 16,
Paleozoic diamictite, Falkland Islands: Geological Society of America
p. 583–593, doi: 10.1002/jqs.637.
Bulletin, v. 78, p. 37–58, doi: 10.1130/0016-7606(1967)78[37:FAPOLP
Hambrey, M.J., 1985, The Late Ordovician–Early Silurian glacial period:
]2.0.CO;2.
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 51, p. 273–289,
Frakes, L.A., and Crowell, J.C., 1969, Late Paleozoic glaciation: I: South
doi: 10.1016/0031-0182(85)90089-6.
America: Geological Society of America Bulletin, v. 80, p. 1007–1042,
Hambrey, M., and Harland, W., eds., 1981, Earth’s Pre-Pleistocene Glacial
doi: 10.1130/0016-7606(1969)80[1007:LPGISA]2.0.CO;2.
Record: Cambridge, UK, Cambridge University Press, 1044 p.
França, A.B., 1994, Itararé Group: Gondwanan Carboniferous-Permian of the
Hand, S.J., 1993, Palaeogeography of Tasmania’s Permo-Carboniferous gla-
Paraná Basin, Brazil, in Deynoux, M., Miller, J.M.G., Domack, E.W., cigenic sediments, in Findlay, R.H., Unrug, R., Banks, M.R., and Veevers,
Eyles, N., Fairchild, I., and Young, G.M., eds., Earth’s Glacial Record: J.J., eds., Gondwana Eight, Assembly Evolution and Dispersal: Brook-
Cambridge, UK, Cambridge University Press, p. 70–82. field, Vermont, A.A. Balkema, p. 459–469.
França, A.B., and Potter, E.D., 1991, Stratigraphy and reservoir potential of Harrington, H.J., 1947, Explicación de las Hojas Geológicas 33m y 34m, Sierra
glacial deposits of the Itararé Group (Carboniferous-Permian), Paraná de Curamalal y de la Ventana, Provincia de Buenos Aires, Servicio Nacio-
basin, Brazil: American Association of Petroleum Geologists Bulletin, nal Minero Geológico: Boletin Direccion Nacional de Geologia y Min-
v. 75, p. 62–85. eria, No. 61.
França, A.B., Winter, W.R., and Assine, M.L., 1996, Arenitos Lapa–Vila Velha: Harrington, H.J., 1955, The Permian Eurydesma fauna of eastern Argentina:
um modelo de trato de sistemas subaquosos canal-lobos sob influência Journal of Paleontology, v. 29, p. 112–128.
glacial, Grupo Itararé (C-P), bacia do Paraná: Revista Brasileira de Geo- Helal, A.H., 1964, On the occurrence and stratigraphic position of Permo-
ciências, v. 26, p. 43–56. Carboniferous tillites in Saudi Arabia: Geologische Rundschau, v. 54,
Fryklund, B., Marshall, A., and Stevens, J., 1996, Cuenca del Colorado, in p. 193–207, doi: 10.1007/BF01821178.
Ramos, V.A., and Turic, M.A., eds., Geología y Recursos Naturales de Helland-Hansen, W., and Gjelberg, J.G., 1994, Conceptual basis and variability
la Plataforma Continental Argentina, Relatorio: Bueno Aires, Congreso in sequence stratigraphy: A different perspective: Sedimentary Geology,
Geológico Argentino, 13th, and Congreso de Exploración de Hidrocarbu- v. 92, p. 31–52, doi: 10.1016/0037-0738(94)90053-1.
ros, 3rd, p. 135–158. Heller, P.L., Chris Paola, C., Hwang, I., John, B., and Steel, R., 2001, Geo-
Ghienne, J.-F., 2003, Late Ordovician sedimentary environments, glacial morphology and sequence stratigraphy due to slow and rapid base-level
cycles, and post-glacial transgression in the Taoudeni Basin, West Africa: changes in an experimental subsiding basin (XES 96-1): American Asso-
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 189, p. 117–145, ciation of Petroleum Geologists Bulletin, v. 85, p. 817–838.
doi: 10.1016/S0031-0182(02)00635-1. Herbert, C.T., and Compton, J.S., 2007, Depositional environments of the
Ghienne, J.-F., Le Heron, D.P., Moreau, J., and Deynoux, M., 2007, The Late lower Permian Dwyka diamictite and Prince Albert shale inferred from
Ordovician glacial sedimentary system of the West Gondwana platform, the geochemistry of early diagenetic concretions, southwest Karoo Basin,
in Hambrey, M.J., Christofferson, P., Glasser, N.F., and Hubbard, B., eds., South Africa: Sedimentary Geology, v. 194, p. 263–277, doi: 10.1016/
Glacial Sedimentary Processes and Products: International Association of j.sedgeo.2006.06.008.
Sedimentologists Special Publication 39, p. 295–319. Hill, P.R., Aksu, A.E., and Piper, D.J., 1982, The deposition of thin bedded
Ghosh, S.K., 1954, Discovery of a new locality of marine Gondwana formation: subaqueous debris flow deposits, in Saxov, S., and Nieuwenhuis, J., eds.,
Science and Culture, v. 19, p. 620. Marine Slides and Other Mass Movements: New York, Plenum Publish-
Ghosh, S.K., 2003, First record of marine bivalves from the Talchir Forma- ing, p. 263–287.
tion of the Satpura Gondwana Basin, India: Palaeobiogeographic impli- Hughes Clarke, M.W., 1988, Stratigraphy and rock unit nomenclature in the oil
cations: Gondwana Research, v. 6, p. 312–320, doi: 10.1016/S1342 producing area of interior Oman: Journal of Petroleum Geology, v. 11,
-937X(05)70980-1. p. 5–60, doi: 10.1111/j.1747-5457.1988.tb00800.x.
32 López-Gamundí

Isaacson, P.E., and Díaz Martínez, E., 2005, Late Devonian glaciation in west- Le Heron, D.P., and Dowdeswell, J.A., 2009, Calculating ice volumes and ice
ern Gondwana and Laurentia: A major event and its consequences: Men- flux to constrain the dimensions of a 440 Ma North African ice sheet: Jour-
doza, Argentina, Gondwana 12, Abstracts, p. 207. nal of the Geological Society [London], v. 166, p. 277–281, doi: 10.1144/
Isaacson, P.E., Hladil, J., Shen, J.W., Kalvoda, J., and Grader, G., 1999, Late 0016-76492008-087.
Devonian (Fammenian) glaciation in South America and marine offlap Le Heron, D.P., Ghienne, J.-F., El Houicha, M., Khoukhi, Y., and Rubino, J.,
on other continents, in Feist, R., Talent, J.A., and Daurer, A., eds., North 2007, Maximum extent of ice sheets in Morocco during the Late Ordo-
Gondwana: Mid-Paleozoic Terranes, Stratigraphy and Biota: Vienna, vician glaciation: Palaeogeography, Palaeoclimatology, Palaeoecology,
Jahrbuch der Geologischen Bundesantstalt v. 54, p. 239–257. v. 245, p. 200–226, doi: 10.1016/j.palaeo.2006.02.031.
Isaacson, P.E., Díaz-Martínez, E., Grader, G.W., Kalvod, J., Babe, O., and Le Heron, D.P., Craig, J., and Etienne, J.L., 2009, Ancient glaciations and
Devuyst, F.X., 2008, Late Devonian–earliest Mississippian glaciation in hydrocarbon accumulations in North Africa and the Middle East: Earth-
Gondwanaland and its biogeographic consequences: Palaeogeography, Science Reviews, v. 93, p. 47–76.
Palaeoclimatology, Palaeoecology, v. 268, p. 126–142. Lesta, P., Mainardi, E., and Stubelj, R., 1980, Plataforma continental argentina:
Isbell, J.L., Miller, M.L., Wolfe, K.L., and Lenaker, P.A., 2003, Timing of Geología Regional Argentina: Córdoba, Argentina, Academia Nacional
late Paleozoic glaciation in Gondwana: Was glaciation responsible for de Ciencias, v. 2, p. 1577–1602.
the development of northern hemisphere cyclothems?, in Chan, M.A., Levell, B.K., Braakman, J.H., and Rutten, K.W., 1988, Oil-bearing sediments
and Archer, A.W., eds., Extreme Depositional Environments: Mega End of Gondwana glaciation in Oman: American Association of Petroleum
Members in Geologic Time: Geological Society of America Special Paper Geologists Bulletin, v. 72, p. 775–796.
370, p. 5–24. Leventhal, J.S., 1987, Carbon and sulfur relationships in Devonian shales from
Isbell, J.L., Miller, M.L., Lenaker, P.A., Koch, Z., and Askin, R., 2005, How the Appalachian Basin as an indicator of environment of deposition:
extensive was Gondwana glaciation in Antarctica?: Mendoza, Argentina, American Journal of Science, v. 287, p. 33–49.
Gondwana 12, Abstracts, p. 208. Li, Z.X., and Powell, C.M., 2001, An outline of the paleogeographic evolu-
Isbell, J.L., Koch, Z., Szablewski, G.M., and Lenaker, P.A., 2008, Permian gla- tion of the Australian regions since the beginning of the Neoproterozoic:
cigenic deposits in the Transantarctic Mountains, Antarctica, in Fielding, Earth-Science Reviews, v. 53, p. 237–277, doi: 10.1016/S0012-8252(00
C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice )00021-0.
Age in Time and Space: Geological Society of America Special Paper Limarino, C.O., Césari, S.N., Net, L.I., Marenssi, S.A., Gutiérrez, P.R., and
441, p. 59–70. Tripaldi, A., 2002, The Upper Carboniferous postglacial transgression
Jervey, M.T., 1988, Quantitative geological modelling of siliciclastic rock in the Paganzo and Río Blanco Basins (northwestern Argentina): Facies
sequences and their seismic expressions, in Wilgus, C.K., Hastings, B.S., and stratigraphic significance: Journal of South American Earth Sciences,
St. Kendall, C.G., Posamentier, H.W., Ross, C.A., and Van Wagoner, J.C., v. 15, p. 445–460, doi: 10.1016/S0895-9811(02)00048-2.
eds., Sea Level Changes: An Integrated Approach: Society of Economic Limarino, C.O., Marenssi, S.A., Tripaldi, A., and Caselli, A.T., 2004, Fjord sedi-
Paleontologists and Mineralogists Special Publication 42, p. 47–69. mentation in the Late Carboniferous of northwestern Argentina: Florence,
Johnson, M.R., Van Vauuren, C.J., Hegenberger, W.F., Key, R., and Shoko, U., Italy, International Geological Congress, no. 32, abstracts 205, p. 960.
1996, Stratigraphy of the Karoo Supergroup in southern Africa: An over- Limarino, C.O., Tripaldi, A., Marenssi, S., and Fauqué, L., 2006, Tectonic, sea-
view: Journal of African Earth Sciences, v. 23, p. 3–15, doi: 10.1016/ level, and climatic controls on Late Paleozoic sedimentation in the west-
S0899-5362(96)00048-6. ern basins of Argentina: Journal of South American Earth Sciences, v. 22,
Jones, A.T., and Fielding, C.R., 2004, Sedimentological record of the late Paleo- p. 205–226, doi: 10.1016/j.jsames.2006.09.009.
zoic glaciation in Queensland, Australia: Geology, v. 32, p. 153–156, doi: Limarino, C.O., Spalletti, L.A., and Colombo-Piñol, F., 2010, Evolución
10.1130/G20112.1. paleoambiental de la transición glacial postglacial en la Formación Agua
Jones, P.J., and Stump, T.E., 1999, Depositional and tectonic setting of the Lower Colorada (Carbonífero, Sierra de Narváez, noroeste argentino): Revista
Silurian hydrocarbon source rock facies, Central Saudi Arabia: American Geológica de Chile (in press).
Association of Petroleum Geologists Bulletin, v. 83, p. 314–332. Lindsay, J.F., 1970, Depositional environment of Paleozoic glacial rocks in
Juan, R.C., de Jager, J., Russell, J., and Gebhard, I., 1996, Flanco norte de la the central Transantarctic Mountains: Geological Society of America
Cuenca del Colorado, in Ramos, V.A., and Turic, M.A., eds., Geología Bulletin, v. 81, p. 1149–1172, doi: 10.1130/0016-7606(1970)81[1149
y Recursos Naturales de la Plataforma Continental Argentina, Relatorio: :DEOPGR]2.0.CO;2.
Buenos Aires, Congreso Geológico Argentino, 13th, and Congreso de Lindsay, J.F., 1997, Permian postglacial environments of the Australian Plate,
Exploración de Hidrocarburos, 3rd, p. 117–133. in Martini, I.P., ed., Late Glacial and Postglacial Environmental Changes,
Keeley, M.L., and Massoud, M.S., 1998, Tectonic controls on the petroleum Quaternary, Carboniferous–Permian and Proterozoic: Oxford, UK, Oxford
geology of NE Africa, in MacGregor, D.S., Moody, R.T.J., and Clark- University Press, p. 213–229.
Lowes, D.D., eds., Petroleum Geology of North Africa: Geological Soci- López-Gamundí, O.R., 1983, Modelo de sedimentación glacimarina para la For-
ety [London] Special Publication 132, p. 265–281. mación Hoyada Verde, Paleozoico superior de la provincia de San Juan:
Keidel, J., 1916, La geología de las sierras de la provincia de Buenos Aires y sus Revista de la Asociación Geológica Argentina, v. 38, p. 60–72.
relaciones con las montañas de Sud Africa y los Andes: Anales Ministerio López-Gamundí, O.R., 1984, Origen y Sedimentología de las diamictitas del
de Agricultura de la Nación, Buenos Aires, Sección Geología, 11 (3). Paleozoico Superior de la República Argentina [Ph.D. thesis]: Buenos
Kneller, B., Milana, J.P., Buckee, C., and Ja’aidi, O., 2004, A depositional record Aires, University of Buenos Aires, 321 p.
of deglaciation in a paleofjord (Late Carboniferous [Pennsylvanian] of López-Gamundí, O.R., 1987, Depositional models for the glaciomarine
San Juan Province, Argentina): The role of catastrophic sedimentation: sequences of Andean Late Paleozoic basins of Argentina: Sedimentary
Geological Society of America Bulletin, v. 116, p. 348–367, doi: 10.1130/ Geology, v. 52, p. 109–126, doi: 10.1016/0037-0738(87)90018-2.
B25242.1. López-Gamundí, O.R., 1989, Postglacial transgressions in Late Paleozoic
Konert, G., Afifi, A.M., Al-Hajri, S.A., and Droste, H.J., 2001, Paleozoic stra- basins of western Argentina: A record of glacioeustatic sea level rise:
tigraphy and hydrocarbon habitat of the Arabian Plate: GeoArabia, v. 6, Palaeogeography, Palaeoclimatology, Palaeoecology, v. 71, p. 257–270,
p. 407–442. doi: 10.1016/0031-0182(89)90054-0.
Kruck, W., and Thiele, J., 1983, Late Palaeozoic glacial deposits in the Yemen López-Gamundí, O.R., 1991, Thin-bedded diamictites in the glaciomarine
Arab Republic: Geologisches Jahrbuch, Reihe B, Regionale Geologie Hoyada Verde Formation (Carboniferous), Calingasta-Uspallata Basin,
Ausland, v. 46, p. 3–29. western Argentina: A discussion on the emplacement conditions of sub-
Le Blanc Smith, G., and Eriksson, K.A., 1979, A fluvioglacial and glaciola- aqueous cohesive debris flows: Sedimentary Geology, v. 73, p. 247–256,
custrine deltaic depositional model for Permo-Carboniferous coals of doi: 10.1016/0037-0738(91)90087-T.
the northeastern Karoo basin, South Africa: Palaeogeography, Palaeo- López-Gamundí, O.R., 1997, Glacial-postglacial transition in the Late Paleozoic
climatology, Palaeoecology, v. 27, p. 67–84, doi: 10.1016/0031-0182(79 basins of southern South America, in Martini, I.P., ed., Late Glacial and
)90094-4. Postglacial Environmental Changes, Quaternary, Carboniferous–Permian
Le Heron, D.P., and Craig, J., 2008, First-order reconstructions of a Late Ordo- and Proterozoic: Oxford, UK, Oxford University Press, p. 147–168.
vician Saharan ice sheet: Journal of the Geological Society [London], López-Gamundí, O.R., and Buatois, L.A., 2010, this volume, Introduction: Late
v. 165, p. 19–29, doi: 10.1144/0016-76492007-002. Paleozoic glacial events and postglacial transgressions in Gondwana, in
Transgressions related to the demise of the Late Paleozoic Ice Age 33

López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial Carboniferous lower Unayzah sandstones, eastern central Saudi Arabia:
Events and Postglacial Transgressions in Gondwana: Geological Society GeoArabia, v. 11, p. 105–152.
of America Special Paper 468, doi: 10.1130/2010.2468(00). Melvin, J., Sprague, R.A., and Heine, C.J., 2010, this volume, From bergs to
López-Gamundí, O.R., and Martínez, M., 2000, Evidence of glacial abrasion in ergs: The late Paleozoic Gondwanan glaciation and its aftermath in Saudi
the Calingasta-Uspallata and western Paganzo basins, mid-Carboniferous Arabia, in López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic
of western Argentina: Palaeogeography, Palaeoclimatology, Palaeoecol- Glacial Events and Postglacial Transgressions in Gondwana: Geological
ogy, v. 159, p. 145–165, doi: 10.1016/S0031-0182(00)00044-4. Society of America Special Paper 468, doi: 10.1130/2010.2468(02).
López-Gamundí, O.R., and Rossello, E.A., 1998, Basin fill evolution and Mésigos, M., 1953, El Paleozoico Superior de Barreal y su continuación aus-
palaeotectonic patterns along the Samfrau geosyncline: The Sauce Grande tral, Sierra de Barreal (Provincia de San Juan): Revista de la Asociación
basin–Ventana foldbelt (Argentina) and Karoo basin–Cape foldbelt Geológica Argentina, v. 8, p. 65–109.
(South Africa): Geologische Rundschau, v. 86, p. 819–834, doi: 10.1007/ Milani, E.J., and De Wit, M.J., 2008, Correlations between the classic Paraná and
s005310050179. Cape–Karoo sequences of South America and southern Africa and their
López-Gamundí, O.R., Espejo, I.S., Conaghan, P.J., and Powell, C.McA., 1994, basin infills flanking the Gondwanides: Du Toit revisited, in Pankhurst,
Southern South America, in Veevers, J.J., and Powell, C.McA., eds., R.J., Trouw, R.A.J., Brito Neves, B.B., and De Wit, M.J., eds., West
Permian-Triassic Pangean Basins and Foldbelts along the Panthalassan Gondwana: Pre-Cenozoic Correlations across the South Atlantic Region:
Margin of Gondwanaland: Geological Society of America Memoir 184, Geological Society [London] Special Publication 294, p. 319–342.
p. 281–329. Miller, J.M.G., 1989, Glacial advance and retreat sequences in a Permo-
Loutit, T.S., Hardenbol, J., Vail, P.R., and Baum, G.R., 1988, Condensed Carboniferous section, central Transantarctic Mountains: Sedimentology,
sections: The key to age dating and correlation of continental margin v. 36, p. 419–430, doi: 10.1111/j.1365-3091.1989.tb00617.x.
sequences, in Wilgus, C.K., Hastings, B.S., St. Kendall, C.G., Posamen- Miller, J.M.G., 1996, Glacial sediments, in Reading, H.G., ed., Sedimentary
tier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., Sea-Level Changes: Environments: Processes, Facies and Stratigraphy (3rd edition): Cam-
An Integrated Approach: Society of Economic Paleontologists and Min- bridge, Massachusetts, Blackwell Science, p. 454–484.
eralogists Special Publication 42, p. 183–213. Miller, M.A., and Melvin, J., 2005, Significant new biostratigraphic horizons in
Lüning, S., Craig, J.D.K., Loydell, Štorch, P., and Fitches, B., 2000, Lower the Qusaiba Member of the Silurian Qalibah Formation of central Saudi
Silurian ‘hot shales’ in North Africa and Arabia: Regional distribution and Arabia, and their sedimentologic expression in a sequence stratigraphic
depositional model: Earth-Science Reviews, v. 49, p. 121–200. context: GeoArabia, v. 10, p. 49–92.
Mahmoud, M.D., Vaslet, D., and Husseini, M.I., 1992, The Lower Silurian Miller, M.F., and Collinson, J.W., 1994, Late Paleozoic post-glacial inland sea
Qalibah Formation of Saudi Arabia: An important hydrocarbon source filled by fine-grained turbidites: Mackellar Formation, central Transant-
rock: American Association of Petroleum Geologists Bulletin, v. 76, arctic Mountains, in Deynoux, M., Miller, J.M.G., Domack, E.W., Eyles,
p. 1491–1506. N., Fairchild, I., and Young, G.M., eds., Earth’s Glacial Record: Cam-
Mángano, M.G., and Buatois, L.A., 2004, Ichnology of Carboniferous tide- bridge, UK, Cambridge University Press, p. 215–233.
Miller, M.F., Collinson, J.W., and Frisch, R.A., 1991, Depositional model and
influenced environments and tidal flat variability in the North American
history of a Permian post-glacial black shale: Mackellar Formation, Cen-
Midcontinent, in McIlroy, D., ed., The Application of Ichnology to Palae-
tral Transantarctic Mountains, in Ulbrich, H., and Rocha-Campos, A.,
oenvironmental and Stratigraphic Analysis: Geological Society [London]
eds., Gondwana Seven Proceedings: São Paulo, Instituto de Geociencias
Special Publication 228, p. 157–178.
University of São Paulo, p. 201–215.
Mángano, M.G., Buatois, L.A., Limarino, C.O., Tripaldi, A., and Caselli, A.,
Mitchell, C., Taylor, G.K., Cox, K.G., and Shaw, J., 1986, Are the Falkland
2003, El icnogénero Psammichnites Torell, 1870 en la Formación Hoyada
Islands a rotated microplate?: Nature, v. 319, p. 131–134, doi: 10.1038/
Verde, Carbonífero Superior de la cuenca Calingasta-Uspallata: Ameghin-
319131a0.
iana, v. 40, p. 601–608. Montañez, I., Tabor, N.J., Niemeier, D., DiMichele, W.A., Frank, T.D., Fielding,
Marshall, J.E.A., 1994, The Falkland Islands: A key element in Gondwana C.R., Isbell, J.L., Birgenheier, L.P., and Rygel, M.C., 2007, CO2-forced
paleogeography: Tectonics, v. 13, p. 499–514, doi: 10.1029/93TC03468. climate and vegetation instability during late Paleozoic deglaciation: Sci-
Martin, J.R., Redfern, J., and Aitken, J.F., 2008, A regional overview of the late ence, v. 315, p. 87–91, doi: 10.1126/science.1134207.
Paleozoic glaciation in Oman, in Fielding, C.R., Frank, T.D., and Isbell, Mussett, A.E., and Taylor, G.K., 1994, 40Ar-39Ar ages for dykes from the Falk-
J.L., eds., Resolving the Late Paleozoic Ice Age in Time and Space: Geo- land Islands with implications for the break-up of southern Gondwana-
logical Society of America Special Paper 441, p. 175–186. land: Journal of the Geological Society [London], v. 151, p. 79–81, doi:
Martínez, M., 1993, Hallazgo de fauna marina en la Formación Guandacol 10.1144/gsjgs.151.1.0079.
(Carbonıfero) en la localidad de Agua Hedionda, San Juan, Precordillera Naqvi, S.M., Rao, D., and Narain, H., 1974, The protocontinental growth of the
Nororiental, Argentina: Compte Rendus XII Congrès International de la Indian shield and the antiquity of its rift valleys: Precambrian Research,
Stratigraphie et Géologie du Carbonifère et Permien, v. 2, p. 291–296. v. 1, p. 345–398, doi: 10.1016/0301-9268(74)90005-9.
Martini, I.P., and Rocha-Campos, A.C., 1991, Interglacial and early post-glacial, Niyogi, D., 1961, Pattern of Talchir sedimentation in Burhai Gondwana basin,
lower Gondwana coal sequences in the Paraná Basin, Brazil, in Ulbrich, Bihar, India: Journal of Sedimentary Petrology, v. 31, p. 63–71.
H.P., and Rocha Campos, A.C., eds., Gondwana Seven Proceedings: São Nystuen, J.P., 1985, Facies and preservation of glaciogenic sequences from
Paulo, Instituto de Geociencias University of São Paulo, p. 317–336. the Varanger Ice Age in Scandinavia and other parts of the North Atlan-
McClure, H.A., 1978, Early Paleozoic glaciation in Arabia: Palaeogeography, tic region: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 51,
Palaeoclimatology, Palaeoecology, v. 25, p. 315–326, doi: 10.1016/0031 p. 209–229, doi: 10.1016/0031-0182(85)90086-0.
-0182(78)90047-0. Ottone, E., 1991, Palynologie du Carbonifère supérieur de la coupe de Mina
McClure, H.A., 1980, Permian–Carboniferous glaciation in the Arabian Pen- Esperanza, Bassin Paganzo, Argentine: Revue de Micropaleontologie,
insula: Geological Society of America Bulletin, v. 91, p. 707–712, doi: v. 34, p. 118–135.
10.1130/0016-7606(1980)91<707:PGITAP>2.0.CO;2. Ovenshine, A.T., 1970, Observations of iceberg rafting in Glacier Bay, Alaska
McGillivray, J.G., and Husseini, M.I., 1992, The Paleozoic petroleum geology and the identification of ancient ice rafted deposits: Geological Society
of central Arabia: American Association of Petroleum Geologists Bulle- of America Bulletin, v. 81, p. 891–894, doi: 10.1130/0016-7606(1970)81
tin, v. 76, p. 1473–1490. [891:OOIRIG]2.0.CO;2.
McLachlan, I.R., and Anderson, A., 1973, A review of the evidence for marine Pasley, M.A., Gregory, W.A., and Hart, G.F., 1991, Organic matter variations in
conditions in Southern Africa during Dwyka times: Palaeontologia Afri- transgressive and regressive shales: Organic Geochemistry, v. 17, p. 483–
cana, v. 15, p. 37–64. 509, doi: 10.1016/0146-6380(91)90114-Y.
McLachlan, I.R., Tsikos, H., and Cairncross, B., 2001, Glendonites (pseudo- Pasley, M.A., Riley, G.W., and Nummedal, D., 1993, Sequence stratigraphic
morphs after ikaite) in Late Carboniferous Marine Dwyka beds in South- significance of organic matter variations: Example from the Upper Cre-
ern Africa: South African Journal of Geology, v. 104, p. 265–272, doi: taceous Mancos Shale of the San Juan Basin, New Mexico, in Katz, B.J.,
10.2113/1040265. and Pratt, L.M., eds., Source Rocks in a Sequence Stratigraphic Frame-
Melvin, J., and Sprague, R.A., 2006, Advances in Arabian stratigraphy: Ori- work: American Association of Petroleum Geologists Studies in Geology
gin and stratigraphic architecture of glaciogenic sediments in Permian- 37, p. 221–241.
34 López-Gamundí

Posamentier, H.W., and Allen, G.P., 1999, Siliciclastic Sequence Stratigraphy— geochemical proxies—Response to climate evolution and sedimentary
Concepts and Applications: SEPM (Society for Sedimentary Geology) response: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 240,
Concepts in Sedimentology and Paleontology 7, 210 p. p. 184–203, doi: 10.1016/j.palaeo.2006.03.059.
Posamentier, H.W., Allen, G.P., James, D.P., and Tesson, M., 1992, Forced Schlutter, S.R., 1998, Characteristics of shale deposition in relation to strati-
regressions in a sequence stratigraphic framework: Concepts, examples, graphic sequence systems tracts, in Schieber, J., Zimmerle, W., and
and exploration significance: American Association of Petroleum Geolo- Sethi, P., eds., Shales and Mudstones: Stuttgart, I.E. Schweizerbart’sche
gists Bulletin, v. 76, p. 1687–1709. Verlagsbuchlandlung, p. 79–108.
Powell, R.D., 1984, Glacimarine processes and inductive lithofacies model- Scotese, C.R., and Barrett, S.F., 1990, Gondwana’s movement over the South
ing of ice shelf and tidewater glacier sediments based on Quaternary pole during the Palaeozoic: Evidence from lithological indicators of cli-
examples: Marine Geology, v. 57, p. 1–52, doi: 10.1016/0025-3227(84 mate, in McKerrow, W.S., and Scotese, C.R., eds., Palaeozoic Biogeog-
)90194-4. raphy and Palaeogeography: Geological Society [London] Memoir 12,
Powell, R.D., Dawber, M., McInnes, J.N., and Pyne, A.R., 1986, Observations p. 75–86.
of the grounding-line area at a floating glacier terminus: Annals of Glaci- Scotese, C.R., and Langford, R.P., 1995, Pangea and the paleogeography of the
ology, v. 22, p. 217–223. Permian, in Scholle, P.A., Peryt, T.M., and Ulmer-Scholle, D.S., eds., The
Rao, C.P., and Green, D.C., 1982, Oxygen and carbon isotopes of Early Perm- Permian of Northern Pangea: Paleogeography, Paleoclimates, Stratigra-
ian cold-water carbonates, Tasmania, Australia: Journal of Sedimentary phy, v. 1, p. 3–19.
Petrology, v. 52, p. 1111–1125. Selleck, B.W., Carr, P.F., and Jones, B.G., 2007, A review and synthesis of glen-
Raymond, R., Jr., and Davies, T.D., 1979, Content and form of sulfur in coal: donites (pseudomorphs after ikaite) with new data: Assessing applicabil-
A reflection of peat depositional environments: Geological Society of ity as recorders of ancient coldwater conditions: Journal of Sedimentary
America Abstracts with Programs, v. 11, p. 285. Research, v. 77, p. 980–991, doi: 10.2110/jsr.2007.087.
Revill, A.T., Volkman, J.K., O’Leary, T., Summon, R.E., Boreham, C.J., Bank, Sengupta, S., Chakraborty, A., and Bhattacharya, H.N., 1999, Fossil Polypla-
M.R., and Dewer, K., 1994, Hydrocarbon biomarkers, thermal maturity, cophora (Mollusca) from the upper Talchir sediments of Dudhi Nala,
and depositional setting of tasmanite oil shales from Tasmania, Australia: Hazaribagh, Bihar: Journal of the Geological Society of India, v. 54,
Geochimica et Cosmochimica Acta, v. 58, p. 3803–3822, doi: 10.1016/ p. 523–527.
0016-7037(94)90365-4. Sessarego, H., and Césari, S., 1989, An early Carboniferous flora from Argen-
Roberts, J., Claoue-Long, J.C., Jones, P.J., and Foster, C.B., 1995, SHRIMP tina: Biostratigraphic implications: Review of Palaeobotany and Palynol-
zircon age control of Gondwanan sequences in Late Carboniferous and ogy, v. 57, p. 247–264, doi: 10.1016/0034-6667(89)90023-7.
Early Permian of Australia, in Dunay, R.E., and Hailwood, E.A., eds., Shearman, D.J., and Smith, A.J., 1985, Ikaite, the parent mineral of jarrowsite-
Non-Biostratigraphical Methods of Dating and Correlation: Geological type pseudomorphs: Proceedings of the Geologists’ Association, v. 96,
Society [London] Special Publication 89, p. 145–174, doi: 10.1144/GSL p. 305–314, doi: 10.1016/S0016-7878(85)80019-5.
.SP.1995.089.01.08. Smith, A.G., Hurley, A.M., and Briden, J.C., 1981, Phanerozoic Palaeoconti-
Roberts, R.J., Hunt, J.W., and Thompson, D.M., 1976, Late Carboniferous marine nental World Maps: Cambridge, UK, Cambridge University Press, 102 p.
invertebrate zones of eastern Australia: Alcheringa, v. 1, p. 197–225. Smith, A.J., 1963, Evidence for a Talchir glaciation striated pavement and
Rocha Campos, A.C., and Carvalho, R.G., 1975, Two new bivalves from the boulder bed at Ira, Central India: Journal of Sedimentary Petrology, v. 33,
Permian “Eurydesma Fauna” of eastern Argentina: São Paulo, Boletim p. 739–750.
Instituto Geologico, Universidade São Paulo, v. 6, p. 181–191. Stanistreet, I.G., Le Blanc Smith, G., and Cadle, A.B., 1980, Trace fossils
Rocha-Campos, A.C., and Rösler, O., 1978, Late Paleozoic faunal and floral as sedimentological and paleoenvironmental indices in the Ecca group
successions in Paraná Basin, Southeastern Brazil: São Paulo, Boletim (Lower Permian) of the Transvaal: Transactions—Geological Society of
Instituto Geologico, Universidade São Paulo, v. 9, p. 1–16. South Africa, v. 83, p. 333–344.
Rocha Campos, A.C., and dos Santos, P.R., 1981, The Itararé Subgroup, Stephenson, M.H., and Osterloff, P.L., 2002, Palynology of the deglaciation
Aquidauana Group and San Gregorio Formation, Paraná Basin, southeast- sequence represented by the Lower Permian Rahab and Lower Gharif
ern South America, in Hambrey, M.J., and Harland, W.B., eds., Earth’s members, Oman: American Association of Stratigraphic Palynologists
Pre-Pleistocene Glacial Record: Cambridge, UK, Cambridge University Contribution Series 40, p. 1–41.
Press, p. 842–852. Sterren, A., and Cisterna, G., 2006, La fauna de Levipustula en la Formación
Rocha Campos, A.C., dos Santos, P.R., and Canuto, J.R., 2008, Late Paleozoic Hoyada Verde: control paleoecológico versus resolución bioestratigráfica:
glacial deposits of Brazil: Paraná Basin, in Fielding, C.R., Frank, T.D., Córdoba, Argentina, Congreso Argentino de Paleontología y Bioestrati-
and Isbell, J.L., eds., Resolving the Late Paleozoic Ice Age in Time and grafía, 9th, Acta de la Academia Nacional de Ciencias, p. 192.
Space: Geological Society of America Special Paper 441, p. 97–114. Stollhofen, H., Stanistreet, I.G., Bangert, B., and Grill, H., 2000, Tuffs, tec-
Rogala, B., James, N.P., and Reid, C.M., 2007, Deposition of polar carbon- tonism and glacially related sea-level changes, Carboniferous–Permian,
ates during interglacial highstands on an Early Permian shelf, Tasma- southern Namibia: Palaeogeography, Palaeoclimatology, Palaeoecology,
nia: Journal of Sedimentary Research, v. 77, p. 587–606, doi: 10.2110/ v. 161, p. 127–150, doi: 10.1016/S0031-0182(00)00120-6.
jsr.2007.060. Storey, B.C., Curtis, M.L., Ferris, J.K., Hunter, M.A., and Livermore, R.A.,
Ross, C.A., and Ross, J.R.P., 1996, Silurian sea level fluctuations, in Witzke, 1999, Reconstruction and break-out model for the Falkland Islands within
B.J., Ludvigson, G.A., and Day, J., eds., Paleozoic Sequence Stratigraphy: Gondwana: Journal of African Earth Sciences, v. 29, p. 153–163, doi:
Views from the North American Craton: Geological Society of America 10.1016/S0899-5362(99)00086-X.
Special Paper 306, p. 187–192. Struckmeyer, H.I.M., and Totterdell, J.M., 1992, Australia, Evolution of a Con-
Runnegar, B., 1979, Ecology of Eurydesma and the Eurydesma fauna, Perm- tinent: Canberra, Australian Bureau of Mineral Resources, 97 p.
ian of eastern Australia: Alcheringa, v. 3, p. 261–285, doi: 10.1080/ Suess, E., Balzer, W., Hess, E.K.F., Muller, P.J., Ungerer, P.J., and Wefer, G.,
03115517908527798. 1982, Calcium carbonate hexahydrate from organic-rich sediments within
Russo, A., Archangelsky, S., Andreis, R.R., and Cuerda, A., 1987, Cuenca the Antarctic shelf: Precursor of glendonites: Science, v. 216, p. 1128–
Chacoparanaense, in Archangelsky, S., ed., El Sistema Carbonífero en la 1131, doi: 10.1126/science.216.4550.1128.
República Argentina: Córdoba, Argentina, Academia Nacional de Cien- Syvitski, J.P.M., Burell, D.C., and Skei, J.M., 1987, Fjords: Processes and Prod-
cias, p. 197–212. ucts: New York, Springer-Verlag, 379 p.
Scasso, R.A., and Mendía, J.E., 1985, Rasgos estratigraficos y paleoambientales Syvitski, J.P.M., Andrews, J.T., and Dowdeswell, J.A., 1996, Sediment deposi-
del Paleozoico de las Islas Malvinas: Revista de la Asociación Geológica tion in an iceberg-dominated glacimarine environment, East Greenland:
Argentina, v. 40, p. 26–50. Basin fill implications: Global and Planetary Change, v. 12, p. 251–270,
Scheffler, K., Hoemes, S., and Schwark, L., 2003, Global changes during Car- doi: 10.1016/0921-8181(95)00023-2.
boniferous–Permian glaciation of Gondwana: Linking polar and equato- Tankard, A.J., Jackson, M.P.A., Eriksson, K.A., Hobday, D.K., Hunter, D.R.,
rial climate evolution by geochemical proxies: Geology, v. 31, p. 605–608, and Minter, W.E.L., 1982, Crustal Evolution of Southern Africa: New
doi: 10.1130/0091-7613(2003)031<0605:GCDCGO>2.0.CO;2. York, Springer-Verlag, 523 p.
Scheffler, K., Buchmann, D., and Schwark, L., 2006, Analysis of late Paleo- Taylor, G.K., and Shaw, J., 1989, The Falkland Islands: New palaeomag-
zoic glacial to postglacial sedimentary successions in South Africa by netic data and their origin as a displaced terrane from southern Africa,
Transgressions related to the demise of the Late Paleozoic Ice Age 35

in Hillhouse, J.W., ed., Deep Structure and Past Kinematics of Accreted geography, Palaeoclimatology, Palaeoecology, v. 70, p. 377–391, doi:
Terranes: American Geophysical Union Geophysical Monograph 50, 10.1016/0031-0182(89)90115-6.
p. 59–72. Visser, J.N.J., 1991, The paleoclimatic setting of the late Paleozoic marine
Theron, J.N., and Blignault, H.J., 1975, A model for the sedimentation of the ice sheet in the Karoo Basin of southern Africa, in Anderson, J.B., and
Dwyka glacials in the southwestern Cape, in Campbell, R.S.W., ed., Ashley, G.M., eds., Glacial Marine Sedimentation: Paleoclimatic Signifi-
Gondwana Geology: Canberra, Australia, Australian National University cance: Geological Society of America Special Paper 261, p. 181–189.
Press, p. 347–356. Visser, J.N.J., 1993, Sea-level changes in a back-arc–foreland transition: The
Thomas, G.S.P., and Connell, R.J., 1985, Iceberg drop, dump, and grounding Late Carboniferous-Permian Karoo basin of South Africa: Sedimentary
structures from Pleistocene glacio-lacustrine sediments, Scotland: Journal Geology, v. 83, p. 115–131, doi: 10.1016/0037-0738(93)90185-8.
of Sedimentary Petrology, v. 55, p. 243–249. Visser, J.N.J., 1997a, A review of the Permo–Carboniferous glaciation in
Thomas, S.G., Frank, T.D., and Fielding, C., 2005, Glendonites as paleoclimate Africa, in Martini, I.P., ed., Late Glacial and Postglacial Environmental
indicators within the Middle Permian Wandrawandian Siltstone, southern Changes, Quaternary, Carboniferous–Permian and Proterozoic: Oxford,
Sydney Basin, Australia: Geological Society of America Abstracts with UK, Oxford University Press, p. 169–191.
Programs, v. 36, no. 5, p. 16. Visser, J.N.J., 1997b, Deglaciation sequences in the Permo-Carboniferous
Tissot, B.P., and Welte, D.H., 1984, Petroleum Formation and Occurrence (2nd Karoo and Kalahari basins of southern Africa: A tool in the analysis of
edition): New York, Springer-Verlag, 699 p. cyclic glaciomarine basin fills: Sedimentology, v. 44, p. 507–521, doi:
Trewin, N.H., MacDonald, D.I.M., and Thomas, C.G.C., 2002, Stratigraphy and 10.1046/j.1365-3091.1997.d01-35.x.
sedimentology of the Permian of the Falkland Islands: Lithostratigraphic Von Brunn, V., 1994, Glaciogenic deposits of the Permo-Carboniferous Dwyka
and palaeoenvironmental links with South Africa: Journal of the Geologi- Group in the eastern region of the Karoo Basin, South Africa, in Deynoux,
cal Society [London], v. 159, p. 5–19, doi: 10.1144/0016-764900-089. M., Miller, J.M.G., Domack, E.W., Eyles, N., Fairchild, I., and Young,
Turner, B.R., Armstrong, H.A., Makhlouf, I.M., and Bourne, T.J., 2005, High G.M., eds., Earth’s Glacial Record: Cambridge, UK, Cambridge Univer-
latitude, east Gondwana glaciation: Glacio-fluvial palaeovalleys inter- sity Press, p. 60–69.
preted as tunnel valleys: Mendoza, Argentina, Gondwana 11, Abstracts, Von Brunn, V., 1996, The Dwyka Group in the northern part of Kwazulu/Natal,
p. 356. South Africa: Sedimentation during late Palaeozoic deglaciation: Palaeo-
Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M., Vail, P.R., Sarg, J.F., geography, Palaeoclimatology, Palaeoecology, v. 125, p. 141–163, doi:
Loutit, T.S., and Hardenbol, J., 1988, An overview of sequence stra- 10.1016/S0031-0182(96)00028-4.
tigraphy and key definitions, in Wilgus, C.K., Hastings, B.S., Kendall, Von Brunn, V., and Gravenor, C.P., 1983, A model for late Dwyka glacioma-
C.G.St.C., Posamentier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., rine sedimentation in the eastern Karoo Basin: Transactions—Geological
Sea Level Changes—An Integrated Approach: Special Publication of the Society of South Africa, v. 86, p. 199–209.
Society of Economic Paleontologists and Mineralogists (SEPM), v. 42, Wickens, H., de V., 1992, Submarine fans of the Permian Ecca Group in the
p. 39–45. SW Karoo Basin: Their origin and reflection on the tectonic evolution of
Van Wagoner, J.C., Mitchum, R.M., Jr., Campion, K.M., and Rahmanian, V.D., the basin and its source areas, in De Wit, M.J., and Ransome, I.G.D., eds.,
1990, Siliciclastic Sequence Stratigraphy in Well Logs, Core and Out- Inversion Tectonics of the Cape Fold Belt, Karoo and Cretaceous Basins
crops: Concepts for High-Resolution Correlation of Time and Facies: of Southern Africa, p. 117–125.
American Association of Petroleum Geologists Methods in Exploration Wignall, P.B., 1991, Model for transgressive black shales?: Geology, v. 19,
Series 7, 55 p. p. 167–170, doi: 10.1130/0091-7613(1991)019<0167:MFTBS>2.3.CO;2.
Vaslet, D., 1990, Upper Ordovician glacial deposits in Saudi Arabia: Episodes, Wignall, P.B., 1994, Black Shales: Oxford, UK, Oxford Monographs in Geol-
v. 13, p. 147–161. ogy and Geophysics 30, 127 p.
Veevers, J.J., and Powell, C.M., 1987, Late Paleozoic glacial episodes in Gond- Williams, E.G., and Keith, M.L., 1963, Relationship between sulfur in coals
wanaland reflected in transgressive-regressive depositional sequences in and the occurrence of marine roof beds: Economic Geology and the Bul-
Euramerica: Geological Society of America Bulletin, v. 98, p. 475–487, letin of the Society of Economic Geologists, v. 58, p. 720–729.
doi: 10.1130/0016-7606(1987)98<475:LPGEIG>2.0.CO;2. Williams, R., Jr., and Ferrigno, J.G., eds., 2009, Chapter A, The State of the
Veevers, J.J., and Tewari, R.C., 1995, Gondwana master basin of Peninsular Earth’s Cryosphere at the Beginning of the 21st Century, in Glaciers,
India between Tethys and the interior of the Gondwanaland province of Global Snow Cover, Floating Ice, and Permafrost and Periglacial Envi-
Pangea: Geological Society of America Memoir 187, 72 p. ronments: U.S. Geological Survey Professional Paper 1386-A (in press).
Veevers, J.J., Cole, D.I., and Cowan, E.J., 1994, Southern Africa: Karoo Basin Winn, R.D., and Steinmetz, J.C., 1998, Upper Paleozoic strata of the Chaco-
and Cape Fold Belt, in Veevers, J.J., and Powell, C.McA., eds., Permian- Parana basin, Argentina, and the great Gondwana glaciation: Journal of
Triassic Pangean Basins and Foldbelts along the Panthalassan Margin of South American Earth Sciences, v. 11, p. 153–168, doi: 10.1016/S0895
Gondwanaland: Geological Society of America Memoir 184, p. 223–280. -9811(98)00007-8.
Venkatachala, B.S., and Tiwari, R.S., 1988, Lower Gondwana marine incur- Wopfner, H., and Casshyap, S.M., 1997, Transition from freezing to subtropical
sions: Periods and pathways: Palaeobotanist, v. 36, p. 24–29. climates in the Permo-Carboniferous of Afro-Arabia and India, in Mar-
Vesely, F.F., and Assine, M.L., 2004, Seqüências e tratos de sistemas deposi- tini, I.P., ed., Late Glacial and Postglacial Environmental Changes, Qua-
cionais do Grupo Itararé, norte do estado do Paraná: Revista Brasileira de ternary, Carboniferous-Permian and Proterozoic: Oxford, UK, Oxford
Geociencias, v. 34, p. 219–230. University Press, p. 192–212.
Visser, J.N.J., 1983, An analysis of the Permo-Carboniferous glaciation in the Wopfner, H., and Kreuser, T., 1986, Evidence for Late Paleozoic glaciation in
marine Kalahari Basin, South Africa: Palaeogeography, Palaeoclimatology, southern Tanzania: Palaeogeography, Palaeoclimatology, Palaeoecology,
Palaeoecology, v. 44, p. 295–315, doi: 10.1016/0031-0182(83)90108-6. v. 56, p. 259–275, doi: 10.1016/0031-0182(86)90098-2.
Visser, J.N.J., 1986, Lateral lithofacies relationships in the glacigene Dwyka Wright, R., Anderson, J.B., and Fisco, P.P., 1983, Distribution and association
Formation in the western and central parts of the Karoo Basin: Transac- of sediment gravity flow deposits and glacial/glacial-marine sediments
tions—Geological Society of South Africa, v. 89, p. 373–383. around the continental margin of Antarctica, in Molnia, B., ed., Glacial-
Visser, J.N.J., 1987a, The palaeogeography of part of southwestern Gond- Marine Sedimentation: New York, Plenum Publishing, p. 265–300.
wana during the Permo–Carboniferous glaciation: Palaeogeography, Zalán, P.V., Wolff, S., Conceição, J.C., Marques, A., Astolfi, M.A., Vieira, I.S.,
Palaeoclimatology, Palaeoecology, v. 61, p. 205–219, doi: 10.1016/0031 Appi, V.T., and Zanotto, O.A., 1990, Bacia do Paraná, in Raja Gabaglia,
-0182(87)90050-2. G.P., and Milani, E.J., eds., Origem e Evolução das Bacias Sedimentares:
Visser, J.N.J., 1987b, The influence of topography on the Permo–Carbonifer- Rio de Janeiro, Petrobras, p. 135–168.
ous glaciation in the Karoo Basin and adjoining area, southern Africa,
in McKenzie, G.D., ed., Gondwana Six: American Geophysical Union
Geophysical Monograph 41, p. 123–129.
Visser, J.N.J., 1989, The Permo-Carboniferous Dwyka Formation of Southern
Africa: Deposition by a predominantly subpolar marine ice sheet: Palaeo- MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

From bergs to ergs: The late Paleozoic Gondwanan


glaciation and its aftermath in Saudi Arabia

John Melvin
Ronald A. Sprague
Christian J. Heine
Saudi Aramco, Box 12646, Dhahran 31311, Saudi Arabia

ABSTRACT

The late Paleozoic (Carboniferous–Permian) Gondwanan glaciation is repre-


sented in the subsurface of eastern and central Saudi Arabia by the Hercynian (or
pre-Unayzah) unconformity and the lower part of the overlying Unayzah Formation.
The subsequent postglacial transgression is manifest in the upper members of the
Unayzah Formation as well as the lowermost clastic deposits of the overlying Khuff
Formation. Its component sediments result from ongoing climatic amelioration fol-
lowing the demise of the ice age, as well as tectonic influences related to the creation
of the Neotethys Ocean.
The Unayzah Formation is subdivided into four stratigraphic members. Thus,
directly overlying the Hercynian unconformity in many places are sandstones and
minor conglomerates of the Unayzah C member. These were laid down within a wide-
spread, braided glaciofluvial depositional system. They represent glacial outwash
produced during times of glacial retreat throughout the duration of the late Paleo-
zoic glaciation. An unknown number of glacial readvances occurred that significantly
deformed these retreat-phase outwash sands and gravels, creating major glacially
tectonized push moraine nappes. Those are interpreted from a number of distinctive
and discrete shear zones that are uniquely associated with the Unayzah C member.
The upper surface of the Unayzah C member is an unconformity that marks the final
subglacial surface at the time of maximum advance of the ice.
The terminal melt-out phase of the Gondwanan glaciation is represented by the
Unayzah B member. Paleomagnetic evidence suggests that this member was depos-
ited at high latitudes, ~75° S. This member comprises a large number of depositional
facies that are essentially glaciolacustrine in character. Those facies include (1) small-
scale ice-contact push moraines indicative of minor glacial readvance, (2) ice-proximal
sublacustrine debris flows (massive diamictites) and associated gravity flow deposits,
and (3) ice-distal, sublacustrine stratified diamictites, ripple cross-laminated sand-
stones, and laminated mudrocks. Facies associations within the Unayzah B member
consistently show evidence of sustained glacial retreat and flooding of the landscape
by filling and spilling over of numerous glacial lakes. This flooding sequence probably

Melvin, J., Sprague, R.A., and Heine, C.J., 2010, From bergs to ergs: The late Paleozoic Gondwanan glaciation and its aftermath in Saudi Arabia, in López-
Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana: Geological Society of America Special
Paper 468, p. 37–80, doi: 10.1130/2010.2468(02). For permission to copy, contact [email protected]. ©2010 The Geological Society of America. All
rights reserved.

37
38 Melvin et al.

represents the maximum climatically related postglacial transgressive event in Saudi


Arabia. In the western part of the study area there is evidence that the ice remained
longer, and it is tentatively interpreted as a local center for high altitude (“alpine”)
glaciation.
Deposition of the Unayzah B member was terminated abruptly by a drainage
event that is marked by a widespread sharp contact with the overlying unnamed mid-
dle Unayzah member. The latter member displays no unequivocally glacially related
depositional facies, and paleomagnetic data suggest that it was deposited at ~55° S. It
is dominated by red floodplain siltstones and very fine–grained sandstones that con-
tain relatively isolated bodies of coarser fluvial and eolian sandstones. These eolianites
display possible cold-climate characteristics. This is particularly true in the western
part of the study area, which is consistent with a relatively sustained, high altitude
ice cap in that area. The unnamed middle Unayzah member is capped in many places
by a paleosol horizon. This represents a hiatus of unknown but probably prolonged
duration and thus suggests a disconformable contact between the unnamed middle
Unayzah member and the overlying Unayzah A member.
The Unayzah A member is dominated by sediments that are strongly character-
istic of terrestrial deposition in a semiarid to arid environment (including ephemeral
lakes and streams as well as eolian deposits). Paleomagnetic data suggest paleolati-
tudes ~28° S. The continental eolian clastic deposits of this member in places display a
cyclicity in their stratal architecture that is related to fluctuations in the paleo–water
table. These fluctuations are possibly related to distant marine transgression, which
is supported by the occurrence of a distinctive bioturbated sandstone very close to
the top of the Unayzah A member. That marine-influenced sandstone is observed in
widely separated localities at either end of the study area and may represent the final
breakthrough of transgressive marine waters close to the end of Unayzah A time.
In several places the uppermost deposits of the Unayzah A member are char-
acterized by thick paleosols. These represent a prolonged period of nondeposition,
interpreted to be directly related to thermal doming of the Arabian plate prior to
rifting and opening of the Neotethys Ocean, and the consequent formation of the pre-
Khuff unconformity, which terminated Unayzah deposition. Overlying the pre-Khuff
unconformity are various siliciclastic facies of the eponymous Basal Khuff Clastics
member of the Khuff Formation. The depositional facies of the lowermost Basal
Khuff Clastics range from shallow marine in the southeastern part of the study area
to predominantly fluvial in the west. This reflects the westward-directed transgression
of the Khuff Formation following thermal collapse in the wake of the rifting that cre-
ated the Neotethys Ocean. That tectonically related transgression reached its fullest
expression with deposition of the carbonates and evaporites that dominate the upper
members of the Khuff Formation. This stratigraphic evolution of the late Paleozoic
in Saudi Arabia can confidently be correlated in sequence stratigraphic terms with
coeval sediments laid down across the Arabian Peninsula.

INTRODUCTION unconformity in Saudi Arabia. This unconformity is widely rec-


ognized in the subsurface across the study area of eastern central
The Arabian plate was subjected to a major episode of tec- Saudi Arabia (Fig. 1A), and it truncates older formations rang-
tonic deformation during the middle Carboniferous (Husseini, ing in age from Proterozoic to Early Carboniferous. This tectonic
1992; McGillivray and Husseini, 1992; Wender et al., 1998; event was closely followed by the inception of the Late Paleozoic
Al-Hajri and Owens, 2000; Konert et al., 2001; Sharland et al., Ice Age in Arabia (Sharland et al., 2001; Al-Husseini, 2004). The
2001; Al-Husseini, 2004). There followed a period of erosion depositional products of that late Paleozoic Gondwanan glacial
that lasted from Namurian to early Westphalian (Serpukhovian to episode were laid down in the first instance upon the Hercyn-
Bashkirian) times (Al-Husseini, 2004), and which is represented ian unconformity. The subsequent postglacial transgression was
in the stratigraphic record by the Hercynian (or pre-Unayzah) affected during the Permian by tectonic events that related to
Late Paleozoic Gondwanan glaciation in Saudi Arabia 39

Figure 1. Location maps. (A) Map of


eastern central Saudi Arabia, showing
the locations of the wells studied. Note
also the location of the Al Batin Arch
(Faqira et al., 2009) in the western part
of the study area. (B) Map of the Arabian
Peninsula, showing the principal study
area for this paper, as well as areas in
Oman (A), Yemen (B), and southwest-
ern Saudi Arabia (C), where glaciogenic
sediments of Carboniferous–Permian
age crop out.

opening of the Neotethys Ocean. The geological record of these where it crops out in the southwestern Wajid region (Helal, 1964;
events is preserved in Saudi Arabia in the siliciclastic deposits Kellogg et al., 1986; McClure, 1980; McClure and Young, 1981;
of the Unayzah Formation and the clastic rocks that occur at the McClure et al., 1988) (Fig. 1B) and dated as Late Carboniferous
base of the overlying Khuff Formation in Saudi Arabia, and cul- (Stephanian = Moscovian to Gzhelian) to Early Permian (Sak-
minates in the carbonates that dominate in the higher parts of the marian) in age (Cameron and Hemer in McClure, 1980; McClure
Khuff Formation. and Young, 1981). Since the late 1980s a large number of wells
The depositional record of the Gondwanan glaciation on the have penetrated the Khuff and Unayzah Formations in the sub-
Arabian plate has been well documented from Oman (Braakman surface of Saudi Arabia as part of an aggressive exploration and
et al., 1982; Hughes Clarke, 1988; Levell et al., 1988; Al-Belushi exploitation program in the search for oil and non-associated gas.
et al., 1996; Aitken et al., 2004; Osterloff et al., 2004a) as well The recovery of many thousands of feet of core in that drilling
as from northern Yemen (Roland, 1979; Kruck and Thiele, 1983) effort has enabled much progress to be made in understanding the
(Fig. 1B). In Saudi Arabia, boulder-bearing diamictites of inferred origin of these formations and their constituent members, as well
glacial origin have been recognized in the Juwayl Formation as their stratigraphic distribution (architecture) in the subsurface
40 Melvin et al.

of central and eastern Saudi Arabia. Detailed studies are ongoing, Al-Duaiji, 1995). This is recognizable throughout the subsurface
and some results have already been released (e.g., Melvin and (Al-Husseini, 2004) and is overlain by interbedded sandstones,
Heine, 2004; Melvin et al., 2005; Melvin and Sprague, 2006). shales, and thin carbonates of the Basal Khuff Clastics member
The current paper discusses the nature, origin, and stratigra- of the Khuff Formation (Fig. 2).
phy of siliciclastic deposits representing the Unayzah Formation Notwithstanding the tripartite subdivision of the subsurface
in the subsurface of east-central Saudi Arabia, as well as over- Unayzah by Ferguson and Chambers (1991), until very recently
lying siliciclastic sediments that occur at the base of the other- there had existed no rigorous definition of its constituent mem-
wise predominantly carbonate Khuff Formation. The rocks are bers. This lack of definition arises from the intrinsic variability
described and considered in terms of their relationship to the in the wireline-log character of the stratigraphic units among the
Gondwanan glaciation per se, and also, as appropriate, in terms wells. That variability stems from extreme variations in thickness
of the extent to which they reflect the development of the late and facies among the members of the Unayzah, as well as from an
Paleozoic, postglacial history of siliciclastic sedimentation in abundance of erosional and nondepositional hiatuses throughout
Saudi Arabia. Twenty-four representative cored wells were inves- the formation. Melvin and Sprague (2006), in an extensive core-
tigated (Fig. 1A), and a total of 6351 ft (1905.3 m) of core was based study, revisited the lower Unayzah (i.e., the pre-Unayzah
described and interpreted. The cored intervals were integrated A part of the formation). Following a detailed description of the
with conventional wireline logs, and a stratigraphic framework various depositional facies within these rocks, they redefined the
has been constructed that places the various core-derived deposi- stratigraphy with the introduction of a new, “un-named middle
tional facies in their architectural context. Unayzah member” that occurs between the Unayzah B member
and the Unayzah A member of Ferguson and Chambers (1991)
STRATIGRAPHIC SETTING (Fig. 2). The continued extensive use of cores and core-related
information from the present study has enabled further consoli-
Sharland et al. (2001) presented an evaluation of the com- dation of our understanding of the lithostratigraphic relation-
plete sequence stratigraphy of the entire Arabian plate. Those ships within and among the members of the Unayzah and the
authors recognized a number of “Tectonostratigraphic Megase- overlying Basal Khuff Clastics. All of these members of the
quences” (TMS) that provided the foundations for that sequence Unayzah Formation in the subsurface of Saudi Arabia continue
stratigraphic framework. Specifically, their TMS AP5 mega- to be represented by an informal working nomenclature. Given
sequence spans the period from the middle Carboniferous “Her- the enhanced understanding of these rocks that is in large part
cynian” tectonic event to the early Late Permian rifting event that represented herein, the need for a revised, formalized lithostrati-
created the Neotethys Ocean (cf. Bishop, 1995; Loosveld et al., graphic nomenclature scheme within the Unayzah has been rec-
1996) (Sharland et al., 2001). TMS AP5 is the last TMS to be ognized and is being addressed by Saudi Aramco. The results
dominated by siliciclastic sediments prior to the movement of of that reexamination of the stratigraphic nomenclature will be
the plate into subtropical latitudes where carbonate and evapo- published in a separate, future paper.
rite deposition became dominant (cf. Beydoun, 1991; Al-Fares
et al., 1998). The base of this megasequence is marked by the Biostratigraphy
Hercynian unconformity (Fig. 2). The top of TMS AP5 is the
pre-Khuff unconformity of Senalp and Al-Duaiji (1995), Evans It is extremely difficult to obtain definitive age dates from
et al. (1997), and Wender et al. (1998) (Sharland et al., 2001). the Carboniferous-Permian succession of the Arabian plate, and
The current paper will therefore discuss in its entirety the TMS in particular in Saudi Arabia from the Unayzah Formation. There,
AP5 megasequence of Sharland et al. (2001) (viz. the Unayzah the primary biostratigraphic tool has been limited to palynology,
Formation), as well as the lowermost siliciclastic deposits of the owing to a lack of macrofossils in these predominantly terrestrial
superseding TMS AP6 megasequence, namely the Basal Khuff rocks. Furthermore, many well sections within the Unayzah are
Clastics member of the Khuff Formation that abruptly overlies barren of any fossils, including palynomorphs. Nonetheless, fol-
the pre-Khuff unconformity. lowing earlier work by Stephenson (1998), Al-Hajri and Owens
(2000), and Stephenson and Filatoff (2000a, 2000b), Stephen-
Lithostratigraphy son et al. (2003) produced a palynological zonation scheme that
enables correlation to a varying degree among all the various
The Unayzah Formation in the subsurface of eastern cen- members of the Unayzah Formation, as well as the Basal Khuff
tral Saudi Arabia sits directly upon the Hercynian unconformity. Clastics, in Saudi Arabia, with the Al Khlata, Gharif, and the
Historically, it has been subdivided informally into three mem- lower part of the Khuff Formations of Oman (Fig. 2). This paly-
bers, the Unayzah A (youngest), B, and C (oldest) (Ferguson nozonation is supported in part by paleontological studies carried
and Chambers, 1991; McGillivray and Husseini, 1992). It var- out in Oman by Angiolini et al. (2006) and was recently updated
ies considerably in thickness, with the greatest thickness varia- and modified by Stephenson (2006). Thus, the Unayzah C has
tions taken up in the Unayzah B and C members. The Unayzah A been assigned to the OSPZ1 Zone by Stephenson et al. (2003),
member is truncated by the pre-Khuff unconformity (Senalp and and is considered by those authors to be Late Carboniferous
Late Paleozoic Gondwanan glaciation in Saudi Arabia 41

A B
KSA / Oman
KSA subsurface stratigraphy Oman lithostratigraphy
Stratigraphic biostratigraphy
Stephenson et al. (2003)
TMS*
units (Sharland et al., 2001)
This paper and
Angiolini et al. (2006)
Levell et al.
(1988)
Osterloff et al.
(2004a, 2004b)
Late
Changhsingian
TMS Khuff DS
Wuchiapingian AP6 Lower P17
Basal Khuff thru
Khuff OSPZ 6 DS
Middle

Capitanian
P19

Gharif Formation
Wordian hiatus hiatus Upper
OSPZ 5 DS P15
Roadian Gharif
PERMIAN

Kungurian Middle
Unayzah A OSPZ 4 DS P13
member Gharif
Early

Artinskian hiatus

Unnamed middle c Lwr


Unayzah member
OSPZ 3 b
hiatus a Gharif DS P10
Sakmarian
Unayzah B Rahab Shale DS P8
TMS
member OSPZ 2 Extensive
AP5 glacio-lacustrine DS P6
Asselian
hiatus
CARBONIFEROUS

DS CP
Unayzah C
Gzhelian

OSPZ 1
member
Lower
(multiple
Late

Al Khlata
hiatuses)
Moscovian

DS C30

Hercynian Unconformity

hiatus hiatus

Figure 2. Late Carboniferous-Permian stratigraphy of Saudi Arabia (KSA) and Oman. Note: (A) Tectonostratigraphic
Megasequences (TMS) of Sharland et al. (2001); (B) biostratigraphic zonation (OSPZ1–OSPZ6) of Stephenson et al.
(2003) that provides critical linkage between the successions of Saudi Arabia and Oman.

(Stephanian = Moscovian to Gzhelian) in age. It is equated with 1988). Those rocks are defined by Osterloff et al. (2004a) as the
the lower part of the Al Khlata Formation in Oman, specifically upper part of depositional sequence DS CP, as well as DS P6 and
depositional sequence DS C30 and the lower part of DS CP of DS P8 in Oman (Fig. 2).
Osterloff et al. (2004a) (Fig. 2). Stephenson and Filatoff (2000a), The correlation of rock units between Saudi Arabia and
Stephenson et al. (2003), and Stephenson (2004) showed that the Oman using the OSPZ3 Zone (and its three subbiozones) is
Unayzah B member is Early Permian (Asselian to Sakmarian) more challenging. These subbiozones are restricted in occur-
in age and is characterized by palynomorph assemblages of the rence, and so may not be recognizable throughout Arabia owing
OSPZ2 Zone. This is equated with the upper part of the Al Khlata to either paleophytogeographical variation or hiatus (Stephen-
Formation of Oman (including the Rahab Shale of Levell et al., son et al., 2003). Thus, although the Haushi Limestone at the
42 Melvin et al.

top of the lower Gharif member in Oman was shown by Angio- UNAYZAH C MEMBER
lini et al. (2006) to correlate with the OSPZ3c Subbiozone, in
Saudi Arabia Stephenson et al. (2003) considered the OSPZ3 Depositional Characteristics
Biozone to be absent, at least in part, and so to represent a de-
positional hiatus at the time of deposition of the lower Gharif A total of 1160 ft (348 m) of core from the Unayzah C mem-
member. Melvin and Sprague (2006) suggested that their “un- ber was examined in detail in this study. Characteristically, it is a
named middle Unayzah member” of the Unayzah in Saudi Ara- hard, quartz-cemented sandstone unit that is extremely variable
bia may be at least partially the terrestrial lateral equivalent of in thickness, reflecting deposition upon a strongly undulating
the marine lower Gharif member of Oman (Fig. 2). This would surface (the Hercynian unconformity). Although absent in some
give it some degree of equivalence with the OSPZ3 Biozone of places, it can be present as several hundred feet of sediment only
Stephenson et al. (2003) and suggests correlation in Oman with a few kilometers away. In general, it displays the most distinctive
depositional sequence DS P10 of Osterloff et al. (2004b). In this and reproducible wireline-log signature of any of the members of
context, Stephenson (2006, written commun.) acknowledged the Unayzah Formation in the Saudi Arabian subsurface. Thus it
that in Saudi Arabia the OSPZ3 Zone, rather than being absent, is characterized by a monotonous, very low API gamma-ray log
could alternatively be placed “against some barren section” in from most wells, and similarly an essentially featureless sonic log
the Unayzah. Any hiatus in Saudi Arabia thus would have been with very low (fast) sonic transit times. These featureless wire-
of more limited duration than originally implied by Stephenson line traces are interrupted locally by high gamma and slow sonic
et al. (2003). Angiolini et al. (2006) demonstrated that the lower spikes, respectively. In core, the sandstones are arranged in multi-
Gharif member in Oman is late Sakmarian in age; by implication story bedsets that are commonly several tens of feet thick (Fig. 3).
this age would also apply to the unnamed middle Unayzah mem- Within these bedsets, individual beds pass upward from sharp,
ber. The Unayzah A member in Saudi Arabia has been broadly erosional basal surfaces, locally strewn with intraformational
correlated with the OSPZ4 Biozone (and possibly the OSPZ3b mudclasts and pebbles and cobbles of both massive and lami-
and OSPZ3c Subbiozones) by Stephenson et al. (2003). This nated sandstone (Fig. 3A, 10.5–12.5 ft; Fig. 3B, 1–9 ft; Fig. 4A),
correlation suggests a general equivalence to the middle Gharif into upward-fining, poorly to moderately sorted, very coarse– to
member of Oman (depositional sequence DS P13 of Osterloff medium-grained sandstones. Primary depositional structures are
et al., 2004b) (Fig. 2). hard to identify in these sandstones, although in places the rocks
Stephenson et al. (2003, p. 486) observed that the palyno- may display low-angle planar lamination, or less commonly
logical changeover between OSPZ4 and OSPZ5 is “probably trough cross-lamination (Fig. 3). In some wells (e.g., well 13)
the greatest recorded in the Permian palynological succession in beds within the Unayzah C member contain abundant mud clasts,
Oman and Saudi Arabia.” They also noted how in Oman there up to 3 cm long, that are uniformly dispersed throughout the bed,
is an abrupt change in depositional environment, as well as in and oriented more or less parallel to bedding (Fig. 3B). Notwith-
chemostratigraphy and heavy mineral assemblages, between the standing the above-described features, very commonly the rock
middle and upper members of the Gharif Formation. Further- is either “massive” or characterized only by an extremely diffuse,
more, there is evidence for local incision and paleosol develop- poorly sorted fabric that passes up into discontinuous, crinkly,
ment in the upper part of the middle Gharif member (Stephenson argillaceous laminations that may be enhanced by stylolitization
et al., 2003; Osterloff et al., 2004b), which suggests a significant (see below). Rarely, thin (centimeter scale) beds of gray-green,
depositional hiatus between the middle and upper Gharif mem- sandy siltstone separate the sandstone bedsets.
bers. It will be demonstrated in the present paper that similar The uppermost contact of the Unayzah C member has been
changes occur between the top of the Unayzah A member and cored in a number of wells, and it is in all cases extremely sharp.
the Basal Khuff Clastics, and that they are thus similarly sepa- Melvin and Sprague (2006) described how in well 8, the contact
rated by an unconformity of considerable significance. The Basal is also extremely irregular, with reentrants that are infilled by sed-
Khuff Clastics and the upper Gharif member are therefore prob- iment of the overlying Unayzah B member. Beneath that irregular
ably equivalent lithostratigraphic units, as was implicit in strati- contact, the uppermost 1 ft (30 cm) of the Unayzah C member in
graphic diagrams presented by Sharland et al. (2001, p. 85) and this well displays a high degree of brecciation, with a network of
Stephenson et al. (2003, p. 470). These units are not, however, abundant, intersecting clay-lined fractures. This has been inter-
time-equivalent, as the upper Gharif member is characterized by preted to be an immature paleosol horizon (Melvin and Sprague,
the OSPZ5 Palynozone, whereas the Basal Khuff Clastics mem- 2006). In most wells where the upper contact has been cored, the
ber is definitively represented by the younger OSPZ6 Zone (Ste- Unayzah C member is overlain by the Unayzah B member (e.g.,
phenson, 2006). This diachroneity is considered by the present Figs. 3A and 3B). In well 11 it is directly overlain by distinctive
authors to be a reflection of the onset in Oman of the major trans- facies of the Unayzah A member (Fig. 3C). All of these strati-
gression that culminated in widespread deposition of carbonates graphic relationships at the top of the Unayzah C member point
of the Khuff Formation and which is related to the opening of the strongly to that contact as an unconformity, the significance of
Neotethys Ocean. which is discussed below.
Figure 3. Core logs from wells 7, 13, and 11, showing the fundamental character of the Unayzah C member. Note that in all wells, the
abundance of “structural features” (stylolites, fractures, etc.) is highlighted in gray, and that these are uniquely limited to the Unayzah
C member. In addition, note the intensely sheared zones in well 7 (between 0 and 10 ft) and well 13 (19–20 ft, 57–60 ft, 75–85 ft).
The rocks in well 13 are characterized by an abundance of mud clasts dispersed throughout the beds. The Unayzah C member can be
overlain by sediments of either the Unayzah B member (wells 7 and 13) or the Unayzah A member (well 11). The key to all symbols
used in core logs in this figure and all other relevant figures throughout the paper is presented in Appendix Figure A1.
Figure 4. Characteristics of glaciotectonic shear zones within the Unayzah C member. (A) Core photograph illus-
trating a typical shear zone within the Unayzah C member in well 7 (cf. Figure 3A, 0–10 ft). This occurs between
14,108 ft and 14,115 ft, and displays evidence for both brittle and ductile shear. It is overlain by intraformational
conglomerate of mud clasts and laminated sandstones (14,104.5 ft to 14,106.5 ft), followed by undeformed mas-
sive sandstones from 14,104.5 ft to the top of the core. (B) Gamma-ray log through the Unayzah C member in well
14, showing how spiky, high API values are associated with shear deformation in core. The core examples B(i) and
B(ii) are arrowed on the gamma-ray log. The otherwise monotonous, low API gamma trace reflects the relatively
featureless rock that characterizes the Unayzah C member away from the shear zones (example at B[iii] is also
arrowed on the wireline log).
Late Paleozoic Gondwanan glaciation in Saudi Arabia 45

Post-Depositional Characteristics sandstones and conglomerates analogously represent extensive


glaciofluvial outwash deposits associated with retreat phases of
In many of the wells that have penetrated and cored the the late Paleozoic Gondwanan ice sheets in Saudi Arabia.
Unayzah C member (e.g., wells 6, 7, 8, 11, 13, 14, 17, and The occurrence of an extensive line of end-moraine com-
24), distinctive zones of intense subhorizontal shear have been plexes known as the Rehburg Line (Fig. 6) that is associated
observed (Melvin and Sprague, 2006). These shear zones occur with the north European Pleistocene glacial outwash deposits is
with associated low-angle overfolds and range in thickness from of special significance with regard to the Unayzah C member.
2 ft (0.6 m) to >20 ft (6 m). Examples are illustrated in Fig- This line extends >500 km from the North Sea in the west to
ures 3A, 3B, and 4. The shear zones display a complex variety of Hanover, Germany, in the east (Bennett, 2001) and marks the
features ranging from brittle shear to ductile shear to simple soft- approximate extent of glacier ice during the Rehburg Phase
sediment deformation. These zones are characteristically over- (Drenthe advance) of the Saalian Glaciation (Van der Wateren,
lain and underlain by much thicker intervals of sandstone that 1995). Several push moraine complexes were documented along
are devoid of any such deformation (Figs. 3A and 4B). The shear the Rehburg Line (Van der Wateren, 1985, 1987, 1994), wherein
zones are highly distinctive wherever they are seen in core, and their architecture consists of a number of subhorizontal nappes
all these zones recorded to date clearly correlate with intervals that have been displaced horizontally by the ice, in some places
that display a spiky, high gamma profile on wireline logs (e.g., as much as 6 km. These nappes are bounded above and below
Fig. 4B). Commonly, they are also associated with a strongly by a number of shear zones (Bennett, 2001). The description of
chaotic signature on image logs (M.H. Prudden, 2005, personal the Unayzah C member presented above has highlighted the rec-
commun.). This has enabled tentative identification of the shear ognition of well-developed zones that show a range of deforma-
zones in uncored wells: in such cases the chaotic zones are sepa- tion styles from brittle to ductile shear and between which the
rated by considerable thicknesses (several tens of feet) of appar- rock is essentially undeformed. Furthermore, recent investiga-
ently normally stratified sediment, as has been observed in the tions by the senior author of the Carboniferous-Permian lower
cored examples. Juwayl Formation (Unayzah C equivalent) at the outcrop in the
Wajid region of southwest Saudi Arabia also revealed the pres-
Origin and Evolution of the Unayzah C Member ence of discrete zones of low-angle dislocation (including shear
and overfolding) within sandstones that are interpreted as glacial
The character of the multistory bedsets of the Unayzah C outwash within glacial paleovalleys. (That outcrop work will be
member suggests that it was deposited in an environment that presented in detail in a future publication.) All of this evidence,
was dominated by processes of repeated erosion and deposition. considered in conjunction with the earlier conclusion that the top
The multiple scoured contacts imply a strongly channelized envi- of the Unayzah C member represents an unconformity of some
ronment, and the grain size and primary sedimentary structures magnitude, suggests the following model for the evolution of the
are suggestive of high energy conditions, as indeed are the thick Unayzah C member.
beds with abundant dispersed mud clasts. The diffuse fabric seen Following the mid-Carboniferous tectonic event and the
in many of these sandstones is ascribed to dewatering, and the near-coincident inception of the late Paleozoic Gondwanan South
associated argillaceous crinkly laminations probably represent Polar glaciation, the ice ultimately extended across the southern
elutriated fines associated with that dewatering. The deposi- part of the Arabian plate to the vicinity of wells 13 and 14 in this
tional system of the Unayzah C member was thus dominated by study. That is to say, it spread to paleolatitudes at least as far north
an abundance of channels characterized by rapid, high energy, as the present-day location of the Al Batin Arch (see Fig. 1A)
sandy bedload sedimentation—i.e., an extensive alluvial braided and therefore significantly farther north than had been previously
plain. The thin siltstone beds represent rare low-stage suspension documented (e.g., Stampfli and Borel, 2004). This glacial event
fallout deposits of fine-grained sediment upon the upper surfaces was long-lived and possibly persisted for some 30–45 m.y., from
of the channel braided bars. the mid-Carboniferous (late Visean to early Namurian, or Ser-
As has been discussed above, limited biostratigraphic evi- pukhovian) to Early Permian (Sakmarian) times (Al-Husseini,
dence suggests that the Unayzah C member is Late Carbonif- 2004). Considerable uncertainty exists, however, regarding the
erous (Stephanian = Moscovian-Gzhelian) in age (Stephenson specific timing of the onset of this glaciation in Arabia, not least
et al., 2003). This indicates that it was laid down relatively early because of the possibility of a number of separate phases of gla-
following formation of the Hercynian unconformity, and, further, cial advance and retreat within its overall duration. Melvin and
that deposition commenced during an early stage of the late Paleo- Sprague (2006) and Osterloff et al. (2004a) discussed the likeli-
zoic Gondwanan glaciation. It is clear that these fluvial sediments hood that older glaciogenic deposits may very well have been suc-
have a widespread distribution throughout the subsurface across cessively removed as a result of extensive cannibalization during
the study area (Fig. 5). Similar extensive fluvial sands and grav- later glacial advances in Saudi Arabia and Oman, respectively.
els are identified within postglacial outwash deposits associated In the Unayzah C member in the subsurface of eastern-central
with the several phases of retreat of the Pleistocene ice sheets in Saudi Arabia, each of the inferred glacial retreat phases resulted
northern Europe. It is inferred that the Unayzah C coarse-grained in substantial volumes of glaciofluvial outwash sands and gravels
Figure 5. Stratigraphic section across the study area, showing thickness variations and lateral extent of the Unayzah C member, in relation to the
Unayzah B member and the unnamed middle Unayzah member (“UmUm”). Datum is the top of the unnamed middle Unayzah member. Lower
bounding surface is the Hercynian unconformity (HU). Note the extent of downcutting into the Unayzah B member by the pre-Khuff unconfor-
mity (PKU) in wells 2 and 3. Map shows line of section.

Figure 6. Map of the northwest European


Plain, showing the Rehburg Line of push
moraines. These moraines are associated
with phases of glacial advance of the
Pleistocene Fenno-Scandian ice sheet.
Modified from Bennett (2001).
Late Paleozoic Gondwanan glaciation in Saudi Arabia 47

being deposited across the area upon laterally extensive alluvial 1996; Ruegg, 1983), as well as from Proterozoic glaciogenic sedi-
braided plains. During subsequent readvances the ice sheets pro- ments from Greenland by Moncrieff and Hambrey (1990).
gressed across the glaciofluvial outwash of each preceding retreat
phase. In the process they formed a number of glacially induced, Stratabound, Internally Deformed Deposits
push moraine nappes, similar to those documented by Van der (Push Moraines)
Wateren (1985, 1987, 1994) along the Rehburg Line from the
Pleistocene of the north European Plain (Fig. 6). Ultimately, this This facies is well displayed in two zones in well 7 (Fig. 7,
process created a thick pile of subhorizontal, superimposed units 26.5–42.0 ft and 55–68 ft). At each location the rock displays
of glaciofluvial sands and gravels, separated by distinct shear severe structural deformation in the form of high- to low-angle
zones (Melvin and Sprague, 2006). It is likely that these repeated reverse faulting (listric thrusting) (Fig. 8A) and overfolding.
processes of glacial thrusting associated with readvances of the The rocks comprise sandstones and silty mudrocks that contain
ice sheet were also responsible for the remobilization of pore flu- dispersed pebbles and small cobbles of siltstone and very fine–
ids within the outwash sands, thus giving rise to the widespread grained sandstone (diamictite) (Fig. 8B). Palynological analyses
evidence of dewatering that is observed within these rocks. The of these fine-grained deposits have yielded a heavily reworked
top Unayzah C unconformity is considered to represent the sub- assemblage of palynomorphs of Silurian, Devonian, and Early
glacial surface at the time of the final advance of the late Paleo- Carboniferous age, as well as significant quantities of monosac-
zoic Gondwanan ice sheet in Saudi Arabia. cate pollen that cannot be older than Namurian (Serpukhovian) in
age (the “Cm palynoflora assemblage”: J. Filatoff, 2004, written
UNAYZAH B MEMBER commun.; Fig. 7).
These intervals of deformed rock are stratabound by unde-
The preceding discussion has shown how the Unayzah C formed sediments laid down in a conformable succession (Fig. 7)
member is characterized across the study area by a limited number and are thus interpreted as having been disrupted at, or very shortly
of depositional facies, namely multistory quartzose sandstones after, their time of deposition. Furthermore, it can be shown (see
and rare conglomerates that were deposited upon an extensive below) that they are interstratified with sediments that are demon-
glaciofluvial outwash braided plain. Consequently, that member strably glaciogenic in nature. They have thus been interpreted as
displays a fairly distinctive, monotonous wireline log signature. the product of glacial deformation and specifically identified as
This is in stark contrast to the rocks that make up the Unayzah B the preserved remnants of glacial push moraines (Melvin and
member. Melvin and Sprague (2006) described in detail >1355 ft Sprague, 2006).
(406.5 m) of core from the Unayzah B member in 13 wells, Bennett (2001) described how push moraines display a wide
and identified 8 distinctive depositional facies from within that range of morphologies at a range of scales, from a few meters to
stratigraphic unit. The characteristics of those various facies are features that extend for several kilometers. The overthrust out-
presented below. In well 7, continuous core was recovered, not wash sands and gravels described earlier from the Unayzah C
only from the entire Unayzah B member as it is represented in member fall into the latter category. The stratabound, internally
that well but also from the complete section through the overly- deformed deposits in the Unayzah B member are much smaller
ing “un-named middle Unayzah member” (Melvin and Sprague, features, with an average thickness of ~14 ft (4.2 m). Boulton
2006), as well as the uppermost parts of the underlying Unayzah et al. (1999) identified four broad categories of push moraines,
C member. This cored interval thus represents a crucial record of wherein the style of deformation involves either fans of imbri-
the stratigraphic relationships in the Unayzah in that well, and is cate thrusts or superimposed subhorizontal overthrusted nappes.
reproduced as a reference section in Figure 7. Regarding the former, the smallest are generally no more than
16.4 ft (5 m) high and can be found in both terrestrial and sub-
Sediment Fracture-Fill Features (?Periglacial aqueous environments (Boulton, 1986; Boulton et al., 1999;
Deformation) Bennett, 2001). These small push moraines are more representa-
tive of the stratabound, internally deformed deposits seen in the
In well 4, recovered cores show that the Unayzah B member Unayzah B member in well 7.
rests directly upon the Hercynian unconformity (i.e., the Unayzah In the western part of the study area the Unayzah B member
C member is absent at that location). A laminated sandstone 3 ft in wells 2 and 3 (see Fig. 1A) directly overlies rocks of early
(0.9 m) thick is present ~2 ft (0.6 m) above the unconformity. Paleozoic age, i.e., it rests directly upon the Hercynian unconfor-
It displays a sharp, subvertical fracture in core that is filled with mity. It is 40 ft (12 m) thick in well 2, and 65 ft (19.5 m) thick in
a variety of coarse-grained detritus including mud clasts and well 3, and in each well it is characterized throughout by poorly
granule-sized quartz grains. This feature has been discussed and sorted sediment and by an abundance of small-scale faults and
tentatively interpreted by Melvin and Sprague (2006) as a frost low-angle shear planes (Melvin and Sprague, 2006). This perva-
contraction wedge formed in a periglacial setting. Similar features sive intraformational deformation of the Unayzah B member in
have been documented in modern periglacial settings (French, these wells is also attributed to direct sustained contact with ice.
Figure 7. Core log from well 7, showing the complete section through the unnamed middle Unayzah member (“UmUm”) and the
Unayzah B member in that well, and their respective contacts with the Unayzah A member above and the Unayzah C member
beneath. Note: Cm indicates the locations of the proven occurrence of the heavily reworked “Cm palynofloral assemblage” in the
Unayzah B member (J. Filatoff, 2004, written commun.).
Late Paleozoic Gondwanan glaciation in Saudi Arabia 49

Figure 8. Core photographs from well 7, showing


features observed in the glacially tectonized facies
of the Unayzah B member. (A) Interval of sandstone
exhibiting high-angle, reverse faulted (thrusted) dis-
locations (see Fig. 7, 58–61 ft). (B) Poorly sorted
pebbly sandstone (diamictite), displaying a strongly
sheared fabric (see Fig. 7, 48–51 ft).

A B

Massive, Very Poorly Sorted Pebbly Siltstones (Diamictite) host rock of gray-green (and locally red-brown) siltstone to very
fine–grained sandstone within which occurs an abundance of
Diamictites are poorly sorted sediments containing a wide well-rounded to subangular, dispersed grains of medium to very
range of particle sizes in a relatively fine matrix (Bates and Jack- coarse quartz sand as well as up to 5% coarser material includ-
son, 1980). The massive diamictites described herein are com- ing granules and pebbles of granite, quartz, feldspar, black chert,
monly observed within the Unayzah B member in the Saudi gray siltstone, sandstone, and mudstone (Fig. 10A). Cobble-
Arabian subsurface, being recorded in this study from wells 1, 2, sized clasts of fine-grained sandstone have been observed in
3, 4, 5, 7, 10, and 12. In some cases they are severely deformed places (Fig. 10B).
(e.g., wells 2, 3, and 7: see previous discussion) as a result of Several workers investigating diamictitic deposits in glacial
being incorporated in push moraines. Where they have been settings have identified a number of subfacies, including lodg-
identified in core they range in thickness from ~6 ft (1.8 m) in ment tillites, proximal and distal water-lain tillites, and debris
well 10 to 14 ft (4.2 m) in well 12, to >200 ft (60 m) in well flows (e.g., Visser, 1982; Levell et al., 1988; Moncrieff and Ham-
5 (Fig. 9). In general the diamictites display almost no stratal brey, 1990). The latter authors conceded that “there are numerous
fabric and are extremely poorly sorted. Thus they comprise a cases in which the interpretation remains open to question.” We
Figure 9. Core logs from wells 10, 12, and 5, showing associations of glaciolacustrine depositional
facies in the Unayzah B member. (A) Sediment gravity flows in well 10 (0–25 ft; note the dropstone
at 17 ft), overlain by multiple laminasets of ripple cross-laminated sandstones (25–44 ft) and a thin
diamictite (44–50 ft), followed by mudrock in the uppermost 5 ft. (B) Laminated mudrock (0–7.5 ft)
in well 12, overlain by massive diamictite (7.5–20.0 ft). This is followed by stratified diamictite
(20–28 ft) and sediment gravity flows (28–41 ft) in the upper part. (C) Extremely thick interval of
massive diamictite in well 5 on top of a unit of sediment gravity flows (0–10 ft). Wireline logs in this
well suggest a total >200 ft (60 m) of diamictite.
Late Paleozoic Gondwanan glaciation in Saudi Arabia 51

A B

Figure 10. Core photographs showing features asso-


ciated with massive and stratified diamictites in the
Unayzah B member. (A) Typical example of massive
diamictite, showing the very poorly sorted nature
of the sediment, and the apparently random orien-
tation of clasts (well 12). (B) Cobble-sized clast of
fine-grained sandstone in massive diamictite (well
5). (C) Stratified diamictite in well 7, showing fine-
grained glaciolacustrine laminite with a thin (2 cm)
C D interval of (type 1) very poorly sorted diamictite. Note
the extreme fissility in the lower part of the laminite
subfacies. (D) Cobble-sized clast of diamictitic mate-
rial (till pellet) in type 1 stratified diamictite. Note the
irregular, diffuse edges to the till pellet (well 12).

concur with Melvin and Sprague (2006), who discussed in detail sorted pebbly siltstone material (diamictite) (Fig. 10C). Two
the massive diamictites of the Unayzah B member and concluded types of this diamictite are observed. Type 1 beds are 0.2–4.0 in.
that they most likely represent resedimented glacial debris laid (0.5–10.0 cm) thick and have indistinct upper and lower bound-
down at the bottom of glacial lakes as debris flow deposits. aries. They consist of silty mudstone that contains abundant dis-
persed grains of fine- to very coarse–grained sand, as well as rare
Interstratified Pebbly Siltstones and Laminated granules of quartz, feldspar, and granite. Significantly, they also
Mudstones: Stratified Diamictite commonly contain distinctive pelletoid clasts of material that is
itself very poorly sorted (diamictitic) in nature. These pelletoid
This distinctive facies within the Unayzah B member has clasts range in size from a few millimeters to 15 cm (Fig. 10D),
been recorded from wells 4, 7 (Fig. 7, 68–85 ft), and 12. It is are commonly flattened parallel to bedding, and show diffuse or
heterolithic in character and comprises a host rock of very fine– “ragged” edges. Type 2 beds are 2–12 in. (5–30 cm) thick, with
grained, laminated mudstone (with associated rare, very thin beds sharp, commonly loaded lower bed boundaries, and are similarly
of rippled, very fine–grained sandstone in some places), within very poorly sorted. The most extensive development of this depo-
which occur thin beds (millimeter to decimeter scale) of poorly sitional facies recorded to date occurs in well 7 (Fig. 7, 68–85 ft).
52 Melvin et al.

There, the type 1 beds become increasingly rarer toward the top Multistory Graded Sandstones: Glaciolacustrine Gravity
of the 17 ft interval, where the laminated mudrocks predominate: Flow Deposits
this interval of stratified diamictite effectively fines upward in
this well. This facies is recognized in core from wells 7, 9, 10, 12, 13,
The laminated mudrocks represent the background sedi- and 17. It is most fully developed in well 9, where it is >230 ft
mentation of this depositional facies. To date, they have proved (69 m) thick. There it comprises sandstone beds that are 1–5 ft
to be palynologically barren (N.P. Hooker, 2008, personal com- (0.3–1.5 m) thick, displaying sharp, commonly erosional basal
mun.) and are interpreted as lake-bottom deposits (as opposed to contacts, and in many cases they are clearly graded (Fig. 11).
marine). In the type 1 diamictite beds the diffuse nature of the bed They display an abundance of fluid escape structures, dominated
boundaries is suggestive of rain-out through the water column of by elutriation pillars and dish structures (cf. Lowe and LoPiccolo,
debris derived from icebergs floating on the surface of the water 1974). These sandstones are organized into (at least) two upward-
body, probably during warm season meltout. The identification thinning and -fining bedset packages (Fig. 11). A similar package
of the poorly sorted pelletoid clasts is significant in this regard, as of sharp based, graded sandstones occurs in well 10 (Fig. 9A,
they represent “till pellets” sensu Ovenshine (1970). That author 0–25 ft), and in that well a cobble sized, angular sandstone clast
considered such deposits to identify uniquely the existence of was recognized, “floating” on the top of one of the graded beds
glacier ice in very close proximity to the environment of deposi- (Fig. 9A, 17 ft). Other occurrences of sharp-based graded sand-
tion. The type 2 diamictites, with their sharp bed contacts, more stones in the Unayzah B member have been described in detail
likely represent debris flow deposition on the bottom of glacial from other wells by Melvin and Sprague (2006).
lakes. In well 7 the progressive diminution up-section of type 1 The graded sandstones described above from well 9 (Fig. 11)
diamictite beds in favor of the laminated mudrocks (described display all the characteristics of rapidly deposited, highly fluid-
above) suggests increasingly ice-distal conditions through time ized sediment gravity flows such as have been described by Lowe
and/or progressive deepening of the lake. and LoPiccolo (1974). The organization of the beds into thick,
upward-thinning and upward-fining packages is significant. It
Nonstratified Pebbly Siltstone: Rain-Out Diamictite suggests sustained deposition in an environment within which,
for each package of sediment, the deposits became increasingly
These rocks comprise a subfacies of the stratified diamic- distal with respect to their source. It is proposed that these grav-
tites, described above. In well 8, directly overlying the Unayzah ity flow deposits represent sublacustrine outpouring of sediment
C member, and constituting the entire Unayzah B member in this from melting glacier ice. The two distinct upward-fining pack-
well, is a 2 ft (0.67 m) thick bed of diamictite. It is mud sup- ages of sandstones may represent either the successive retreat of
ported and very poorly sorted, and it displays no internal strati- two different glacial sources (e.g., valley glaciers?), or they may
fication. It contains an abundance of dispersed grains of fine- to reflect an element of glacial readvance and retreat in the location
very coarse–grained sand as well as rarer dispersed granules and of this well. That these sublacustrine gravity flow sandstones in
small pebbles of quartz and sedimentary rock fragments. Melvin the Unayzah B member have a glacial origin is supported by the
and Sprague (2006) illustrated how, at the top of this bed, there noted occurrence of the anomalously large clast described from
occurs a small (1 cm) pebble with its long axis oriented perpen- well 10 (Fig. 9A). This is interpreted as a dropstone, which melted
dicular to bedding. In the lower part of the bed, indistinct mot- out of ice floating on the lake surface, and fell upon the lake-
tling is observed that is suggestive of bioturbation (Melvin and bottom gravity flow beneath. In well 13, dropstones have been
Sprague, 2006). Similar mottled diamictites are seen in well 7, recognized in fine-grained lake deposits (Melvin and Sprague,
within the stratified diamictite facies. 2006) that were subsequently overlain by upward-thinning and
The noted occurrence of the pebble with its long axis ori- -fining packages of gravity flow deposits, similar to the deposits
ented perpendicular to bedding supports an origin for these rocks described herein at well 9. Furthermore, the overall trend of those
related to melt-out from floating ice in a glacial lake. Similar deposits in well 13 was to become thinner and finer upward, sug-
rain-out diamictites of glacial origin were described and dis- gesting that the sediment supply was becoming more distant (i.e.,
cussed from the Karoo in South Africa by Visser (1982), albeit the rocks became more ice-distal in character), and/or the lake
on a larger scale. The possible bioturbation in these Unayzah B was becoming deeper.
sediments is suggestive of a seasonal aspect to their deposition
and supported by its recognition in the stratified diamictites in Ripple-Drift Cross-Laminated Sandstone: Sublacustrine
well 7. Thus, in relatively warm periods, coarse-grained detritus Glacial Outwash
was released from melting ice floes in a glaciolacustrine setting;
having settled upon the lake bottom, the sediment was colonized This depositional facies has been observed only in core
by organisms that were suited to the relatively mild conditions. from well 10, where it occurs in a unit that is ~20 ft (6 m) thick
When harsher conditions prevailed, sediment rain-out was mini- (Fig. 9A, 25–45 ft), resting directly above the interval of sublacus-
mized, faunal activity dwindled, and sedimentation was reduced trine gravity flow sandstones mentioned above. It comprises three
to suspension settling of very fine–grained silt and clay. bedsets, 3–11 ft (0.9–3.3 m) thick, each separated from the other
Late Paleozoic Gondwanan glaciation in Saudi Arabia 53

Figure 11. Wireline gamma-ray (GR) log, showing cored interval


and associated core log from well 9, showing thick development of
multiple beds of highly fluidized sediment gravity flows (as indicated
by abundance of dish structure and elutriation pipes). Note how the
beds are arranged in thick packages that show a general upward thin-
ning and fining trend (see text for full discussion). These gravity flow
sandstones sit directly upon an interval (uncored) with a high gamma-
ray-log signature, which has yielded samples of the heavily reworked
Cm palynoflora (J. Filatoff, 2004, written commun.).

by a thin interval of siltstone. Each bedset consists of fine- to very


fine–grained sandstone characterized by multiple, thin (1–3 cm)
laminasets of climbing ripple-drift cross-lamination. Type A
ripple-drift cross-lamination of Jopling and Walker (1968) domi-
nates the succession, although it is seen to pass upward into type
B ripple-drift cross-lamination in the uppermost 3 ft (0.9 m).
The sandstones that display type A ripple-drift cross-
lamination have been interpreted to represent sustained bedload
transport of sediment that originated as subaqueous outwash
material in a glacial lake (Melvin and Sprague, 2006). Analogous
sublacustrine glacial outwash deposits, dominated by laminasets
of ripple-drift cross-laminated sandstones, have been described
from the Pleistocene of Canada, where they are associated
with lake-bottom kame deltas (Gustavson et al., 1975) as well
as subaqueous esker fan deposits (Rust and Romanelli, 1975).
The uppermost sets of type B ripple-drift cross-lamination in
well 10 in the present study suggest decreasing energy, whereby
increased fallout from suspended load prevailed over bedload
transport. This inferred loss of energy is taken to indicate that the
original source of sediment (melting ice) had become somewhat
farther removed from the location of well 10, i.e., this succession
of ripple-drift cross-laminated sandstones represents an increas-
ingly ice-distal facies association.

Mudrock Facies: Distal Glaciolacustrine Deposits

Fine-grained sediments of this depositional facies have been


observed in a number of wells (e.g., wells 7, 10, 12, 13, and 17)
(Melvin and Sprague, 2006). In well 10 (Fig. 9A, 50.5–69.5 ft),
19 ft (5.7 m) of dark gray mudrock overlies a 6 ft (1.8 m) thick
diamictite. The lowermost 10 ft (3 m) of this mudrock deposit
comprises a series of stacked, thin (centimeter scale) muddy
siltstone beds that are sharp based and are distinguished from
each other only by very thin (millimeter scale) claystone part-
ings. This lower unit fines upward gradationally into increasingly
argillaceous, homogeneous dark gray mudrock; in the uppermost
0.13 ft (4 cm) it becomes a pale gray claystone.
The muddy siltstones of the lower part of this mud-prone
interval display the characteristics of very fine–grained gravity
flow deposits (“distal turbidites”) and pass upward into essentially
featureless mudstone, interpreted as suspension fallout deposits.
They have been proven to be palynologically barren (J. Filatoff,
2004, written commun.) and are considered to have been laid
54 Melvin et al.

down on the bottom of a lake. The overall upward-fining of these sage from type A to type B ripple-drift cross-lamination (sensu
muddy sediments from siltstone to claystone is considered to rep- Jopling and Walker, 1968) suggests increasingly ice-distal condi-
resent a gradual deepening of the lake through time. tions. They are directly overlain by a massive diamictite deposit
that is only ~6 ft (1.8 m) thick. Its limited thickness suggests
Origin and Evolution of the Unayzah B Member an intermediate setting in terms of its proximity to any glacial
source (especially when compared with the very great thickness
The foregoing discussion has highlighted the large number of the same facies at well 5). The uppermost 19 ft (5.7 m) of the
of depositional facies that characterize the Unayzah B member. Unayzah B succession in well 10 comprises ~10 ft (3 m) of mud-
Except for the inferred periglacial frost wedging described ear- prone distal turbidites that pass up into 9 ft (2.7 m) of featureless
lier, and the case of wells 2 and 3 in the western part of the study mudrock (Fig. 9A, 50.5–69.5 ft). These fine-grained sediments
area (to be discussed below), each of those facies is interpreted to were laid down in an extremely ice-distal setting. This sedimen-
have been deposited in a lacustrine environment (palynological tary succession at well 10 thus represents a passage through time
evidence for a marine setting being consistently lacking). In most from a relatively ice-proximal setting to one that was distinctly
cases there is evidence, direct or indirect, to suggest a glacial ice-distal. The clear inference from this evidence is that, relative
influence upon that environment. Thus, the stratabound, inter- to the location of well 10, either the ice was becoming increas-
nally deformed deposits in well 7 have been interpreted above ingly distant (i.e., retreating) or the lake was becoming deeper,
to represent the remains of glacial push moraines and as such or both. It was noted above how Melvin and Sprague (2006)
provide strong evidence of an ice-contact setting. The remain- observed similarly that in well 13, a series of upward-thinning
ing depositional facies are indicative of a spectrum of glacially and fining packages of gravity flow deposits in the Unayzah B
related environments, ranging from ice-proximal to ice-distal. member displayed an overall trend toward being thinner and finer
Considering the volumes of sediment they represent, the >200-ft- upward, again suggesting that the sediment supply (from the ice)
thick, massive, very poorly sorted pebbly siltstones (diamictites) was becoming more distant or the lake was becoming deeper.
at well 5 (Fig. 9C), and the >230-ft-thick sequence of multistory In well 7 the sediments of the Unayzah B member that
graded sandstones (glaciolacustrine gravity flow deposits) in well directly overlie the top Unayzah C unconformity consist of ~23 ft
9 (Fig. 11) most likely represent deposition in an ice-proximal (6.9 m) of sublacustrine gravity flow deposits (Fig. 7, 3–26.5 ft).
sublacustrine environment. In the latter case the organization These are interpreted as relatively ice-proximal, sublacustrine fan
of the gravity flow deposits may suggest either multiple glacial deposits. They are overlain (Fig. 7, 26.5–42.0 ft) by the lower
sources or possibly some degree of glacial readvance, as has been of two units of stratabound, internally deformed deposits (push
discussed earlier. Ice-distal settings are suggested by the very moraines), inferred to represent ice-contact conditions. The suc-
fine–grained sediment (mudrock facies) seen in well 10 as well cession continues upward with ~13 ft (3.9 m) of massive, very
as by the interstratified pebbly siltstones and laminated mud- poorly sorted pebbly siltstones (diamictite) (Fig. 7, 42.0–55.0 ft),
stones (stratified diamictite) that are developed in wells 7 and and that is again interpreted to be a relatively ice-proximal
12. Sediments indicative of deposition in settings intermediate deposit. The return of an ice-contact setting is seen in the second
in the “proximal-distal spectrum” are represented by nonstrati- of the stratabound, internally deformed deposits (push moraines)
fied pebbly siltstone (rain-out diamictite) and ripple-drift cross- (Fig. 7, 55–68 ft). The highest part of the Unayzah B member
laminated sandstone (sublacustrine glacial outwash), as well as in well 7 (Fig. 7, 68–85 ft) is characterized by well-developed
the thinner bedded occurrences of massive, very poorly sorted and sustained deposition of interstratified pebbly siltstones and
pebbly siltstones (diamictite) and multistory graded sandstones laminated mudstones (stratified diamictite). Those mud-prone
(glaciolacustrine gravity flow deposits). sediments represent a significantly more ice-distal setting than
The assignment of these various depositional facies to a any of the underlying sediments. Furthermore, the upward dim-
position in the “ice proximal-distal spectrum” is substantiated, inution of the amounts of glacially derived sand grains in this
not only by their grain size, thickness, and general sedimento- facies in this well (described earlier) does itself strongly imply
logical characteristics, but also by their associations relative to continued retreat of the ice from this location. Thus, in well 7
each other from well to well. That is to say, the vertical associa- the Unayzah B member displays evidence for an overall gradual
tion of depositional facies commonly suggests an overall passage retreat of the ice whereby ice-proximal conditions gave way to
in any given location from an ice-proximal setting to one that was ice-distal conditions, but during which at least two distinct ice-
ice-distal. This is well displayed in well 10 (Fig. 9A). There, a contact deformational events occurred. That is to say, the overall
lowermost interval of at least 25 ft (7.5 m) of medium- to coarse- retreat of the ice was interrupted by at least two minor readvances
grained multistory graded sandstones (glaciolacustrine gravity of the ice sheet in this location. The possibility of minor glacial
flow deposits) (with dropstones) represents ice-proximal subla- readvance within the Unayzah B member was also inferred ear-
custrine fan deposition. This is overlain by ~20 ft (6 m) of multi- lier from the architecture of the gravity flow sandstones described
ple laminasets of ripple-drift cross-laminated sandstone (Fig. 9A, from well 9.
25–45 ft). These sediments were interpreted earlier also to be sub- It is clear from the foregoing examples from individual
lacustrine glacial outwash detritus, within which the upward pas- wells that across most of the study area the vertical association of
Late Paleozoic Gondwanan glaciation in Saudi Arabia 55

depositional facies in the Unayzah B member generally displays Locally minor readvances of the ice occurred, with the creation
an increasingly ice-distal aspect up-section. In all cases this is of minor push moraine deposits. In general, however, the terminal
interpreted as representing the steady retreat of the Gondwanan retreat of the ice was reflected and preserved in the sedimentary
ice sheet in early Permian times. Locally, there is evidence for record in the form of increasingly ice-distal deposits. The land-
some minor readvance of the ice (e.g., at wells 7 and 9). Support scape changed from one of numerous isolated and deep glacial
for these sedimentological conclusions can be found in the paly- lakes to one that was awash with glacial meltwaters as the lakes
nology. The Late Carboniferous–Early Permian assemblages that overspilled their boundaries and became connected. This termi-
characterize the OSPZ2 palynozone of Stephenson et al. (2003) nal (maximum) flood scenario is preserved at well 8, where the
(which in turn characterizes the Unayzah B member in Saudi thin rain-out diamictite directly overlies the top of the Unayzah
Arabia) comprise a variety of spores, pollen, and fresh-water C unconformity and appears to link the much deeper glacial lake
algae (N.P. Hooker, 2008, written commun.). These assemblages deposits that characterize wells 7 and 9 (Fig. 12).
can be interpreted as showing a progression from a cold, dry cli- In the western part of the study area, around wells 2 and 3
mate in the Late Carboniferous to a slightly warmer and wetter (Fig. 12) a somewhat different situation prevailed. It was noted
climate in the Early Permian. The latter represents the transition earlier how in each of these two wells the Unayzah B member is
from a glacial to a deglacial phase, and is reflected in the palyno- represented in its entirety by relatively thick intervals of glacially
flora by the change from low diversity spores and monosaccate deformed sediment, which are attributed to sustained, direct con-
pollen-dominated assemblages (glacial phase) to high diversity tact with the ice. Furthermore, it is clear from the lack of relevant
spores with fresh-water algae (deglacial phase) (N.P. Hooker, depositional facies that when the ice did eventually melt in this
2008, written commun.). The palynofloral assemblages of the area, the meltwaters were not ponded in topographic lows to form
deglacial phase are commonly derived from the massive dia- significant lakes such as has been described widely from across
mictites described above as being relatively common within the the region. The possible corollary exists therefore that the western
Unayzah B member. part of the study area may have been topographically high during
Figure 12 is a stratigraphic section that is hung using the Unayzah B times. This is a not altogether unlikely scenario, given
top of the “un-named middle Unayzah member” (Melvin and the proximity of these westerly locations to the Al Batin Arch
Sprague, 2006) as a datum. It incorporates both the Unayzah of Faqira et al. (2009) (see Fig. 1A). Indeed, given the sugges-
B member and the “un-named middle Unayzah member.” It is tion by some workers (e.g., Eyles, 1993) that ice coverage may
clear that the Unayzah B member exhibits considerable thick- have extended over to present-day east Africa, it is interesting to
ness variation across the study area. Specifically the thickest sec- speculate upon the possibility of a center of high altitude (alpine)
tions of this stratigraphic unit are seen in wells 5 and 9: it has glaciation in western central Saudi Arabia at this time.
been discussed above how these two wells contain sediments that
are the most ice-proximal in their character. The thinnest occur- UNNAMED MIDDLE UNAYZAH MEMBER
rence of the Unayzah B member is at well 8 (Fig. 12), where it is
represented by only 2 ft (0.6 m) of nonstratified pebbly siltstone The unnamed middle Unayzah member was recognized and
(rain-out diamictite) resting directly upon the top of the Unayzah defined in the extensive core-based study of the lower Unayzah
C unconformity. Considered in the context of the preceding dis- Formation carried out by Melvin and Sprague (2006). Its facies
cussion, these stratigraphic relationships suggest the following characteristics show that it is very different from the glaciogenic
model for the evolution of the Unayzah B member. Unayzah B member, and yet it displays a distinctive character
Following the various retreats and inferred readvances of the that sets it aside also from the Unayzah A member. It is a stratal
Gondwanan ice sheets that led to the deposition and deforma- unit that represents a transition from the Unayzah B member to
tion, respectively, of the Unayzah C member (see earlier discus- the Unayzah A member, even though its boundaries are sharp
sion), the subglacial surface of the final readvance was effectively and readily identified in core. Among the wells examined in this
preserved as the top of the Unayzah C unconformity. Thereafter, study, the unnamed middle Unayzah member occurs in a num-
the final melting and ultimate withdrawal of the ice took place. ber of situations. Thus it may be identified directly overlying the
As the ice melted, the subglacial topography was exposed, and Unayzah C member, as at well 6 (Fig. 12), although more com-
the lows began to be filled with the meltwaters from the retreat- monly it sits upon the glaciogenic facies of the Unayzah B mem-
ing ice sheets, with the consequent formation of abundant glacial ber. In places it is absent, as at well 11 (Fig. 12), and elsewhere
lakes. Initially the deepest of those lakes were infilled with large it may have been removed by erosion at the pre-Khuff unconfor-
volumes of resedimented glacial outwash material in the form mity, as seen in wells 2 and 3 (Fig. 12). In well 7 the unnamed
of massive, very poorly sorted ice-proximal debris flows (dia- middle Unayzah member appears to have been completely cored,
mictites) (e.g., well 5, Fig. 9C) and highly fluidized gravity flow and its relationships to the Unayzah B and A members are clearly
sands that developed significant lake bottom turbidite deposits exposed (Fig. 7). This well thus serves as a valuable reference
(e.g., well 9, Fig. 12). As the ice continued to retreat, more melt- well. The nature of the individual depositional facies and the
water and more sediment were supplied, and the numerous gla- stratigraphic boundaries of the unnamed middle Unayzah mem-
cial hollows were filled and spilled over with rising floodwaters. ber are discussed below.
56 Melvin et al.

Figure 12. Stratigraphic section across the study area, showing facies architecture within the Unayzah B member and the overlying unnamed
middle Unayzah member. Datum is the top of the unnamed middle Unayzah member. In wells where this sediment package is thick (e.g., wells
7, 9), the lowermost parts of the section are dominated by glaciogenic deposits (Unayzah B member). These deposits are overlain in those wells
by nonglacial, red lacustrine siltstones and associated sandstones of the floodplain facies of the unnamed middle Unayzah member. Where the
overall package is thin (e.g., well 6), it comprises only the latter member. Note the effects of the pre-Khuff unconformity (PKU) in wells 2 and
3. Map shows line of section. Modified from Melvin and Sprague (2006). GR—gamma-ray.

Fluvial Sandstones grained, ripple cross-laminated sandstones (Fig. 13A). These


characteristics of the sandstones, and their occurrence embedded
The best example of fluvial sandstones in the unnamed within red very fine–grained (silty) deposits, suggest deposi-
middle Unayzah member is seen in core from well 8. There, tion in a channelized, heavily oxidized (i.e., terrestrial) setting.
Melvin and Sprague (2006) identified a sandstone interval that They are thus interpreted to be fluvial in origin, and their archi-
is 30.5 ft (9.15 m) thick (Fig. 13A, 29.5–60.0 ft). It directly over- tecture of stacked shallow channel deposits encased and isolated
lies some 10 ft (3 m) of red sandy siltstones (alluvial floodplain within finer–grained floodplain deposits further suggests depo-
deposits: see below) and occurs ~18 ft (5.4 m) above the top of sition in relatively high-sinuosity or even anastomosing rivers.
the very thin Unayzah B member in this well. It is also over- Fluvial channel sandstone deposits, isolated within fine-grained
lain by a thick interval of 45 ft (13.5 m) of red sandy siltstones floodplain siltstones, were described from the Cutler Formation
of the alluvial floodplain facies (see below). Fine- to medium- (Permian to Pennsylvanian) of New Mexico by Eberth and Miall
grained, moderately sorted sandstones occur in stacked beds that (1991), and were similarly interpreted to be anastomosed river
are 2–4 ft (0.6–1.2 m) thick. They display sharp, erosional con- deposits. In well 7 there occurs a similar interval of sandstones
tacts, with local mud-clast concentrations, and pass upward into that shows cross-bedding and an upward-fining profile; these are
low-angle trough cross-lamination overlain in places by finer– also considered to be of fluvial origin (Fig. 13B, 41–50 ft).
Late Paleozoic Gondwanan glaciation in Saudi Arabia 57

Figure 13. Core logs from wells 8, 7, and 6, illustrating the nature of the unnamed middle Unayzah member (“UmUm”). Note in par-
ticular the distinct sandstone units overlying siltstone intervals, and the similarities and differences displayed among them. See text
for detailed discussion.

Eolian Sandstones dark gray siltstone are also present. Overlying these sandstones
in well 6 is an interval 5 ft (1.5 m) thick comprising well-sorted,
This facies is well displayed in the unnamed middle Unayzah well-rounded, and frosted grains that occur in high-angle, grain-
member in wells 7 and 6 (Fig. 13B and 13C). In well 7, there size-segregated cross-laminations (Fig. 13C, 40–45 ft). The
is an interval of ~17 ft (5.1 m) of low-angle to flat-laminated, cross-laminations are disrupted by small-scale synsedimentary
well-sorted fine-grained sandstones wherein the grains are well faults (Fig. 13C, 42–43 ft). Synsedimentary deformation in
rounded and appear frosted (Fig. 13B, 53–70 ft). In well 6, tex- similar facies in the unnamed middle Unayzah member has been
turally similar flat-laminated sandstones are interbedded with observed in several other places. It is particularly well developed
more poorly sorted sandstones, displaying well-developed adhe- in western parts of the study area in central Saudi Arabia, in the
sion ripples (Fig. 13C, 23–40 ft). Thin (centimeter scale) beds of vicinity of well 24 (see Fig. 1A).
58 Melvin et al.

Melvin and Sprague (2006) described how the base of the beds showing adhesion ripples represent damper, sandy inter-
unnamed middle Unayzah member in well 7 features a thin dune conditions, and the thin siltstone interbeds suggest short-
(2 ft; 0.6 m) red sandstone that rests abruptly upon the dark gray lived, shallow bodies of water in an otherwise eolian dominated
stratified diamictite at the top of the Unayzah B member in that environment. The overlying cross-laminated sandstone interval
well (Fig. 13B, 4–6 ft). That sandstone has high-angle cross- in well 6 is interpreted to be a residual eolian dune deposit. The
laminations that become quite diffuse in the lower levels of the synsedimentary dislocations may represent gravitational collapse
bed. It also displays local nested concentrations of subrounded on the dune slip face. Alternatively, Melvin and Sprague (2006)
to subangular, granule-sized grains. It has a sharp, granule- and suggested the possibility that these dislocations may relate to
pebble-strewn basal contact with the underlying glaciolacustrine sediment readjustment in response to melting of bodies of ice or
stratified diamictites of the Unayzah B member. The latter rocks snow that may have been trapped during dune migration. Analo-
display considerable cracking in their uppermost part, and the gous features were described from modern cold-climate eolian-
red sandstone is seen to penetrate deeply into those dislocations ites from the Rocky Mountains of Colorado by Ahlbrandt and
(Fig. 14A). Andrews (1978) and from coeval Permian deposits of the Gond-
The flat-laminated sandstones described above from well wanan sequence of Australia (Williams et al., 1985).
6 and well 7 have the character of eolian sand sheet deposits, Ostensibly, the thin red sandstone that occurs at the base
such as were well documented by Fryberger et al. (1979). The of the unnamed middle Unayzah member in well 7 (Fig. 13B,

A B

Figure 14. Core photographs showing the nature of


the lower and upper contacts of the unnamed middle
Unayzah member in well 7. (A) Gray, silty glacio-
lacustrine mudstone at the top of the Unayzah B
member, abruptly overlain by red eolian sandstones
at the base of the unnamed middle Unayzah member
(“UmUm”) (see Fig. 13B, 5 ft). At the contact, note
how the glaciolacustrine deposits are cracked, and
the coarse basal sediment of the red sandstones has
deeply invaded the cracks. (B) Red siltstone and silty,
very fine–grained sandstone. The mottled and rooted
lower interval is the uppermost part of the unnamed
middle Unayzah member (“UmUm”) in this well,
and represents a paleosol (Fig. 13B, 72.5–77 ft). The
abruptly supradjacent siltstones are assigned to the
Unayzah A member.
Late Paleozoic Gondwanan glaciation in Saudi Arabia 59

4–6 ft) displays the character of a residual, basal eolian dune places these sand grains appear to be concentrated within subver-
sandstone deposit. The juxtaposition of such terrestrial sediment tical cracks within the sediment. This interval in well 6 displays
relative to the underlying glaciolacustrine stratified diamictites characteristic subspherical fractures throughout. Similar features
has significant stratigraphic implications that are discussed fur- have also been noted at the top of the unnamed middle Unayzah
ther below. member in well 9.
These variegated, texturally immature rocks are interpreted
Red Sandy Siltstones: Alluvial Floodplain Deposits as paleosols. In most places where they are recognized, they rep-
resent the stratigraphically highest occurring facies within the
The unnamed middle Unayzah member is commonly char- unnamed middle Unayzah member.
acterized by extensive deposits of sandy siltstones and silty, very
fine–grained sandstones that are generally red, ranging from red- Origin and Evolution of the Unnamed Middle
dish brown to reddish purple, and locally red-green variegated. Unayzah Member
They are well displayed in core from well 6 (Fig. 13C, 10.5–
22.0 ft) and well 7 (Fig. 13B, 7–33 ft). In well 8, 10 ft (3 m) of red The contact that separates the Unayzah B member from the
siltstones are seen below the fluvial sandstones described above unnamed middle Unayzah member is extremely sharp (Fig. 14A)
(Fig. 15, 19.5–29.5 ft), and a further 47 ft (14.1 m) of red sandy and is recognized across the study area (Fig. 12). Melvin and
siltstones are present above those sandstones (Fig. 15, 60–107 ft). Sprague (2006) interpreted the contact to represent a regionally
In that well the highest parts of the unnamed middle Unayzah disconformable surface. The various facies that characterize the
member show a distinct, albeit subtle upward-coarsening grain uppermost part of the underlying Unayzah B member represent
size trend (Fig. 15, 107–129 ft). Although generally these red widespread lake-dominated sedimentation at the time of maxi-
sandy siltstones display very diffuse bedding characteristics, mum flooding of glacial meltwaters across the landscape in Early
in places they are interbedded with red, silty very fine–grained Permian times, as discussed above. The (possible) cold-climate
sandstones that can display irregular lamination or occur as very eolian, fluvial, and fine-grained alluvial red-bed deposits of the
thin (centimeter scale), current-rippled beds. Rarely, these beds unnamed middle Unayzah member clearly represent a very
have sharp upper contacts, showing evidence of sand-filled desic- different, terrestrially dominated depositional setting and sug-
cation cracks. gest that a highly significant drainage event marked the end of
These fine-grained red beds were laid down under highly Unayzah B deposition.
oxidizing conditions, as suggested by the widespread red color- Martini and Brookfield (1995) described how, in the Pleisto-
ation. They are interpreted to have been deposited in a terrestrial cene Bowmanville Bluffs in Ontario, Canada, a very sharp demar-
environment dominated by low energy conditions. The generally cation occurs between clay-rich glaciolacustrine rhythmites and
diffuse bedding characteristics are indicative of heavily water- overlying sand-prone sediments. They suggested that the wide-
logged sediment, and the local development of ripple lamination spread distribution of this contact and the nature of the overlying
suggests the presence of minor current activity. Thus this facies successions were the result not of erosion but of a sudden drop in
is considered to represent subaqueous deposition within shallow lake level caused by the retreat of a glacier.
floodplain lakes. Locally, these lakes were at times infilled to the Teller et al. (2002) noted that during the last deglaciation
point of exposure and desiccation. In that context, the upward- of the Quaternary period, melting of the Laurentide Ice Sheet
coarsening interval at the top of the unnamed middle Unayzah in North America led to the release of very large volumes of
member in well 8 (Fig. 15, 107–129 ft) is tentatively interpreted stored precipitation to the oceans. There were a number of lakes
as a small floodplain lacustrine delta. along the margin of the Laurentide Ice Sheet, whose confining
ice or sediment dams failed during the deglaciation. Of these,
Paleosols proglacial Lake Agassiz was by far the largest, covering a total
of more than one million square kilometers over its 4000 yr his-
In well 7, the uppermost part of the unnamed middle Unayzah tory (Teller et al., 2002). Frequent changes in lake levels of Lake
member is represented by ~4 ft (1.2 m) of red-purple-gray var- Agassiz have been documented. These changes were abrupt and
iegated argillaceous and silty very fine–grained sandstones often involved the release of several thousand cubic kilometers of
(Fig. 13B, 72–76 ft; Fig. 14B). These sandstones are tightly water. The exact time for each lake drawdown (outburst) can only
silica cemented and poorly sorted, containing an abundance of be estimated, but it is believed that lake levels for most phases
dispersed (floating) grains of medium- to coarse-grained sand. could have been drawn down in a matter of months to a few years
They are characterized by well-developed root traces up to 1 ft (Teller et al., 2002). Similarly, in northern Russia during the last
(30 cm) long, cutans, and other indicators of pedogenic develop- glaciation of the Quaternary Period, the North Polar ice sheets
ment (Fig. 14B). In well 6 the eolian sandstone described earlier expanded and blocked north-flowing rivers such as the Yenissei,
from the unnamed middle Unayzah member is overlain by ~4 ft Ob, and Pechora. As a result, south of the ice sheets a number
(1.2 m) of poorly sorted sediment (Fig. 13C, 46–50 ft) that con- of large ice-dammed lakes formed that were considerably larger
tains abundant floating grains of medium to coarse sand. In many than any lake on Earth today (Mangerud et al., 2004). The final
Figure 15. Core log through ~240 ft (72 m) of core in well 8, extending from the uppermost part of the Unayzah C member to the top
of the Unayzah A member. Note the stratigraphic boundaries and the overall upward-coarsening profile from 132 to 238 ft. Modified
from Melvin et al. (2005).
Late Paleozoic Gondwanan glaciation in Saudi Arabia 61

drainage of the best mapped lake (namely Lake Komi in the work has identified several different depositional facies (sum-
Pechora Lowlands of northern Siberia) was modeled, and it was marized below), and these and their resultant facies associations
concluded that it probably emptied within a few months (Man- have permitted the recognition of three major depositional envi-
gerud et al., 2004). These results dramatically demonstrate the ronments within the Unayzah A member. All of these environ-
likelihood of geologically instantaneous drainage events related ments are strongly indicative of a highly arid continental setting,
to glaciolacustrine settings at times of terminal glacial retreat. and hence are suggestive of a significant and sustained climate
Following the probably dramatic drainage event that is change that became manifest at the beginning of deposition of
inferred to have terminated glaciolacustrine deposition of the the Unayzah A member.
Unayzah B member, the Permian landscape was dominated
for the most part by very low lying alluvial floodplains that Semiarid Ephemeral Lake Deposits
accommodated deposition of the large volumes of fine-grained
(silt-sized) material that now constitutes the greater part of the These rocks characterize the lower intervals of the Unayzah
unnamed middle Unayzah member. A number of river systems A member almost everywhere it occurs. They comprise silty, very
traversed these flood basins, but they were probably relatively fine–grained sandstones with minor sandy siltstones that range
isolated in occurrence, as is indicated from the limited evidence in color from brick red and red brown to buff yellow and pale
in the apparent relative isolation of their resulting channel depos- to dark gray. They display variably lenticular to indistinct crin-
its (Figs. 13A and 15). Toward the end of the period of deposition kly lamination, and locally small-scale ripple cross-lamination
of the unnamed middle Unayzah member, relatively drier condi- is seen. Rarely, over-steepened and folded sediment is observed.
tions became established, accompanied by a reduction in sedi- Elsewhere, small-scale vertical features have been noted and
ment supply, and the fluvial sands were reworked into a number interpreted as syneresis cracks. These very fine–grained rocks are
of isolated eolian deposits (Figs. 13B and 13C). The possibility specifically characterized by the common occurrence of thin,
that these sandstones may represent cold-climate eolianites has flat-lying (horizontal) laminae, commonly only one or two grains
considerable implications for the Permian paleogeography of the thick (Fig. 15, 136–186 ft). The laminae comprise grains of well-
study area, particularly in the western part. There, around well 24 rounded, medium to coarse sand. Commonly, bedding and lami-
(Fig. 1A), these postulated cold-climate eolian deposits are par- nation within these rocks are disrupted by subvertical, sand-filled
ticularly well developed. It was mooted earlier that the possibility cracks that have a downward penetration of 2–5 cm.
exists that a high-altitude Permian glaciation was active and cen- These fine-grained silty sediments dominate the lower parts
tered on the Al Batin Arch (see Fig. 1A). If such was the case the of the Unayzah A member, as was noted above. As such, they
likelihood that these deformed eolian sandstones of the unnamed commonly overlie the similar red siltstones that were described
middle Unayzah member are indeed cold-climate deposits, laid earlier from the unnamed middle Unayzah member, separated
down at a relatively high altitude, is rendered significantly more from them in most places only by the paleosols that mark the
credible. Significantly, ongoing palynological studies from the end of deposition of the latter unit. The subtle but critical dif-
same general area appear also to suggest assemblages having ference between these two siltstone facies lies in the abundance
“montane” affinities (N.P. Hooker, 2008, personal commun.). of coarser-grained sand laminae, as well as a greater number of
Across the study area as a whole, active deposition of the sand-filled vertical cracks, that are observed within the lower
unnamed middle Unayzah member ultimately ceased, and soil- zones of the Unayzah A member.
forming processes prevailed. These resulted in the formation The siltstones within the unnamed middle Unayzah member
of the paleosols that in many places characterize the top of the have been interpreted above as the depositional infill of alluvial
unnamed middle Unayzah member. Those soils represent a depo- floodplain lakes. Sedimentological evidence for those lakes being
sitional hiatus and consequently mark a disconformable surface relatively long-lived bodies of standing water is much greater
that separates the unnamed middle Unayzah member from the than is seen in the lower Unayzah A member. In the latter case,
overlying Unayzah A member. the sediment is similarly interpreted to be of a shallow lacustrine
origin, wherein waterlain deposits are indeed preserved (as indi-
UNAYZAH A MEMBER cated by the ripple forms and evidence of slumping). However,
there is much more abundant evidence for fluctuation of the
The Unayzah A member has a widespread distribution across lake levels, and consequent repeated and widespread exposure
the study area. Generally its thickness ranges from 150 to 300 ft and desiccation within these lower Unayzah A silty, very fine–
(45–90 m) but does not display the extreme variation recognized grained sandstones. That evidence includes the common occur-
in the Unayzah C or Unayzah B members. Although identified rence of subvertical sand-filled cracks (desiccation cracks) and
in the subsurface across the area of investigation, the Unayzah the abundant horizontal laminae of coarser-grained sand. These
A member nonetheless is difficult to correlate in detail from features are interpreted as being related to quasi-planar adhesion
wireline logs. This is because of the intrinsic variability in laminae (sensu Hunter, 1980) or adhesion laminations (Kocurek
depositional facies distribution, confirmed from examination of and Fielder, 1982) that were laid down upon exposed damp sur-
>3150 ft (945 m) of Unayzah A core in the present study. This faces. The lowermost interval of the Unayzah A member is thus
62 Melvin et al.

interpreted to represent deposition within an ephemeral lake basin flood events. The repeated occurrence of desiccation in the finer,
setting (playa lakes) and as such indicates much drier conditions mud-draped sediment emphasizes the episodic nature of deposi-
than those that prevailed during deposition of the unnamed mid- tion of these beds in a semiarid environment; i.e., these are the
dle Unayzah member. sporadic deposits of ephemeral streams. The overall depositional
profile of the Unayzah A member at well 8 (Fig. 15, 132–238 ft)
Semiarid Ephemeral Stream Deposits displays a clear upward-coarsening trend, passing gradationally
from a siltstone-dominated regime upward into medium- and
These deposits constitute the greater part of the upper Unayzah coarse-grained sandstones. The siltstone-dominated interval rep-
A member in a number of wells across the study area. Thus, for resents the semiarid ephemeral lake (playa lake) deposits of the
example in well 8, the upper part of the Unayzah A member dis- lower Unayzah A member, discussed above. The upward-coars-
plays a number of interbedded sandstones and siltstones (Fig. 15, ening depositional trend that overlies these playa lake sediments
186–238 ft). The sandstones are thin-bedded units, rarely more in well 8 represents the progradation into the lake of a terminal
than ~1 ft (30 cm) thick, that make up the lower parts of upward- alluvial fan, or a lake-marginal bajada, that was traversed by a
fining couplets. They are fine to medium grained, moderately number of shallow, ephemeral stream channels.
to poorly sorted, and generally either massive or flat laminated
with sharp, erosional lower contacts (locally with numerous Eolian Erg Center to Erg Margin Deposits
small mud clasts). These sandstones can be single units, or in
places they comprise thin amalgamated beds (Fig. 16A). They Although the semiarid ephemeral stream deposits described
are commonly overlain by finer-grained beds of similar thickness above are encountered in the upper Unayzah A member in sev-
that form the upper parts of the depositional couplets referred eral places in the subsurface of eastern and central Saudi Ara-
to above. These beds comprise very fine–grained silty sandstone bia, they do not everywhere represent the dominant depositional
that displays well-developed ripple cross-lamination (Fig. 16B). facies association of that stratigraphic unit. In many places the
In places these silty, rippled sandstones contain an abundance uppermost part of the Unayzah A member consists of a complex
of mud drapes that display well-developed desiccation features of sandstone facies that can be ascribed in general to an eolian
(Fig. 15, 194–196 ft; Fig. 16C). setting (Fig. 17), and which are summarized below.
The lower, sharp-based and coarser sandstones of these Eolian dune cross-bedded sandstones are common
upward-fining couplets are interpreted as having been deposited throughout the upper Unayzah A member. They comprise fine-
rapidly by flash floods, and the overlying, finer-grained, and ripple- to medium-grained, well to very well sorted sandstones, with
laminated sediment represents the waning flow deposits of those very well rounded and frosted grains of quartz sand that occur

A B C

Figure 16. Core photographs illustrating aspects of the semiarid ephemeral stream facies association of the upper Unayzah A member (see text
for discussion). (A) Amalgamated medium-grained sandstones representing flash flood deposition. Note the sharp, scoured, and loaded contacts
(arrows), suggestive of rapid, high energy sedimentation. (B) Argillaceous, ripple laminated fine- to very fine–grained sandstones indicative of
waning flow conditions following flash flood events. (C) Abundance of desiccated clay drapes in the ripple laminated facies (arrows).
Figure 17. Core log through ~210 ft (63 m) of core in well 21, representing the various facies identified in the eolian facies associa-
tion of the upper Unayzah A member. Note in particular the abrupt horizontal upper terminations of the cross-bedded facies in many
places. The facies are identified as follows: (i) eolian dune cross-bedded sandstones; (ii) eolian sand sheet sandstones; (iii) paleosols;
(iv) sandy interdune deposits (damp interdunes); (v) silty interdune deposits (wet interdunes). Modified from Melvin et al. (2005).
64 Melvin et al.

in pronounced high angle (>30°), grain size–segregated cross- medium-grained sandstones, with very well rounded and frosted
laminations. Those laminations may be very closely spaced grains of sand that occur in low angle to flat, grain size–segregated
(“pin-striped,” sensu Fryberger and Schenk, 1988) (Fig. 18A) laminations (Fig. 17, 90–98 ft; Fig. 18B). Many sets of these
and inversely graded, where they are interpreted as wind-ripple laminations have a pin-striped appearance, suggesting an origin
laminations (subcritically climbing translatent strata: Kocurek from wind-ripple migration. Flat-based, lenticular (convex-up)
and Dott, 1981) that formed on the slip faces of eolian dunes. accumulations of coarse-grained, well rounded sand are also
Alternatively, the cross-laminations are a few centimeters thick, observed within these deposits. It is possible in places to distin-
in which case they are better interpreted as grain flow cross-strata guish very low angle truncations separating different sets of the
(sensu Kocurek and Dott, 1981) that formed by gravity sliding of low-angle laminated sand. Some examples of this facies display
sand down the dune slip faces. These eolian dune cross-bedded disruption of the laminae, attributed to plant roots, as well as to
deposits are readily identified on downhole image logs, which, other unknown forms of faunal bioturbation (insects? spiders?)
crucially, allows this facies to be recognized in uncored wells. (Fig. 18B). All of these characteristics compare favorably with
Eolian sand sheet deposits are also widely recognized features seen in eolian sand sheets, as they were described by
within the upper Unayzah A member. They comprise fine- to Fryberger et al. (1979).

C
A

B
D
Figure 18. Core photographs illustrating facies representative of the eolian erg center to erg margin facies association in the upper
Unayzah A member. (A) Pronounced high-angle cross-lamination of the eolian dune facies. These laminations include very closely
spaced (pin-striped) wind ripple laminations (wr) and thicker grain flow laminations (gf). (B) Very low-angle wind ripple lamination
(wr) of the eolian sand sheet depositional facies. Note: low-angle truncations (arrowed) and subtle high-angle disruptions (insect bur-
rows?) (arrowed, ib). These laminated sands overlie an interval at the bottom of the photograph that shows indistinct adhesion ripple
(ar) structures. (C) Irregular crinkly laminations that are associated with damp interdune settings. White patches are irregular occur-
rences of anhydrite cement. (D) Disrupted and brecciated sandstone typical of many paleosols found within the arid eolian facies
association of the upper Unayzah A member.
Late Paleozoic Gondwanan glaciation in Saudi Arabia 65

Sandy interdune deposits (damp interdunes) are com- ephemeral fluvial) dominated, the repeatability of facies is clear,
monly recognized in the upper Unayzah A member. They consist and the facies boundaries are seen to be sharp (Fig. 17). In partic-
of moderately sorted, fine- to medium-grained sandstones that ular, the eolian dune cross-bedded sandstones are in many cases
contain a heterogeneous assemblage of depositional structures, abruptly truncated by a horizontal contact with the overlying
including irregular to crinkly, variably continuous laminations facies (Fig. 19A; Fig. 17, 24 ft, 42 ft, 125 ft, 151 ft, 163 ft). These
and very thin (centimeter scale) beds with well-developed adhe- abrupt horizontal truncations of eolian cross-strata are identified
sion ripple laminations (Fig. 17, 84–90 ft; Fig. 18C). The occur- as Stokes surfaces (after Stokes, 1968) and are attributed to a ris-
rence of adhesion ripple laminations suggests that these sediments ing water table in the Permian eolian depositional environment
were laid down upon a damp substrate (cf. Kocurek and Fielder, (see Fryberger et al., 1988). The stratigraphic significance of the
1982). The irregular nature of the laminations is reminiscent of identification in the core of such rises in the paleo–water table
textures observed in damp interdune, or sandy sabkha, environ- was emphasized by Melvin et al. (2005).
ments (e.g., see Fryberger et al., 1983). Given the association of Subsidiary to the current work, a field-scale study was under-
these sediments with facies (described above) that were clearly taken of the Unayzah A member at the southern end of the giant
deposited in an arid, eolian-dominated setting, these irregularly Ghawar structure in eastern Saudi Arabia that included wells
laminated sandstones of the upper Unayzah A member are simi- 15 through 20 (see Fig. 1A). Details of that study are discussed
larly interpreted to represent a damp interdune environment in by Melvin et al. (2010). There, the lower part of the Unayzah A
close proximity to the paleo–water table. member (described above as dominated by ephemeral lake facies)
Silty interdune deposits (wet interdunes) comprise siltstones was seen in cores from many of the wells to terminate upward in
and silty, very fine–grained sandstones showing variably continu- a thin (decimeter scale), subtle upward-fining trend, with a pas-
ous to lenticular and crinkly lamination. In places small-scale rip- sage from silty, very fine–grained sandstone to siltstone sensu
ple cross-lamination is seen, suggesting subaqueous deposition. stricto (Fig. 19B). That upward-fining trend can be interpreted as
These sediments occur in intervals that are rarely >1 ft (30 cm) to a deepening of the lower Unayzah A lake. The surface represent-
2 ft (60 cm) thick. They show many depositional similarities to ing the top of this vertical, upward-fining trend among the wells
the semiarid lake deposits described above. The significant dif- has been interpreted further to represent in a lateral dimension
ference lies in the thickness of their occurrence. They are thus the maximum extent of the lake (MEL) (Melvin et al., 2010).
interpreted not as ephemeral (playa) lake sediments but rather Above this MEL horizon, each well within the subsidiary study
as the deposits of temporary, very shallow interdune ponds that area displays its own assemblage of the various upper Unayzah
formed at times when the water table rose above the depositional A member eolian depositional facies (described above). Within
surface within an otherwise arid (desert) environment. these assemblages, respectively, a number of cyclical rises in the
Sandy paleosol horizons have also been recognized at a paleo–water table can be recognized based on the identification
number of localities within the eolian-dominated upper Unayzah of Stokes surfaces in core, as well as on image logs, and the gen-
A member. They are 0.3–4.0 ft (0.09–1.2 m) thick (Fig. 17, eral associations of the various facies. Crucially, when the data
53–57 ft, 126 ft, 151 ft), and comprise poorly sorted, very fine to from each well are hung stratigraphically using the MEL horizon
medium-grained sandstones, with patchy carbonate cementation as a datum, many of the interpreted rises in the paleo–water table
and showing varying degrees of disrupted texture (Fig. 18D). that are recognized above the MEL in the various wells correlate
These paleosol horizons suggest that the paleo–water table was exactly, irrespective of the depositional facies within which they
relatively high at the time of their formation. are found (Fig. 20) (Melvin et al., 2010). This correlation extends
Individually, the above-described eolian and eolian-related along a distance of ~65 km. Thus, a correlatable layering scheme
facies occur commonly throughout the upper part of the Unayzah is crucially established that is related to paleo–water table fluc-
A member. The mutual associations of these facies, as well as tuations within the eolian-dominated Unayzah A member. The
their abundances relative to each other, vary markedly from well correlation demonstrably carries through different depositional
to well, and indeed from field to field across the study area. Thus facies tracts and is intrinsically established on sequence strati-
in some places the upper part of the Unayzah A member is domi- graphic principles.
nated (almost to the exclusion of any other facies) by stacked
eolian dune cross-bedded sandstones that are clearly seen in both Terminal Facies of the Uppermost Unayzah A Member
core and image-log data. These dune-dominated occurrences are
interpreted as representative of eolian erg-center settings. Else- The preceding discussion has described how the Unayzah
where a much greater mix of the various eolian facies (described A member is characterized by a wide variety of different deposi-
above) is seen: those mixed facies associations are interpreted tional facies that generally fall into three distinctive facies asso-
as eolian erg-margin deposits. Similar eolian-related facies asso- ciations, all of which are strongly indicative of deposition within
ciations have been described from the Permian Cedar Mesa a semiarid to arid setting. Thus this member almost universally in
Sandstone in Utah by Mountney and Jagger (2004). If well 21 its lower part comprises playa lake deposits that pass upward into
(Fig. 17) is taken as a typical example of a well within which the either ephemeral stream facies or eolian erg to erg margin facies.
facies of the upper Unayzah A member are eolian (as opposed to The ultimate termination of these conditions is represented in
Figure 19. Core photographs illustrating features of
stratigraphic significance within the eolian facies
association of the upper Unayzah A member. (A) Abrupt
upper horizontal termination of pronounced eolian
dune cross-bedding. This is a “Stokes surface” and is
overlain in this example by sand sheet and interdune
facies. It represents a rise through the dune deposits
of the paleo–water table. Note: Large white spots are
postdepositional nodules of anhydrite. (B) Upward-
fining of very fine–grained sandstones into siltstone
at the top of the lower Unayzah A member ephemeral
lake deposits in well 18. The abrupt contact with over-
lying coarser (eolian) sandstones defines the maximum
extent of the lake (MEL) in this well. See text for full
discussion.

A B

Figure 20. Stratigraphic cross section through the Unayzah A member in selected wells at the southern end of the Ghawar structure (see Fig. 1A).
The section has the MEL (maximum extent of the lake) horizon (see text for discussion) as its datum. Note how a number of maximum wetting
horizons above the MEL can be identified (from facies relationships seen in core and image logs), and that these appear to be correlatable along
the length of the section, irrespective of the facies within which they are found. These wetting cycles are interpreted to be related to fluctuations
in the paleo–water table. PKU—pre-Khuff unconformity.
Late Paleozoic Gondwanan glaciation in Saudi Arabia 67

many wells across the study area by very different and distinctive a large number of depositional facies are readily identified. It is
facies at the top of the Unayzah A member (Fig. 21). also evident that most of those facies reflect deposition within
Thus in well 24 the stratigraphically highest eolian deposits a significantly arid environment. Thus, depending upon paleo-
(Fig. 21A, 1–9 ft) are overlain by 21 ft (6.3 m) of finer-grained geographical location within the study area, depositional facies
“interdune” deposits that differ from those normally seen in the associations reflect either (1) arid to semiarid conditions that are
upper Unayzah A member in that they display minor evidence characteristic of terminal alluvial fans and bajadas or (2) wide-
of burrowing. Those sandstones are overlain by a highly distinc- spread eolian erg systems, both of which encroached upon ephem-
tive interval 7 ft (2.1 m) thick, showing an abundance of burrows eral lakes (playas) whose dimensions fluctuated throughout time.
(Fig. 21A, 32–39 ft; Fig. 22A). Above this intensely bioturbated Stratigraphic analysis of one of these erg systems by Melvin
zone there occurs in well 24 a thick development of paleosols et al. (2010) revealed a high degree of cyclicity among facies that
(Fig. 21A, 39–83 ft). Those paleosols appear to comprise at least results in a layering scheme that is correlatable both within and
five individual paleosol deposits each of which displays a number between depositional facies tracts (see earlier discussion).
of probable soil “zones.” The facies characterizing these zones This internal stratigraphy within the eolian-dominated
vary from very argillaceous, poorly sorted sandstones with an Unayzah A member has been interpreted to be a direct reflection
abundance of cutans and possible rooting features to extremely of fluctuations in the level of the paleo–water table throughout the
well developed, massive to nodular silcretes (e.g., Fig. 21A, time of its deposition. These correlatable rises in the paleo–water
68–70 ft; Fig. 22B). This 44 ft (13.2 m)–thick interval of paleo- table are intrinsically allostratigraphic in nature, and it is tempting
sols is superseded in this well by a heavily rubbleized interval of to consider their cyclicity as possibly having its basis in the Earth’s
red silty mudstone that displays in its lower part an abundance orbital fluctuations. If the Gondwanan Permo-Carboniferous ice
of very thin (millimeter scale) laminations of fine- to medium- sheet was still significantly (albeit very distally) in existence dur-
grained sandstone. ing deposition of the Unayzah A member, and if such orbital fluc-
This sequence in well 24 can be compared favorably with the tuations had an impact on the melting of the ice and the consequent
highest part of the upper Unayzah A member in well 21. In the postglacial global transgression, then it may be that even within
latter well, some 300 km distant from well 24 (see Fig. 1A), the the terrestrial rocks of the Unayzah A member there is a record of
highest eolian dune sand is overlain by ~20 ft (6.0 m) of paleo- the pulsatory nature of that transgression. It appears to be highly
sol deposits (Fig. 21B, 10–30 ft). Those rocks subsequently pass significant in this regard that the stratigraphically highest of these
upward through some upward-fining sandstones into an interval terrestrial deposits in locations so widely separated as well 24
of low-angle laminated sandstone wherein the laminations are and 21 (Fig. 1A) are superseded by a thin zone of sandstone that
severely disrupted by intense burrowing (Fig. 21B, 39–45 ft; displays a distinctly marine signature in the form of intense bur-
Fig. 22C). Overlying these burrowed sediments is another inter- rowing activity (Figs. 21 and 22). This would suggest that the
val of thickly developed paleosols (Fig. 21B, 45–76 ft). Again, episodic rises in base level inferred from the cyclicity identified
these paleosols appear to comprise a number (five or six) of from rises in the paleo–water table did indeed ultimately manifest
stacked, individual paleosol deposits, within which various facies themselves with a breakthrough of marine waters at the very end
representing soil zonation occur. Those zones are similar to those of the time of deposition of the Unayzah A member.
identified in well 24 (see above) in that they vary from very argil- It is clear that the end of deposition of the Unayzah A mem-
laceous, poorly sorted sandstones with an abundance of cutans ber in most areas was marked by a (probably prolonged) period
and apparent root-related structures to well-developed massive to of minimal deposition to nondeposition. This was reflected
nodular silcretes (e.g., Fig. 22D). It appears that, notwithstand- initially by the intensely bioturbated sandstone (indicating low
ing the great distance that separates wells 24 and 21, there are rates of sedimentation) and was manifest ultimately by the
great similarities between the two wells in terms of the facies development of the very thick and pervasive paleosols seen in
associations in the uppermost parts of the Unayzah A member. many places at the top of this stratal unit. The widespread extent
Specifically there appears to have been a thin, probable marine of those paleosols testifies to the development of a sustained,
event, identified by intensive burrowing, that was followed in high water table that may indicate a change in climate (decreas-
both cases by the development of thick, stacked paleosol depos- ing aridity). The significance of this inferred prolonged period
its. In well 24 these paleosols are overlain by fine-grained red of minimal deposition to nondeposition in terms of the overall
beds; in well 21 there is a 5 ft gap of non-recovery of core that is tectono-stratigraphic evolution of Saudi Arabia is discussed in
superseded by sandstones of shallow-marine origin and ascribed the final section of this paper.
to the Basal Khuff Clastics member of the Khuff Formation (see
later discussion). BASAL KHUFF CLASTICS MEMBER OF THE
KHUFF FORMATION
Origin and Evolution of the Unayzah A Member
Senalp and Al-Duaiji (1995) observed that the Unayzah
It is clear from the foregoing discussion that the Unayzah Formation at outcrop is truncated by an unconformity known
A member is a highly complex stratigraphic unit, within which as the pre-Khuff unconformity. In the subsurface of the study
Figure 21. Core logs through the uppermost
Unayzah A member at (A) well 24 and (B) well
21. These wells are ~300 km apart. Nonethe-
less, they display a remarkable similarity in their
stratigraphic organization. Note in particular the
occurrence in each well of a pronounced biotur-
bated sandstone zone (bs) ~25–30 ft (7.5–9.0 m)
above the highest occurrence of eolian dune
cross-bedded sandstone (dxb). In each case the
bioturbated zone is overlain by a similar thick-
ness (30–35 ft, 9.0–10.5 m) of paleosols (ps) that
display similar internal stratigraphic organiza-
tion. See text for full discussion.
Late Paleozoic Gondwanan glaciation in Saudi Arabia 69

A B C

Figure 22. Core photographs showing aspects of the uppermost Unayzah


A member in wells 24 and 21. (A) Intensely bioturbated sandstone in
well 24 (Fig. 21A, 33–38 ft). (B) Pervasive development of silcrete in
well 24 (Fig. 21A, 68–70 ft). (C) Intensely bioturbated sandstone in well
21 (Fig. 21B, 38–43 ft). (D) Pervasive development of silcrete in well 21
(Fig. 21B, 53 ft).

area (Fig. 1A), this unconformity separates the sandstones of Upward-Fining Units of the Basal Khuff Clastics
the Unayzah from the overlying Khuff Formation. Although the
Khuff Formation is widely recognized as a carbonate-dominated In many places in the western part of the study area, and
stratigraphic unit (Sharland et al., 2001; Vaslet et al., 2005), its exemplified at wells 24 and 23 (Fig. 23A and Fig. 23B), the pre-
lowermost deposits in many places are characterized by a series Khuff unconformity is abruptly overlain by pebble conglomer-
of alternating sandstones, shales, and thin carbonates that sit ates and very coarse–grained, pebbly sandstones that fine upward
directly upon the pre-Khuff unconformity. Those deposits are over several feet into argillaceous fine-grained sandstones and
described herein as the Basal Khuff Clastics member of the (ultimately) dark gray mudstones. The upward-fining sandstones
Khuff Formation. They include the unit identified by Senalp and conglomerates have clasts that are variably subrounded to
and Al-Duaiji (1995) and Vaslet et al. (2005) as the Ash Shiqqah angular in nature and polymict in composition. Well 24, for exam-
member of the Khuff Formation. The depositional facies of the ple, displays an assemblage of angular detritus, including quartz,
Basal Khuff Clastics member change in their character increas- feldspar, jasper, granite, sandstone, siltstone, and mudstone. The
ingly in a westward direction, as will be discussed below. Fur- mudstones that cap these upward-fining units in places display
thermore, the higher carbonate members of the Khuff Formation rooted textures in their upper part and commonly pass up into
eventually overstep the Basal Khuff Clastics member, as well thin bioturbated sandstone and associated bioturbated mudrocks
as the Unayzah and ultimately all older Paleozoic deposits, in with calcareous concretions (e.g., well 23, Fig. 23B).
the same direction. In extreme westerly locations, these upper The upward-fining character of these coarse-grained sedi-
Khuff carbonates rest directly upon Proterozoic rocks of the Ara- ments in the western parts of the study area, and the lack of
bian Shield. marine indicators except higher in the section, suggest that the
Figure 23. Selected core logs through the Basal Khuff Clastics member across the study area. (A) In well 24 the pre-Khuff uncon-
formity (PKU) is overlain by a coarse breccia that fines upward into a highly carbonaceous paleosol. That in turn is overlain by
~5 ft (1.5 m) of thin-bedded marine bioturbated sandstones that pass up into gray mudstones (well 24). (B) In well 23 the pre-Khuff
unconformity is abruptly overlain by upward-fining pebbly sandstone that passes upward into gray pedogenically altered mudrock.
That soil is overlain by very thin, bioturbated fine-grained sandstone. (C) At well 15 the pre-Khuff unconformity is overlain by a
mudrock-dominated interval that contains a few significant bioturbated sandstones. The mudrock facies show a cyclicity wherein
fissile shales pass upward into more blocky and highly carbonaceous mudstones. (D) At well 17 the section directly above the pre-
Khuff unconformity is characterized by a number of upward-fining cycles each of which passes into well-developed argillaceous
and carbonaceous soils. These are superseded by a mudrock-dominated succession that contains bioturbated sandstones in its lower
part, and within which can be seen a similar subtle cyclicity as at well 15. (E) The Basal Khuff Clastics member at well 21, in the
eastern part of the study area, solely comprises shallow-marine depositional facies, including bioturbated sandstones and mudstones,
and cross-bedded sandstones. See text for full discussion of these wells and their relationships.
Late Paleozoic Gondwanan glaciation in Saudi Arabia 71

lowermost parts of the Basal Khuff Clastics member in those 24 (Fig. 23A) the paleosol described above from the Basal Khuff
areas are characterized by fluvial deposition. The coarseness of Clastics is abruptly overlain by ~5 ft of thin-bedded (decimeter
the sandstones is indicative of a significant drop in base level scale) coarse-grained sandstones that are burrowed throughout.
(and/or uplift in the source area) pursuant to the creation of the The depositional facies described above from well 21 were
pre-Khuff unconformity. laid down in a shallow-marine environment wherein the thin sets
of cross-bedded sandstone represent offshore marine bars, and
Paleosols of the Basal Khuff Clastics the various bioturbated facies are indicative of slightly deeper
(or otherwise protected) conditions that allowed a prolifera-
In a number of the wells that have cored the Basal Khuff tion of infaunal activity. The bioturbated sandstone in well 17
Clastics member, distinctive paleosols are recognized that are, was likewise deposited in a marine setting. It is probable that
significantly, quite different in character from those soils identi- the large cobble-sized clasts of siliceous rock were eroded by
fied and earlier described from the top of the Unayzah A member. storms from silcretes in the uppermost Unayzah paleosols in
Thus in wells 24 (Fig. 23A, 35–40 ft) and 17 (Fig. 23D, 11–28 ft) nearby locations. The coarse-grained (i.e., sand-sized) marine
fine- to very fine–grained, poorly sorted and very argillaceous detritus observed in wells 21 and 17 is not recognized in well
and carbonaceous sandstones are identified low in the Basal 15. There, the first definitive indications of marine deposition are
Khuff Clastics section, overlying upward-fining coarser-grained seen in the thin bioturbated sandstone that occurs ~5 ft above
sediment described above. The example from well 17 comprises the pre-Khuff unconformity (Fig. 23C, 13–14 ft; Fig. 24D). Sig-
four small-scale, upward-fining beds that display a high degree nificantly, palynological data from that zone show an influx of
of vertical rooted texture throughout (Fig. 23D, 11–28 ft). These abundant scolecodont debris that can be interpreted as indicating
paleosols of the Basal Khuff Clastics member generally occur a significant marine influx at that point (J. Filatoff, 2004, personal
low in that stratal unit, directly subjacent to rocks displaying evi- commun.). It is clear that a similar marine depositional signature
dence of a marine influence in their deposition. This association is also evident in well 24 in the western part of the study area
leads to their interpretation as representing coastal marshlands in (Fig. 23A, 41–47 ft).
low-lying (estuarine) areas. As such they represent the first signs
of the onset of the Khuff marine transgression. Mudrocks of the Basal Khuff Clastics

Marine Sandstone Deposits of the Basal Khuff Clastics In wells 17 (Fig. 23D) and 15 (Fig. 23C) the marine sands
described above are overlain by up to 40 ft (12 m) of dark gray
In well 21 (Fig. 1A) the thick paleosols described above from mudrocks. They locally contain thin-walled bivalve shells and
the uppermost Unayzah A member (see Fig. 21B) are separated display a number of subtle cycles, each of which is ~5 ft (1.5 m)
by ~6 ft (1.8 m) of non-recovery of core from a unit that is at least thick. Within each cycle, there is a passage upward from fissile,
21 ft (6.3 m) thick, and comprises sandstones and mudstones that gray, silty mudstone to blocky and crumbly, highly carbonaceous
are entirely different in character (Fig. 23E, 12–33 ft). Those sed- and rooted mudstone (Fig. 23C, 15–47 ft; Fig. 23D, 45–77 ft).
iments are assigned to the Basal Khuff Clastics member in this These fine-grained depositional cycles are interpreted to repre-
well. They comprise (1) interbedded dark gray, silty mudstones sent successive periods of lagoonal infill in a marginal marine
and thin (centimeter scale), heavily bioturbated fine-grained setting, each one terminating in a coastal swamp.
sandstones (Fig. 24A); (2) massive, intensively bioturbated sand-
stones up to 3 ft (0.9 m) thick; and (3) cross-bedded, fine-grained Origin and Evolution of the Basal Khuff Clastics Member
sandstones that occur in bedsets up to 5 ft (1.5 m) thick, with of the Khuff Formation
individual sets up to 0.5 ft (0.15 m) thick (Fig. 24B), and which
only rarely display any bioturbation. In well 17 (Fig. 23D), the It is clear from the foregoing not only that the Basal Khuff
paleosols described above from the lowermost parts of the Basal Clastics member represents an extremely heterogeneous assem-
Khuff Clastics member are overlain by a heavily bioturbated blage of depositional facies but also that that variation is geo-
sandstone unit 6 ft (1.8 m) thick that contains large (cobble-sized) graphically constrained. Thus in the southeast of the study area
clasts of very fine–grained silica rock (Fig. 24C). Those clasts are around well 21 the lowermost Basal Khuff Clastics member
probably derived from nearby nodular silcrete horizons within is specifically characterized by significant deposits of shallow-
the uppermost Unayzah A paleosol (see earlier discussion). In marine sandstones. Somewhat to the north and west, around well
well 15 the pre-Khuff unconformity is seen clearly to truncate 17, the lowermost Basal Khuff Clastics consist of very different
eolian sandstones of the Unayzah A member (Fig. 23C). This sediments, comprising relatively thin, upward-fining sandstones
unconformity is overlain by a very thin (1–2 cm) sandstone bed, that have been interpreted above to be highly argillaceous and
above which the sequence is dominated by dark gray carbona- carbonaceous paleosols. Those paleosols are believed to have
ceous mudstones. Within those mudstones a very thin (0.3 ft, formed in a low-lying coastal plain or estuarine marsh environ-
0.09 m) bioturbated sandstone (Fig. 24D) occurs ~5 ft above the ment. Above these paleosols in well 17, bioturbated marine sand-
pre-Khuff unconformity in this well (Fig. 23C, 13–14 ft). In well stones are recognized, although they are not as fully developed as
72 Melvin et al.

A B
D

Figure 24. Core photographs showing some characteristics of the marine sandstone facies of the Basal Khuff Clastics member.
(A) Interbedded mudstones and fine-grained sandstones, displaying intense burrowing activity (well 21). (B) Cross-bedded sandstones
of the offshore shallow-marine-bar environment (well 21). (See also Fig. 23E.) (C) Intensely bioturbated marine sandstone, contain-
ing a large clast of chert rock (Ch). The latter is considered to have been ripped out of nearby paleosol deposits by storms (well 17).
(D) Thin, sharp-based bioturbated sandstone overlying gray mudrock (m). The sandstone is interpreted to be a relict shallow-marine
storm deposit laid down below storm wave base (well 15). See text for further discussion.

those that occur in well 21; at well 15, shallow-marine sandstone a fluvial deposit that is ultimately transgressed by a thin marine
is also recognized, but there it occurs only as a very thin bed sandstone (Fig. 23B). At well 24 a basal fluvial breccia passes
(Fig. 24D) within an otherwise mudstone-dominated sequence. up into a paleosol that is in turn capped by ~5 ft (1.5 m) of thin-
Those mudstones show a subtle cyclicity suggestive of repeated bedded bioturbated sandstones (Fig. 23A, 32–47 ft).
infill of lagoons in a coastal setting. In other words, the Basal The clear distinction within the lowest parts of the Basal
Khuff Clastics in this location display a subtly less marine signa- Khuff Clastics between shallow-marine depositional facies in the
ture than equivalent deposits somewhat to the southeast at well southeast and continental (fluvial) facies in the west demonstrates
21. Even farther to the west in the study area, at well 23 the Basal the gradual encroachment of the marine transgression from the
Khuff Clastics member is represented in its lowermost part by southeast to the west in earliest Khuff times. The transgression
Late Paleozoic Gondwanan glaciation in Saudi Arabia 73

reached its maximum expression with deposition of the various tiate the much closer association of the Unayzah B member with
carbonate-evaporite cycles of the Khuff Formation, most recently the unnamed middle Unayzah member in mid- to high southerly
described by Vaslet et al. (2005). latitudes, from the near tropical setting of the Unayzah A mem-
ber. These observations are discussed further below.
PALEOMAGNETIC STUDIES
DISCUSSION
In the course of conducting these stratigraphic studies of the
Unayzah and lowermost Khuff Formations, an opportunity arose Sharland et al. (2001) discussed how Tectonostratigraphic
to carry out a paleomagnetic pilot study on core samples from Megasequence TMS AP5 is the last megasequence of the Arabian
two of the wells, wells 7 and 22. A total of 38 samples were col- plate to have been dominated by siliciclastic sediments (Unayzah
lected under appropriately controlled, nonmagnetic conditions, Formation). Only the Basal Khuff Clastics member, overlying the
and analyzed with a view to determining whether or not the paleo- pre-Khuff unconformity at the base of TMS AP6, represents any
latitude at the time of their deposition could be identified. The further siliciclastic deposition in Saudi Arabia prior to the north-
samples were taken from rocks that were independently assigned ward drift of the Arabian plate into subtropical latitudes where
to the Unayzah B member, the un-named middle Unayzah mem- carbonate and evaporite deposition predominated (cf. Beydoun,
ber, and the Unayzah A member. The results of this pilot study 1991; Al-Fares et al., 1998).
are presented in Figure 25. TMS AP5 is marked at its base in Saudi Arabia by the Her-
Although the samples from each stratigraphic unit show a cynian unconformity, which represents erosion associated with
range of values, it is clear that each unit has its own signature the mid-Carboniferous “Hercynian” tectonic inversion event
with respect to its interpreted paleolatitude and that there is no (Al-Husseini, 2004). Those erosional processes were intensified
overlap in the respective data sets. Thus the samples from the with the almost coincident inception of the late Paleozoic Gond-
Unayzah B member were deposited in paleolatitudes represented wanan glaciation in Arabia. The upper boundary of TMS AP5
by a mean value ~75° S. The unnamed middle Unayzah mem- is the pre-Khuff unconformity, dated as mid-Tatarian in Saudi
ber appears to have been laid down when the study area was at Arabia by Stephenson and Filatoff (2000b). It represents erosion
a (mean) paleolatitude of ~55° S. Deposition of the Unayzah A following the early Late Permian rifting associated with the cre-
member did not occur until the area of investigation lay in paleo- ation of the Neotethys Ocean (cf. Bishop, 1995; Loosveld et al.,
latitudes of ~27° S. These data clearly reflect the northward drift 1996). It can thus be considered to be the “Break-up Unconfor-
of the Arabian plate throughout the Early to Middle Permian, as mity” associated with the continental rifting and spreading of the
was independently documented by previous authors (Beydoun, Sanandaj-Sirjan and central Iran terranes from the Arabian plate
1991; Al-Fares et al., 1998). They also appear to clearly differen- (Sharland et al., 2001). Angiolini et al. (2003), however, believe
that in the Sakmarian of Oman there are two tectono-eustatic
transgressive events that are related to the end of the Gondwanan
Unayzah A glaciation and the concomitant tectonic evolution during rifting
and initial opening of the Neotethys Ocean. Those authors thus
consider the “Break-up Unconformity” to be represented by an
unconformity that is seen within TMS AP5, and which formed
Unnamed middle much earlier than the pre-Khuff unconformity (Angiolini et al.,
Unayzah member
2003). Clearly, the evolution of Tectonostratigraphic Megase-
Younger

quences TMS AP5 and TMS AP6 on the Arabian plate represents
Unayzah B a complex interplay of tectonic, climatic, and eustatic relation-
ships. An attempt is made below to discuss that evolution, in the
context of the depositional and stratigraphic development of the
various units described above within the Unayzah, as well as the
0 30 60 90 Basal Khuff Clastics member of the Khuff Formation of Saudi
Paleolatitude (degrees S) Arabia. Inevitably, that discussion necessitates consideration of
the equivalent stratal units elsewhere on the Arabian plate, spe-
Figure 25. Diagram based on paleomagnetic data, showing relative
position with respect to paleolatitude of the Unayzah B member, the cifically Oman.
unnamed middle Unayzah member, and the Unayzah A member. These As has been discussed above, the late Paleozoic South Polar
data sets individually display varying degrees of spread but nonethe- glaciation that affected so much of the Gondwanan continent is
less appear to show no overlap with each other. They illustrate clearly considered by Al-Husseini (2004) possibly to have commenced
the northward drift of the Arabian plate from Asselian–earliest Sak- in Arabia as early as middle Carboniferous (late Visean–early
marian time (Unayzah B member) to Artinskian–Kungurian(?) times
(Unayzah A member). The results were obtained from core plugs col- Namurian or Serpukhovian) times, and to have persisted for
lected under controlled (nonmagnetic) conditions and were analyzed 30–45 m.y. until the Early Permian (Sakmarian). This provided
by E. Hailwood (CoreMagnetics). ample time for an unknown number of glacial advances, retreats,
74 Melvin et al.

and readvances to have taken place, both in Saudi Arabia and of Stephenson et al. (2003). Melvin and Sprague (2006) proposed
Oman. Such multiple events are implicitly recognized in Oman that the unnamed middle Unayzah member red beds in Saudi
from palynological evidence within the lower Al Khlata (Oster- Arabia can be correlated with the lower Gharif member in Oman
loff et al., 2004a), which presumably reflects different interstadial and therefore may be considered generally correlative with the
depositional periods. In Saudi Arabia the palynology is not so OSPZ3 Zone. This implied that in Saudi Arabia MFS P10 would
forthcoming. It is tempting to think that the glacial readvances somehow have equivalence within the wholly terrestrial deposits
that have been postulated herein in relation to the shear zones of the unnamed middle Unayzah member of Melvin and Sprague
within the Unayzah C member may be correlative by some (2006). Ongoing palynological studies are providing encourag-
means with the Oman section. This is clearly an important area ing results that support this original proposition of Melvin and
for future study. Sprague (2006) that there is at least some stratigraphic equiva-
The final retreat phase of the Gondwanan glaciation in Arabia lence between the unnamed middle Unayzah member and the
(and by implication the onset of climatic amelioration) is mani- lower Gharif member (N.P. Hooker, 2008, personal commun.). In
fest in the great variety of predominantly glaciolacustrine depo- Oman, Angiolini et al. (2003) considered the paleoclimatic and
sitional facies recognized and described above in the Unayzah B paleotectonic characteristics of the Saiwan Formation at outcrop
member in Saudi Arabia as well as in the upper Al Khlata For- (equivalent in the subsurface to the Haushi Limestone at the top
mation (including the Rahab Shale) of Oman (Osterloff et al., of the lower Gharif member). They concluded from paleontologi-
2004a). Paleomagnetic evidence from the current study appears cal, paleoecological, and petrographic evidence that the opening
to suggest that at this time Saudi Arabia lay in high southerly of the Neotethys Ocean commenced north of Oman in about mid-
latitudes ~75° S. Notwithstanding minor readvances such as Sakmarian times (Angiolini et al., 2003). From the foregoing dis-
have been recognized in the local occurrence of stratabound push cussion of the sedimentary history of the Unayzah Formation, it
moraine deposits (Fig. 7), the overall character of the Unayzah B would appear nonetheless that these proposed early stages of the
member across most of the study area illustrates ongoing glacial Neotethys marine transgression did not find recognizable expres-
retreat, evident in the continuing filling and spilling over of an sion in the terrestrial unnamed middle Unayzah member in Saudi
environment that was dominated by glacial lakes. This wholesale Arabia. The significant drainage event postulated earlier as mark-
flooding of the postglacial landscape in Saudi Arabia is reflected ing the end of Unayzah B member times, and, most significantly,
in the clearly “transgressive” (deepening) nature of the glaciola- the end of the Gondwanan glaciation in Saudi Arabia, may well
custrine depositional facies associations. In the western part of have been initiated by distant tectonic events related to an incipi-
the study area the Unayzah B is characterized predominantly by ent breach as a first step toward opening of the Neotethys Ocean.
glacially induced deformation features. These are considered to In Oman, Guit et al. (1995) suggested the presence of
reflect the longer term presence of significant bodies of ice in a regional intra-Gharif unconformity related to a drop in rela-
these locations, and possibly even an element of long-lived alpine tive sea level at the top of the lower Gharif member, and Oster-
glaciation associated with highlands of the Al Batin Arch. loff et al. (2004b) also described how the base of the middle
The uppermost deposits of the Unayzah B member across Gharif marks an overall regression. In Saudi Arabia the contact
most of the subsurface of eastern Saudi Arabia (and of the Rahab between the unnamed middle Unayzah member and the overly-
Shale in Oman, Osterloff et al., 2004a) provide clear sedimento- ing Unayzah A member is considered to be a sequence bound-
logical evidence of the maximum melt-out of the Gondwanan ice ary that is at least disconformable on the subcropping formation
sheet. It remains to be seen whether or not these rocks represent (Melvin and Heine, 2004). It heralds the onset of a major shift in
the maximum flooding event of the entire Tectonostratigraphic paleoclimatic conditions, under which deposition of the Unayzah
Megasequence TMS AP5. That in turn demands a consideration A member became dominated by arid to semiarid eolian and
of the extent to which the postglacial transgression is reflected by ephemeral stream sedimentation. That climatic change was prob-
purely climatic (i.e., warming) controls, or by overriding plate ably brought about by the northward drift of the Arabian plate
tectonic factors. into lower southerly latitudes (see Beydoun, 1991; Al-Fares et al.,
The only maximum flooding surface (MFS) that has hitherto 1998). Northward drift is clearly reflected in the paleomagnetic
been formally recognized in TMS AP5 was identified in Oman data presented in the current paper (Fig. 25). Those data demon-
by Sharland et al. (2001) as MFS P10, which is represented by strate a large amount of northward migration toward tropical lati-
a bioturbated shale directly below the Haushi Limestone. This is tudes and can be inferred to represent a significantly long period
the “maximum flooding shale” of Guit et al. (1995) and occurs of time between deposition of the unnamed middle Unayzah
within the postglacial lower Gharif member (see Fig. 2). Stephen- member and the Unayzah A member. This in turn would suggest
son and Osterloff (2002) showed that MFS P10 is marked by the that the stratigraphic boundary between these two members rep-
occurrence of the acritarch Ulanisphaeridium omanensis, and resents a hiatus of considerable significance.
Stephenson et al. (2003) subsequently assigned it to the OSPZ3b The earlier discussion of the Unayzah A member has high-
Subbiozone. Angiolini et al. (2006) showed that the marine lighted the identification in places of a number of zones within
deposits of the overlying Haushi Limestone are Early Permian this stratigraphic unit, each of which is related to a demonstra-
(Sakmarian) in age and correlative with the OSPZ3c Subbiozone ble rise in the paleo–water table and is correlatable over tens of
Late Paleozoic Gondwanan glaciation in Saudi Arabia 75

kilometers, irrespective of the depositional facies within which it Khuff Clastics member: the boundary for that changeover is
occurs. The possibility has already been mooted that this cycli- manifest as the pre-Khuff unconformity.
cal rise in the paleo–water table through the sediments of the Osterloff et al. (2004b, p. 115) postulated some equiva-
Unayzah A member may be related to a pulsed rise in distant lence in the sequence development of the uppermost cycle of the
sea level, and as such may be a phantom expression of ongo- upper Gharif member in Oman compared with the upper Basal
ing marine transgression. The recognition of an apparently wide- Khuff Clastics member, while stressing that no age equivalence
spread bioturbated (and therefore probably marine) sandstone is implied. Indeed, Stephenson (2006) recently clarified how in
zone close to the top of the Unayzah A member is highly relevant Oman the upper Gharif member is assigned to palynozone OSPZ5
in this regard. of Stephenson et al. (2003), whereas the Basal Khuff Clastics
The middle Gharif member in Oman is considered the prob- member in Saudi Arabia is assigned definitively to the younger
able stratigraphic equivalent of the Unayzah A member in Saudi palynozone OSPZ6 (see Fig. 2). It has been shown earlier in this
Arabia (Stephenson et al., 2003). It has been described by Oster- paper how in Saudi Arabia the lowermost part of the Basal Khuff
loff et al. (2004b) as consisting “largely of non-marine clastics, Clastics member displays very different depositional facies from
indicated by red continental paleosols and non-marine facies.” place to place, ranging from shallow marine in the southeastern
Significantly, the latter authors also note that a “rapid vertical part of the study area to fluvial deposits in more westerly loca-
interdigitating nature of marine to non-marine clastics” has been tions (Fig. 23). This geographical demarcation of facies clearly
recorded in the middle Gharif member from cores from central reflects the onset of fully marine transgression to the south and
Oman. It is possible that such alternations of facies in Oman east of the study area. That transgression progressed steadily
reflect a more marineward expression of the cyclicity that has westward, as is reflected in (1) the diachroneity of deposits from
been seen and described herein from the predominantly continen- the upper Gharif member to the Basal Khuff Clastics, (2) the het-
tal sediments of the Unayzah A. That cyclicity culminates with erogeneity of the facies tracts within the Basal Khuff Clastics,
the prominent bioturbated (and therefore probably marine) zone and, ultimately, (3) by the superposition of Khuff carbonates over
that occurs in some wells in the uppermost parts of the Unayzah Proterozoic basement rocks in western central Saudi Arabia, as
A member (cf. Fig. 21), suggesting final breakthrough in latest discussed previously.
Unayzah A time by marine waters into the study area of eastern The pre-Khuff unconformity marks the top of the Tectono-
and central Saudi Arabia. stratigraphic Megasequence TMS AP5 of Sharland et al. (2001).
Thereafter the uppermost Unayzah A member displays Those authors describe how “progressive thermal uplift (the
widespread development of generally thick paleosols, suggesting precursor to continental rifting and spreading) is interpreted to
a prolonged period of nondeposition, with no marine indicators. have occurred … culminating in the regional ‘pre-Khuff uncon-
This is interpreted to be a reflection of the up-doming that has formity’ at the top of this megasequence” (Sharland et al., 2001,
been postulated by Sharland et al. (2001) in relation to the ther- p. 85). The change in Oman from continental or marginal-marine
mal uplift that preceded continental rifting and spreading, and deposits of the Gharif Formation to fully marine Khuff Forma-
which culminated in the regional pre-Khuff unconformity. tion carbonates in the Wordian (Broutin et al., 1995) (equivalent
It was noted earlier that the palynological changeover to early Kazanian) is interpreted “to reflect marine flooding fol-
between OSPZ4 and OSPZ5 is “probably the greatest recorded lowing thermal collapse of the new Arabian plate passive mar-
in the Permian palynological succession in Oman and Saudi gin with Neotethys” (Sharland et al., 2001, p. 89). Those authors
Arabia” (Stephenson et al., 2003). This palynological change observe that “in north Arabia the transition probably occurs
finds expression in Oman also in changes in depositional (envi- slightly later, at the base of the Tatarian (equivalent to middle
ronmental), chemostratigraphic, and mineralogical (heavy min- Capitanian) . . . possibly indicating a very rapid ‘unzipping’ effect
erals) signatures between the middle Gharif and upper Gharif from southeast to northwest.” The present paper presents data that
members (Stephenson et al., 2003; Osterloff et al., 2004b). It fully support these concepts. Furthermore, it seems logical that in
is thus strongly implicit that a significant depositional hiatus lithostratigraphic terms the upper Gharif in Oman is indeed an
exists between these two members. Osterloff et al. (2004b) noted older equivalent of the Basal Khuff Clastics member in Saudi
how the upper Gharif member consists of four cycles (namely Arabia and marks the onset in Oman of the major transgression
cycles 5–8), the uppermost of which displays evidence of “tidal/ that ultimately took place across Arabia, pursuant to the creation
estuarine” environments. The preceding discussion has demon- of the Neotethys Ocean.
strated a similar change in Saudi Arabia from the widespread arid
continental setting in the Unayzah A member to the mixed fluvial CONCLUSIONS
to shallow-marine setting of the Basal Khuff Clastics member.
Furthermore, the Basal Khuff Clastics member exhibits a dis- 1. In middle Carboniferous times the Arabian plate was
tinctive and different heavy mineral assemblage relative to the subjected to an episode of major tectonic activity that was fol-
underlying sandstones of the Unayzah A (R.W.O’B. Knox, 2003, lowed by a period of erosion and the creation of the Hercynian
written commun.). It is thus evident that a significant depositional unconformity. These events were closely followed by the incep-
changeover exists between the Unayzah A member and the Basal tion of the late Paleozoic Gondwanan glaciation in Arabia, which
76 Melvin et al.

resulted in deposition of the Unayzah Formation upon the Her- stratal unit. Those rises in the paleo–water table possibly reflect
cynian unconformity. ongoing pulsed phases of a distant marine transgression, an idea
2. The Unayzah C member, where present, sits directly upon that is given support by the recognition of a heavily bioturbated
the Hercynian unconformity. It varies greatly in thickness and sandstone zone at the very top of the Unayzah A member in two
comprises quartzose sandstones that were laid down during mul- wells at either end of the study area. That zone is thought to rep-
tiple retreat phases of the ice sheets upon glaciofluvial outwash resent final breakthrough of the transgressing marine waters. In
braided plains. During subsequent intervening readvances of the most places the very highest parts of the Unayzah A display thick,
ice sheets, these outwash deposits were overridden, cannibal- well-developed paleosol horizons. These represent an extended
ized, and pushed into stacked piles of glacially tectonized (push period of minimal deposition and are considered to reflect the
moraine) deposits, manifest as relatively undeformed sandstones long period of uplift that was related to thermal doming prior
separated by distinct shear zones. The top of the Unayzah C to rifting and collapse that led to the creation of the pre-Khuff
member is an unconformity and represents the final subglacial unconformity. The Unayzah A member is considered equivalent
surface of the Gondwanan ice sheet in Saudi Arabia. to the middle Gharif member of Oman.
3. The terminal melt-out phase of the Gondwanan glacia- 6. The Unayzah A member is truncated by the pre-Khuff
tion is identified in the various facies of the Unayzah B member, unconformity, which marks the upper boundary of Tectono-
which was deposited in high southerly latitudes, ~75° S, based stratigraphic Megasequence TMS AP5 in Arabia. It is overlain by
on paleomagnetic data. These Unayzah B sediments comprise sandstones and shales of the Basal Khuff Clastics member of the
a number of different ice-proximal to ice-distal glaciolacustrine lowermost Khuff Formation. The depositional facies of the Basal
depositional facies. Facies associations consistently demonstrate Khuff Clastics are characterized in southeastern areas by shallow-
progressive retreat of the ice and a concomitant sustained increase marine sandstones and shales that pass westward into more
in flooding by meltwaters of an environment that was character- terrestrial facies (fluvial), reflecting the onset of a major westward-
ized by an abundance of glacial lakes. The top of the Unayzah B directed marine transgression. That transgression reached its full-
member can be taken to represent the climatic maximum flood- est development with the widespread marine carbonates of the
ing surface related to the melting of the Gondwanan ice sheets in Khuff Formation. It is an expression of the successful opening of
Saudi Arabia and is considered equivalent to the top of the Rahab the Neotethys Ocean, and as such represents a tectonically related
Shale in Oman. There is evidence to suggest a possible center of major flooding event.
relatively prolonged upland glaciation in the western part of the 7. The late Paleozoic Gondwanan glaciation per se in Saudi
study area, associated with the Al Batin Arch. Arabia is represented in part by the Hercynian unconformity, as
4. The unnamed middle Unayzah member is separated from well as by the deposition and deformation of the Unayzah C mem-
the underlying Unayzah B member by a sharp stratigraphic ber. Its final retreat from Saudi Arabia is seen in the rocks of the
boundary, considered to be a regional disconformity, which is Unayzah B member, the top of which represents the maximum
interpreted to represent a significant drainage event that marked glacial melt-out. This is the maximum expression of postglacial
the end of the glaciation in Saudi Arabia. Paleomagnetic data sug- transgression that can be ascribed solely to climatic (warming)
gest that the unnamed middle Unayzah member was laid down factors. The end of the glaciation in Saudi Arabia is marked by
in paleolatitudes ~55° S. This member comprises red-brown, an inferred dramatic drainage event, manifest in the disconfor-
alluvial floodplain sandy siltstones, within which occur various mity that separates the glaciolacustrine Unayzah B member from
isolated sandstone facies representing fluvial and possibly cold- the terrestrial deposits of the unnamed middle Unayzah member.
climate eolian deposits. It is commonly capped by a paleosol Thereafter, the rocks are continental in nature, and their relation-
horizon that represents a terrestrial highstand deposit, believed to ships with distant marine transgression (for which there is tan-
be equivalent to the marine Haushi Limestone highstand deposit talizing evidence) must necessarily incorporate the likelihood
at the top of the lower Gharif member in Oman. That those of significant tectonic influences. These tectonic influences are
paleosols represent a prolonged period of limited deposition or ultimately manifest in the pre-Khuff unconformity. Above that
nondeposition is supported by the paleomagnetic evidence. The unconformity the rocks of the Basal Khuff Clastics member of
paleosols are therefore considered to mark a significant discon- the Khuff Formation reflect the marine transgressive flooding
formity between the unnamed middle Unayzah member and the associated with the opening of the Neotethys Ocean.
overlying Unayzah A member.
5. The lower Unayzah A member was deposited in ACKNOWLEDGMENTS
widespread ephemeral (playa) lakes, whereas the upper Unayzah
A sediments were deposited predominantly within arid to semi- We acknowledge the Saudi Arabian Ministry of Petroleum
arid eolian dune fields and associated interdune deposits as well and Mineral Resources and the Saudi Arabian Oil Company
as within ephemeral stream systems and minor playa lakes. In (Saudi Aramco) for granting permission to publish this paper.
places the eolian deposits display an internal stratigraphy that is The evolution of our thoughts on the Unayzah Formation has
related to cyclical fluctuations in the paleo–water table, and which benefited from many discussions with many colleagues at Saudi
is correlatable within and between significant facies tracts in this Aramco. In particular we wish to acknowledge Mark Prudden,
Appendix: Core Log Legend
Planar lamination Mud cracks

Cross lamination
Burrows
Trough cross-
lamination
Rooted horizons
Low-angle lamination
Pedogenic “slickoliths”
Current ripples
Floating sand grains

Climbing ripples Mud clasts

Coarse sand Pebbles, cobbles


lamination
Wispy lamination Chert

Adhesion ripple Carbonaceous layer

Graded bedding Upward-shoaling


interval

Dish structure with Stylolites


elutriation pillars
Soft sediment High-angle stylolites
deformation
Flame structures Vertical fractures

Soft sediment Low-angle shear


loading

Grain size
G S Z M
Gravel Sand Silt Mud/clay
Figure A1. This figure provides a legend to which reference should be made for all figures throughout the paper compris-
ing sedimentological core logs.
78 Melvin et al.

Kent Norton, and Roger Price of Saudi Aramco’s Exploration Boulton, G.S., 1986, Push moraines and glacier-contact fans in marine and terres-
Organization for much stimulating debate. John Filatoff, Nigel trial environments: Sedimentology, v. 33, p. 677–698, doi: 10.1111/j.1365
-3091.1986.tb01969.x.
Hooker, and Merrell Miller provided biostratigraphic informa- Boulton, G.S., van der Meer, J.J.M., Beets, D.J., Hart, J.K., and Ruegg, G.H.J.,
tion, and Stephen Franks shared petrographic data regarding these 1999, The sedimentary and structural evolution of a recent push moraine
enigmatic rocks. Peter Sharland also provided valuable insights complex: Holmstrombreen, Spitsbergen: Quaternary Science Reviews,
v. 18, p. 339–371, doi: 10.1016/S0277-3791(98)00068-7.
on the occurrence of late-glacial drainage events during the Braakman, J.H., Levell, B.K., Martin, J.H., Potter, T.L., and van Vliet, A., 1982,
Pleistocene glaciation. Ernie Hailwood of CoreMagnetics pro- Late Palaeozoic Gondwana glaciation in Oman: Nature, v. 299, p. 48–50,
vided the analyses of the samples selected for the paleomagnetic doi: 10.1038/299048a0.
Broutin, J., Roger, J., Platel, J.-P., Angiolini, L., Baud, A., Bucher, H., Mar-
(paleolatitude) studies that are referenced herein. We nonetheless coux, J., and Al-Hashmi, H., 1995, The Permian Pangea. Phytographic
accept sole responsibility for the ideas presented in this work. We implications of new palaeontological discoveries in Oman (Arabian Pen-
thank George Grover and the Saudi Aramco Publications Review insula): Comptes Rendues Academie Scientifique Paris, t. 321, Serie IIa,
p. 1069–1086.
Committee for their time and effort, and for insightful comments Eberth, D.A., and Miall, A.D., 1991, Stratigraphy, sedimentology and evolution
in reviewing this paper. Kathleen Haughney of the Saudi Aramco of a vertebrate-bearing, braided to anastomosed fluvial system, Cutler For-
Exploration and Producing Information Center was her usual mation (Permian-Pennsylvanian), north-central New Mexico: Sedimen-
tary Geology, v. 72, p. 225–252, doi: 10.1016/0037-0738(91)90013-4.
cheerful self in chasing down some of the more elusive reference Evans, D.S., Bahabri, B.H., and Al-Otaibi, A.M., 1997, Stratigraphic trap in
material. Gene Cousart of Aramco’s Cartography Department the Permian Unayzah Formation, central Saudi Arabia: GeoArabia, v. 2,
demonstrated outstanding commitment and professionalism in p. 259–278.
Eyles, N., 1993, Earth’s Glacial Record and Its Tectonic Setting: Earth-Science
his approach to producing the figures that are presented in this Reviews, v. 35, 248 p., doi: 10.1016/0012-8252(93)90002-O.
paper. Ali Al-Zahrani and Hadi Al-Uraij are thanked for their tire- Faqira, M., Rademakers, M., and Afifi, A.M., 2009, New insights into the Her-
less efforts in arranging for the layout of many thousands of feet cynian Orogeny, and their implications for the Paleozoic hydrocarbon
system in the Arabian Plate: GeoArabia, v. 14, p. 199–228.
of Unayzah core in the Saudi Aramco Core Storage facility. Ferguson, G.S., and Chambers, T.M., 1991, Subsurface stratigraphy, deposi-
tional history, and reservoir development of the Early-to-Late Permian
REFERENCES CITED Unayzah Formation in central Saudi Arabia: Bahrain, Proceedings of the
Society of Petroleum Engineers (SPE) Middle East Oil Show, 7th, SPE
Paper 21394, p. 487–496.
Ahlbrandt, T.S., and Andrews, S., 1978, Distinctive sedimentary features of French, H.M., 1996, The Periglacial Environment (2nd edition): Harlow, UK,
cold-climate eolian deposits, North Park, Colorado: Palaeogeography, Addison Wesley, Longman, 341 p.
Palaeoclimatology, Palaeoecology, v. 25, p. 327–351, doi: 10.1016/0031 Fryberger, S.G., and Schenk, C.J., 1988, Pin stripe lamination: A distinctive
-0182(78)90048-2. feature of modern and ancient eolian sediments: Sedimentary Geology,
Aitken, J.F., Clark, N.D., Osterloff, P.L., Penney, R.A., and Mohiuddin, U., v. 55, p. 1–15, doi: 10.1016/0037-0738(88)90087-5.
2004, Regional core-based sedimentological review of the glacially Fryberger, S.G., Ahlbrandt, T.S., and Andrews, S., 1979, Origin, sedimentary
influenced Permo-Carboniferous Al-Khlata Formation, South Oman Salt features, and significance of low-angle eolian “sand sheet” deposits, Great
Basin, Oman: GeoArabia, Abstract, v. 9, no. 1, p. 16. Sand Dunes National Monument and vicinity, Colorado: Journal of Sedi-
Al-Belushi, J.D., Glennie, K.W., and Williams, B.P.J., 1996, Permo-Carboniferous mentary Petrology, v. 49, p. 733–746.
glaciogenic Al Khlata Formation, Oman: A new hypothesis for origin of Fryberger, S.G., Al-Sari, A.M., and Clisham, T.J., 1983, Eolian dune, interdune,
its glaciation: GeoArabia, v. 1, p. 389–404. sand sheet, and siliciclastic sabkha sediments of an offshore prograding
Al-Fares, A.A., Bouman, M., and Jeans, P., 1998, A new look at the middle sand sea, Dhahran area, Saudi Arabia: American Association of Petro-
to Lower Cretaceous stratigraphy, offshore Kuwait: GeoArabia, v. 3, leum Geologists Bulletin, v. 67, p. 280–312.
p. 543–560. Fryberger, S.G., Schenk, C.J., and Krystinik, L.F., 1988, Stokes surfaces and
Al-Hajri, S.A., and Owens, B., eds., 2000, Stratigraphic Palynology of the the effects of near-surface groundwater-table on aeolian deposition: Sedi-
Palaeozoic of Saudi Arabia: GeoArabia Special Publication 1, Gulf mentology, v. 35, p. 21–41, doi: 10.1111/j.1365-3091.1988.tb00903.x.
PetroLink, Bahrain, 231 p. Guit, F.A., Al-Lawati, M.H., and Nederlof, P.J.R., 1995, Seeking new potential
Al-Husseini, M.I., 2004, Pre-Unayzah unconformity, Saudi Arabia, in Al- in the Early–Late Permian Gharif play, west central Oman, in Al-Husseini,
Husseini, M.I., ed., Carboniferous, Permian and Early Triassic Arabian M.I., ed., Middle East Petroleum Geosciences GEO94: Gulf PetroLink,
Stratigraphy: GeoArabia Special Publication 3, Gulf PetroLink, Bahrain, Bahrain, v. 2, p. 447–462.
p. 15–59. Gustavson, T.C., Ashley, G.M., and Boothroyd, J.C., 1975, Depositional
Angiolini, L., Balini, M.E., Garzanti, E., Nicora, A., and Tintori, A., 2003, Gond- sequences in glaciolacustrine deltas, in Jopling, A.V., and McDonald,
wanan deglaciation and opening of Neotethys: The Al-Khlata and Sai- B.C., eds., Glaciofluvial and Glaciolacustrine Sedimentation: Society
wan Formations of Interior Oman: Palaeogeography, Palaeoclimatology, of Economic Paleontologists and Mineralogists Special Publication 23,
Palaeoecology, v. 196, p. 99–123, doi: 10.1016/S0031-0182(03)00315-8. p. 264–280.
Angiolini, L., Stephenson, M.H., and Leven, E.J., 2006, Correlation of the Helal, A.H., 1964, On the occurrence and stratigraphic position of Permo-
Lower Permian surface Saiwan Formation and subsurface Haushi lime- Carboniferous tillites in Saudi-Arabia: Geologische Rundschau, v. 54,
stone, Central Oman: GeoArabia, v. 11, p. 17–38. p. 193–207, doi: 10.1007/BF01821178.
Bates, R.L., and Jackson, J.A., 1980, Glossary of Geology (2nd edition): Falls Hughes Clarke, M.W., 1988, Stratigraphy and rock unit nomenclature in the oil
Church, Virginia, American Geological Institute, 751 p. producing area of interior Oman: Journal of Petroleum Geology, v. 11,
Bennett, M.R., 2001, The morphology, structural evolution and significance of p. 5–60, doi: 10.1111/j.1747-5457.1988.tb00800.x.
push moraines: Earth-Science Reviews, v. 53, p. 197–236, doi: 10.1016/ Hunter, R.E., 1980, Quasi-planar adhesion stratification—An eolian structure
S0012-8252(00)00039-8. formed in wet sand: Journal of Sedimentary Petrology, v. 50, p. 263–266.
Beydoun, Z.R., 1991, Arabian Plate Hydrocarbon Geology and Potential—A Husseini, M.I., 1992, Upper Palaeozoic tectono-sedimentary evolution of the
Plate Tectonic Approach: American Association of Petroleum Geologists Arabian and adjoining plates: Journal of the Geological Society [Lon-
Studies in Geology 33, 77 p. don], v. 149, p. 419–429, doi: 10.1144/gsjgs.149.3.0419.
Bishop, R.S., 1995, Maturation history of the Lower Palaeozoic of the Eastern Jopling, A.V., and Walker, R.G., 1968, Morphology and origin of ripple-drift
Arabian Platform, in Al-Husseini, M.I., ed., Middle East Petroleum Geo- cross-lamination with examples from the Pleistocene of Massachusetts:
sciences GEO94: Gulf PetroLink, Bahrain, v. 1, p. 180–189. Journal of Sedimentary Petrology, v. 38, p. 971–984.
Late Paleozoic Gondwanan glaciation in Saudi Arabia 79

Kellogg, K.S., Janjou, D., Minoux, L., and Fourniguet, J., 1986, Explanatory Osterloff, P., Penney, R., Aitken, J., Clark, N., and Husseini, M.I., 2004a, Depo-
notes to the geologic map of the Wadi Tathlith Quadrangle, Kingdom of sitional sequences of the Al Khlata Formation, subsurface Interior Oman,
Saudi Arabia: Deputy Minister for Mineral Resources, Ministry of Petro- in Al-Husseini, M.I., ed., Carboniferous, Permian and Early Triassic
leum and Mineral Resources, Kingdom of Saudi Arabia, 27 p., Geosci- Arabian Stratigraphy: GeoArabia Special Publication 3, Gulf PetroLink,
ence Map GM-103C, scale 1:250,000, sheet 20G. Bahrain, p. 61–81.
Kocurek, G., and Dott, R.H., Jr., 1981, Distinctions and uses of stratification Osterloff, P., Al-Harthy, A., Penney, R., Spaak, P., Williams, G., Al-Zadjali, F.,
types in the interpretation of eolian sand: Journal of Sedimentary Petrol- Jones, N., Knox, R., Stephenson, M., Oliver, G., and Al-Husseini, M.I.,
ogy, v. 51, p. 579–595. 2004b, Depositional sequences of the Gharif and Khuff Formations, sub-
Kocurek, G., and Fielder, G., 1982, Adhesion structures: Journal of Sedimen- surface Interior Oman, in Al-Husseini, M.I., ed., Carboniferous, Permian
tary Petrology, v. 52, p. 1229–1241. and Early Triassic Arabian Stratigraphy: GeoArabia Special Publica-
Konert, G., Al-Afifi, A.M., Al-Hajri, S.A., and Droste, H.J., 2001, Paleozoic tion 3, Gulf PetroLink, Bahrain, p. 83–147.
stratigraphy and hydrocarbon habitat of the Arabian Plate: GeoArabia, Ovenshine, A.T., 1970, Observations of iceberg rafting in Glacier Bay, Alaska,
v. 6, p. 407–442. and the identification of ancient ice-rafted deposits: Geological Society
Kruck, W., and Thiele, J., 1983, Late Palaeozoic glacial deposits in the Yemen of America Bulletin, v. 81, p. 891–894, doi: 10.1130/0016-7606(1970)81
Arab Republic: Geologisches Jahrbuch, Reihe B, Regionale Geologie [891:OOIRIG]2.0.CO;2.
Ausland, v. 46, p. 3–29. Roland, N.W., 1979, Geological Map of the Arab Yemen Republic, Sheet
Levell, B.K., Braakman, J.H., and Rutten, K.W., 1988, Oil-bearing sediments Sadah, 1:250,000: Hanover, Germany, Federal Institute of Geoscience
of Gondwana glaciation in Oman: American Association of Petroleum and Natural Resources.
Geologists Bulletin, v. 72, p. 775–796. Ruegg, G.H.J., 1983, Periglacial eolian evenly laminated sandy deposits in
Loosveld, R.J.H., Bell, A., and Terken, J.J.M., 1996, The tectonic evolution of the Late Pleistocene of NW Europe, a facies unrecorded in modern sedi-
interior Oman: GeoArabia, v. 1, p. 28–51. mentological handbooks, in Brookfield, M.E., and Ahlbrandt, T.S., eds.,
Lowe, D.R., and LoPiccolo, R.D., 1974, The characteristics and origins of Eolian Sediments and Processes: New York, Elsevier, Developments in
dish and pillar structures: Journal of Sedimentary Petrology, v. 44, Sedimentology 38, p. 455–482.
p. 484–501. Rust, B.R., and Romanelli, R., 1975, Late Quaternary subaqueous outwash
Mangerud, J., Jakobsson, M., Alexanderson, H., Astakhov, V., Clarke, G.K.C., deposits near Ottawa, Canada, in Jopling, A.V., and McDonald, B.C., eds.,
Henriksen, M., Hjort, C., Krinner, G., Lunkka, J.-P., Moller, P., Murray, Glaciofluvial and Glaciolacustrine Sedimentation: Society of Economic
A., Nikolskaya, O., Saarnisto, M., and Svendsen, J.I., 2004, Ice-dammed Paleontologists and Mineralogists Special Publication 23, p. 177–192.
lakes and rerouting of the drainage of northern Eurasia during the Last Gla- Senalp, M., and Al-Duaiji, A., 1995, Stratigraphy and sedimentation of the
ciation: Quaternary Science Reviews, v. 23, p. 1313–1332, doi: 10.1016/ Unayzah reservoir, central Saudi Arabia, in Al-Husseini, M.I., ed., Middle
j.quascirev.2003.12.009. East Petroleum Geosciences Conference, GEO’94: Gulf PetroLink, Bah-
Martini, I.P., and Brookfield, M.E., 1995, Sequence analysis of Upper Pleis- rain, v. 2, p. 837–847.
tocene (Wisconsinan) glaciolacustrine deposits of the north-shore bluffs Sharland, P.R., Archer, R., Casey, D.M., Davies, R.B., Hall, S.H., Heward,
A.P., Horbury, A.D., and Simmons, M.D., 2001, Arabian Plate Sequence
of Lake Ontario, Canada: Journal of Sedimentary Research, v. 65,
Stratigraphy: GeoArabia Special Publication 2, Gulf PetroLink, Bahrain,
p. 388–400.
371 p., with 3 charts.
McClure, H.A., 1980, Permian–Carboniferous glaciation in the Arabian penin-
Stampfli, G.M., and Borel, G.D., 2004, The TRANSMED transects in space
sula: Geological Society of America Bulletin, v. 91, p. 707–712, doi: 10
and time: Constraints on the paleotectonic evolution of the Mediterranean
.1130/0016-7606(1980)91<707:PGITAP>2.0.CO;2.
domain, in Cavazza, W., Roure, F., Spakman, W., Stampfli, G.M., and
McClure, H.A., and Young, G.M., 1981, Late Paleozoic glaciation in the
Ziegler, P.A., eds., The TRANSMED Atlas—The Mediterranean Region
Arabian peninsula, in Hambrey, M.J., and Harland, W.B., eds., Earth’s
from Crust to Mantle: Berlin, Heidelberg, Springer, p. 53–80.
Pre-Pleistocene Glacial Record: Cambridge, UK, Cambridge University Stephenson, M.H., 1998, Preliminary correlation of palynological assemblages
Press, p. 275–277. from Oman with the Granulatisporites confluens Oppel Zone of the Grant
McClure, H.A., Hussey, E.M., and Kaill, I.J., 1988, Permian-Carboniferous Formation (Lower Permian), Canning Basin, Western Australia: Jour-
glacial deposits in southern Saudi Arabia: Geologisches Jahrbuch, Reihe nal of African Earth Sciences, v. 26, p. 521–526, doi: 10.1016/S0899
B, Regionale Geologie Ausland, v. 68, p. 3–31. -5362(98)00030-X.
McGillivray, J.G., and Husseini, M.I., 1992, The Palaeozoic petroleum geology Stephenson, M.H., 2004, Early Permian spores from Oman and Saudi Ara-
of central Arabia: American Association of Petroleum Geologists Bulle- bia, in Al-Husseini, M.I., ed., Carboniferous, Permian and Early Triassic
tin, v. 76, p. 1473–1490. Arabian Stratigraphy: GeoArabia Special Publication 3, Gulf PetroLink,
Melvin, J., and Heine, C.J., 2004, Sequence stratigraphy of an eolian gas sand: Bahrain, p. 185–215.
Layering in the Permian Unayzah-A reservoir at south Ghawar, Eastern Stephenson, M.H., 2006, Stratigraphic note: Update of the standard Arabian
Saudi Arabia: GeoArabia, Abstract, v. 9, p. 103. Permian palynological biozonation; definition and description of OSPZ5
Melvin, J., and Sprague, R.A., 2006, Advances in Arabian stratigraphy: Ori- and 6: GeoArabia, v. 11, p. 173–178.
gin and stratigraphic architecture of glaciogenic sediments in Permian- Stephenson, M.H., and Filatoff, J., 2000a, Correlation of Carboniferous-
Carboniferous lower Unayzah sandstones, eastern central Saudi Arabia: Permian palynological assemblages from Oman and Saudi Arabia, in Al-
GeoArabia, v. 11, p. 105–152. Hajri, S., and Owens, B., eds., Stratigraphic Palynology of the Palaeozoic
Melvin, J., Sprague, R.A., and Heine, C.J., 2005, Diamictites to eolianites: of Saudi Arabia: GeoArabia Special Publication 1, Gulf PetroLink, Bah-
Carboniferous–Permian climate change seen in subsurface cores from rain, p. 168–191.
the Unayzah Formation, east-central Saudi Arabia, in Reinson, G.E., Stephenson, M.H., and Filatoff, J., 2000b, Description and correlation of Late
Hills, D., and Eliuk, L., eds., 2005 CSPG Core Conference Papers and Permian palynological assemblages from the Khuff Formation, Saudi
Extended Abstracts CD: Calgary, Canadian Society of Petroleum Geolo- Arabia and evidence for the duration of the pre-Khuff hiatus, in Al-Hajri,
gists, p. 237–282. S., and Owens, B., eds., Stratigraphic Palynology of the Palaeozoic of
Melvin, J., Wallick, B.P., and Heine, C.J., 2010, Advances in Arabian stratig- Saudi Arabia: GeoArabia Special Publication 1, Gulf PetroLink, Bahrain,
raphy: Allostratigraphic layering related to paleo–water table fluctuations p. 192–215.
in eolian sandstones of the Permian Unayzah A reservoir, South Haradh, Stephenson, M.H., and Osterloff, P.L., 2002, Palynology of the deglaciation
Saudi Arabia: GeoArabia, v. 15, p. 55–86. sequence represented by the Lower Permian Rahab and Lower Gharif
Moncrieff, A.C.M., and Hambrey, M.J., 1990, Marginal-marine glacial sedi- members, Oman: American Association of Stratigraphic Palynologists
mentation in the late Precambrian succession of East Greenland, in Contribution Series, v. 40, p. 1–32.
Dowdeswell, J.A., and Scourse, J.D., eds., Glacimarine Environments: Stephenson, M.H., Osterloff, P.L., and Filatoff, J., 2003, Palynological biozona-
Processes and Sediments: Geological Society [London] Special Publica- tion of the Permian of Oman and Saudi Arabia: Progress and challenges:
tion 53, p. 387–410. GeoArabia, v. 8, p. 467–496.
Mountney, N.P., and Jagger, A., 2004, Stratigraphic evolution of an aeolian erg Stokes, W.L., 1968, Multiple parallel-truncation bedding planes—A feature of
margin system: The Permian Cedar Mesa Sandstone, SE Utah, USA: Sed- wind-deposited sandstone formations: Journal of Sedimentary Petrology,
imentology, v. 51, p. 713–743, doi: 10.1111/j.1365-3091.2004.00646.x. v. 38, p. 510–515.
80 Melvin et al.

Teller, J.T., Leverington, D.W., and Mann, J.D., 2002, Freshwater outbursts to Vaslet, D., Le Nindre, Y.-M., Vachard, D., Broutin, J., Crasquin-Soleau, S.,
the oceans from glacial Lake Agassiz and their role in climate change dur- Berthelin, M., Gaillot, J., Halawani, M., and Al-Husseini, M.I., 2005, The
ing the last deglaciation: Quaternary Science Reviews, v. 21, p. 879–887, Permian-Triassic Khuff Formation of central Saudi Arabia: GeoArabia,
doi: 10.1016/S0277-3791(01)00145-7. v. 10, p. 77–134.
Van der Wateren, D.F.M., 1985, A model of glacial tectonics, applied to the ice- Visser, J.N.J., 1982, Upper Carboniferous glacial sedimentation in the Karoo
pushed ridges in the Central Netherlands: Geological Society of Denmark basin near Prieska, South Africa: Palaeogeography, Palaeoclimatology,
Bulletin, v. 34, p. 55–74. Palaeoecology, v. 38, p. 63–92, doi: 10.1016/0031-0182(82)90065-7.
Van der Wateren, D.F.M., 1987, Structural geology and sedimentology of the Wender, L.E., Bryant, J.W., Dickens, M.F., Neville, A.S., and Al-Moqbel, A.M.,
Dammer Berg push moraine, FRG, in Meer, J.J.M., ed., Tills and Glacio- 1998, Paleozoic (pre-Khuff) hydrocarbon geology of the Ghawar area,
tectonics: Rotterdam, A.A. Balkema, p. 157–182. eastern Saudi Arabia: GeoArabia, v. 3, p. 273–302.
Van der Wateren, D.F.M., 1994, Proglacial subaquatic outwash fan and delta Williams, B.P.J., Wild, E.K., and Suttill, R.J., 1985, Paraglacial aeolianites:
sediments in push moraines—Indicators of subglacial meltwater activ- Potential new hydrocarbon reservoirs, Gidgealpa Group, southern Cooper
ity: Sedimentary Geology, v. 91, p. 145–172, doi: 10.1016/0037-0738(94 Basin: Australian Petroleum and Exploration Association Journal, v. 25,
)90127-9. p. 291–310.
Van der Wateren, D.F.M., 1995, Structural geology and sedimentology of push
moraines: Processes of soft sediment deformation in a glacial environ-
ment and the distribution of glaciotectonic styles: Mededelingen Rijks
Geologische, v. 54, 167 p. MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

Environmental and paleogeographic implications of glaciotectonic


deformation of glaciomarine deposits within Permian strata of the
Metschel Tillite, southern Victoria Land, Antarctica

John L. Isbell*
Department of Geosciences, University of Wisconsin–Milwaukee, Milwaukee, Wisconsin 53021, USA

ABSTRACT

Popular reconstructions of late Paleozoic glaciation depict a single massive ice


sheet centered over Victoria Land and extending outward over much of Gondwana.
This view is untenable, as interpretations presented here indicate that glaciogenic
strata in southern Victoria Land were deposited in a glaciomarine setting, and that
ice entered the area from at least two different ice centers on opposite sides of the
depositional basin. Reports from other areas also reveal that multiple ice sheets, ice
caps, and alpine glaciers diachronously waxed and waned across Gondwana during
the Carboniferous and Permian.
Glaciogenic rocks of the Lower Permian Metschel Tillite contain glaciotectonic
structures and glaciogenic deposits that include (1) sheared diamictites, (2) thrust
sheets, (3) massive and stratified diamictites, and (4) sheet sandstones. These features
formed as subglacial deforming beds, thrust moraines at glacial termini, and as gla-
ciomarine deposits associated with temperate glaciers. A glaciomarine setting, rather
than a glaciolacustrine setting, is suggested, owing to the abundance of meltwater
plume deposits. A wedge-shaped sandstone body at the base of the overlying Weller
Coal Measures was deposited as a grounding-line fan. Results of this study imply
deposition in ice-marginal glaciomarine settings from ice radiating out of multiple
glacial centers. These findings are significant because multiple glaciers, covering
a given area, contain considerably less ice volume than a single massive ice sheet.
Therefore, the waxing and waning of multiple ice masses during the late Paleozoic
would have influenced global climate and eustatic sea level much differently than
would have a single massive Gondwanan ice sheet.

INTRODUCTION Permian (Fig. 1; Veevers, 1994, 2001; Zeigler et al., 1997; Hyde
et al., 1999; Scotese et al., 1999). In these models, ice flowed radi-
Many paleogeographic reconstructions of the late Paleozoic ally outward from a glacial spreading center over Victoria Land,
Gondwanan Ice Age (LPGIA) depict an immense ice sheet cov- Antarctica, and extended, in one direction, to glaciomarine mar-
ering Gondwana during the Mississippian, Pennsylvanian, and gins in the Ellsworth Mountains and South Africa (Lindsay, 1970;
Barrett, 1991; Veevers, 2001). In the opposite direction, models
*[email protected] show ice flowing out of northern Victoria Land to terrestrial ice

Isbell, J.L., 2010, Environmental and paleogeographic implications of glaciotectonic deformation of glaciomarine deposits within Permian strata of the Metschel
Tillite, southern Victoria Land, Antarctica, in López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in
Gondwana: Geological Society of America Special Paper 468, p. 81–100, doi: 10.1130/2010.2468(03). For permission to copy, contact [email protected].
©2010 The Geological Society of America. All rights reserved.

81
82 Isbell

A. Africa India
S. America

Au
str
ali
a
Antarctica
CTM VL
Figure 1. (A) Map showing the distribution of
glacial basins in Gondwana. (B) Gondwana re-
EM

construction after Powell and Li (1994), showing


the location of a hypothetical ice sheet covering
Gondwana (modified from Ziegler et al., 1997)
NZ with an ice spreading center over Victoria Land,
Antarctica (Lindsay, 1970; Barrett, 1991). Hypo-
thetical flow directions are from models proposed
by Lindsay (1970), Crowell and Frakes (1971),
B. Barrett (1991), and Veevers (2001). CTM—central
Transantarctic Mountains, EM—Ellsworth Moun-
0 km 1500 tains, NZ—New Zealand, VL—Victoria Land.

Antarctica
VL
CTM
EM
Gondwana
Glacial Victoria Land
Basins NZ
Ice Spreading
Center
Gondwanan Ice Sheet 0 km 1500

margins in southern Australia and glaciomarine margins in Tas- deposits occurred prior to deposition of the overlying Lower
mania (Fig. 1; Crowell and Frakes, 1971; Barrett, 1991; Veevers, Permian Weller Coal Measures. The occurrence of the glacial
2001). Throughout Antarctica, early workers identified upper valleys and regional thinning of the glacial strata toward Victo-
Paleozoic diamictites as tillites deposited subglacially from the ria Land led Lindsay (1970) and Barrett (1991) to conclude that
terrestrial portion of the ice sheet as it waxed and waned across southern Victoria Land lay adjacent to a late Paleozoic glacial
Gondwana (e.g., Lindsay, 1970; Barrett and Kyle, 1975; Barrett spreading center and that deposition occurred from the terres-
et al., 1986). The occurrence of an ice spreading center in Victo- trial portion of the ice sheet (Fig. 1B). However, data presented
ria Land supplying ice to the massive Gondwanan Ice Sheet is herein suggest that some of the relief on the preglacial contact
central to many paleogeographic and climatic models for the late separating upper Paleozoic glacial rocks above from Devonian
Paleozoic. These models are based on work conducted prior to strata below resulted from glaciotectonic thrusting along concave
the advent of modern glacial facies models, and therefore their upward shear planes rather than from erosion and development of
validity has not been rigorously tested in Victoria Land, the pro- glacial valleys, and that deposition occurred within a glacioma-
posed terrestrial center of the ice sheet. rine setting.
In Antarctica (Fig. 2), upper Paleozoic glacial deposits thin Even though glaciotectonic structures and their significance
from the central Transantarctic Mountains into southern Victo- have been recognized in Gondwanan rocks in South Africa and
ria Land. There, glaciogenic deposits are discontinuous (Bar- South America (e.g., Visser and Loock, 1982; Rocha-Campos
rett and Kyle, 1975; Collinson et al., 1994). In southern Victoria et al., 2000), descriptions of pre-Pleistocene glaciotectonic fea-
Land, thick glaciogenic deposits are reported as glacial terrestrial tures are scarce. Deformed upper Paleozoic glaciogenic rocks
valley fills (Barrett, 1972; Barrett and Kyle, 1975; McKelvey also occur in Antarctica. The identification and interpretation of
et al., 1977; Barrett and McKelvey, 1981). Barrett and McKel- these strata are enigmatic, as deformation has been attributed to
vey (1981) suggested that the valleys preferentially preserved overriding of the sediments by Late Carboniferous or Permian
the deposits from erosion during postglacial rebound, whereas, ice, slumping along glacial valley walls, and/or owing to tec-
outside the valleys, partial if not complete erosion of the glacial tonic upheaval (Barrett, 1972; Barrett and Kyle, 1975; McKelvey
Permian deposits, southern Victoria Land, Antarctica 83

A B

Figure 2. (A) Map of Antarctic sites showing the location of southern Victoria Land. (B) Location map of southern Victoria Land showing Kennar
Valley, Mount Metschel, and Mount Ritchie.

et al., 1977; Barrett and McKelvey, 1981; McElroy and Rose,


1987; Spörli, 1992).
This paper presents sedimentologic data collected from
Mount Metschel, Mount Ritchie, and Kennar Valley in southern
Victoria Land (Fig. 2) during the 2000–2001 and 2001–2002
austral field seasons. Here, strata of the Devonian Aztec Silt-
stone, the Permian Metschel Tillite, and the basal few meters
of the Weller Coal Measures (Fig. 3) are described and inter-
preted (1) to determine whether a major late Paleozoic ice sheet
spreading center was located in Victoria Land, (2) to determine
the setting in which glaciogenic sediments were deposited (i.e., Figure 3. Generalized stratigraphic section of Devonian and Permian
subglacial, periglacial, glaciomarine), and (3) to document the rocks exposed in southern Victoria Land, Antarctica.
occurrence of glaciotectonic structures within upper Paleozoic
strata in Antarctica. Understanding upper Paleozoic glaciogenic
strata in southern Victoria Land will help to resolve the nature,
timing, and extent of the LPGIA, which is of importance, as this an overview of the occurrence and distribution of Devonian and
glacial interval is one of the best analogues for predicting what upper Paleozoic rocks in this area.
will happen to Earth systems during the transition out of the cur- The Aztec Siltstone is up to 217 m thick and consists of inter-
rent Cenozoic Ice Age. bedded shale, siltstone, and cross-bedded sandstone (Fig. 3). Silt-
stones and shales are a few centimeters to several meters thick,
STRATIGRAPHY whereas sandstones vary from 0.1 to 15 m in thickness (McKel-
vey et al., 1977; McElroy and Rose, 1987). The occurrence of
Glaciotectonic structures and glaciogenic deposits described fining-upward cycles, red beds, rootlet horizons, mud cracks, and
in this paper are found at the top of the Devonian Aztec Siltstone, conchostracan fossils all suggest that deposition occurred in an
in the Lower Permian Metschel Tillite, and at the base of the alluvial setting (McKelvey et al., 1977; McPherson, 1978, 1979).
Lower Permian Weller Coal Measures in southern Victoria Land A microflora containing Geminospora lemurata and fossil fish
(Fig. 3). The rocks are part of the Taylor (Devonian) and Victoria near the top of the unit, including Bothriolepis, Groenlandaspis,
(Permian and Triassic) Groups of the Beacon Supergroup. Barrett and Turinia gondwana, indicate a Middle to Late Devonian age
(1972), Barrett and Kyle (1975), McKelvey et al. (1977), Barrett (Helby and McElroy, 1969; McKelvey et al., 1972; Ritchie, 1975;
and McKelvey (1981), and McElroy and Rose (1987) provide Young, 1988, 1989, 1991; Turner and Young, 1992).
84 Isbell

Glaciogenic strata of the Metschel Tillite unconformably gently dipping (12°–24°) sandstone, mudstone, and paleosols in
overlie rocks of the Aztec Siltstone (Fig. 3). The Metschel Tillite the Aztec Siltstone that are internally folded (folding may be the
is 0–85 m thick and consists of diamictite, conglomerate, sand- result of numerous microscopic thrust faults) and cut by thrust
stone, and shale, which are locally intraformationally deformed faults with centimeter-scale displacements; Fig. 4A); (3) 2 m of
(Barrett and Kyle, 1975; McKelvey et al., 1977; Barrett and pervasively thrust faulted Aztec Siltstone where individual sand-
McKelvey, 1981; McElroy and Rose, 1987). Although the age of stone beds are internally folded and thrust faulted (Fig. 4B); (4) a
the Metschel Tillite is unknown, it is here considered to be Early sharp contact that truncates underlying folded and faulted rocks
Permian on the basis of its position below rocks of the Weller of the Aztec Siltstone below from diamictites of the Metschel
Coal Measures, which contain Lower Permian palynomorphs, Tillite above; (5) 1.5 m of thrust faulted diamictite consisting
and also on the basis of correlation of the Metschel Tillite with of a chaotic mixture of sandstone pods and boudins, including
glaciogenic rocks of the Darwin Tillite exposed 130 km to the material from the Aztec Siltstone, and granite and quartz clasts
south. The Darwin Tillite contains Asselian-Sakmarian palyno- up to 0.3 m in diameter; (6) 2–3 m of pervasively foliated (dips
morphs, as do all other palynomorph-bearing upper Paleozoic of 8°–30°), chaotic to homogenized diamictite containing granite
glaciogenic rocks in Antarctica (cf. Barrett and Kyle, 1975; Kyle, and quartz clasts up to 0.2 m in diameter (Fig. 4C); and (7) locally
1977; Kyle and Schopf, 1982; Lindström, 1995; Askin, 1998). up to 2 m of almost completely homogenized foliated diamic-
Strata of the Weller Coal Measures rest both disconformably tite containing granite, and quartz clasts up to 0.2 m in diameter
and conformably on diamictites of the Metschel Tillite (Fig. 3; (Fig. 4D). Foliation, thrust faults, and axial planes of folds within
McKelvey et al., 1977). Conglomerate and cross-bedded, coarse- the deformed zone dip at up to 44° toward 306°. At the top of
grained sandstone containing quartz pebbles occur at the base of the deformed zone a sharp contact separates highly foliated and
the formation, whereas upward within the unit, fine to coarse- homogenized diamictites below from undeformed, weakly strati-
grained sandstones are interstratified with siltstones, shales, and fied to stratified diamictite above (Fig. 4E).
coals. Strata near the base of the formation contain fossil leaf A similar but much thinner deformed zone occurs on the
impressions of Gangamopteris, Glossopteris, and Cordaites east side of Mount Ritchie near the top of the Metschel Tillite.
(Kyle, 1976; Pyne, 1984; Cúneo et al., 1993). Kyle (1977) cor- There, weakly stratified and stratified diamictite are overlain
related the Weller Coal Measures with Lower Permian (Artin- by a 1-m-thick sandy interval containing numerous small-scale
skian) rocks in Australia (Australian Stage 4 Palynomorph Zone thrust faults, and an overlying 1-m-thick foliated diamictite. The
of Evans, 1969) based on microfloras from the middle and upper small-scale thrust faults dip at up to 45° toward 243°. This foli-
parts of the formation. ated diamictite is overlain by a sharp contact with weakly strati-
fied diamictite or by an erosional surface that separates strata of
GLACIOGENIC DEPOSITS AND GLACIOTECTONIC the Metschel Tillite below from conglomerates and cross-bedded
STRUCTURES sandstones of the Weller Coal Measures above.

Two types of deformed and two types of undeformed facies Interpretation


associations within Devonian and upper Paleozoic strata are The presence of small-scale thrust faults, chaotic bedding–
reported in this paper. The associations are (1) sheared and homog- boudinage structures, and foliated homogenized materials at
enized sandstone and diamictite, (2) large-scale thrust sheets, Mount Metschel suggest that sandstone and shale at the top of
(3) diamictite, and (4) sheet sandstone facies associations. the Aztec Siltstone and diamictite at the base of the Metschel
Tillite were deformed by shear. In general, deformation within
Sheared and Homogenized Sandstone and Diamictite
Facies Association

Description Figure 4. Stratigraphic column of deformed strata of the Devonian


On Mount Metschel (Fig. 2) an asymmetric deformation Aztec Siltstone and Permian Metschel Tillite exposed on the east-
zone separates the 50(+)-m-thick Devonian Aztec Siltstone from facing slopes of Mount Metschel, interpreted to be the deposits of a
the overlying 21.5-m-thick Permian Metschel Tillite (Fig. 4). At subglacial deformation bed (glaciotectonite and deformation till). Ori-
entations of structural features and interpreted ice-flow directions are
this site the Aztec Siltstone consists of 0.1–0.5-m-thick interbeds also shown. (A) Thrust faults with centimeter-scale displacements cut
of cross-laminated and horizontally laminated, fine- to medium- strata of the Aztec Siltstone. Brunton compass for scale. (B) Perva-
grained sandstone, gray to red mudstone, and calcrete-bearing sively thrust faulted sandstone at the top of the Aztec Siltstone. Rock
paleosols. Near the top of the unit, strata pass upward through hammer for scale. (C) Pervasively foliated diamictite and sandstone
a deformed zone, which shows increasing strain upward, into near the base of the Metschel Tillite. Ice axe for scale. (D) Pervasively
foliated and partially homogenized diamictite near the base of the
foliated diamictite of the Metschel Tillite (Fig. 4). The follow- Metschel Tillite. Ice axe for scale. (E) Stratified diamictite with lobe-
ing zones occur from the base to the top of this sequence, which shaped, boulder-bearing sandy diamictite near the top of the Metschel
straddles the Aztec-Metschel contact: (1) undeformed interbeds Tillite at Mount Metschel. Ice axe for scale. vm—vector mean, mvl—
of sandstone and mudstone of the Aztec Siltstone; (2) 2–2.5 m of mean vector length, n—number of measurements.
E Stratified diamictite D

Dipping
foliation
Lobe-shaped sandy diamictite

n=11
Ice flow direction C
(c)
Dip of shear
planes vm=126
mvl=0.95
Diamictite

n=11
Clay

Sand
Silt

f mc
14m
(d)
Undeformed stratified
diamictite with granite
and quartz clasts up to
B
1 m in diameter
Metschel Tillite

Homogenized diamictite with


granite and quartz clasts up to
0.2 m in diameter

Pervasively foliated, diamictite


with granite and quartz clasts
up to 0.2 m in diameter
Thrust faults
Thrust faulted diamictite with sand
boudins and granite and quartz
clasts up to 0.3 m in diameter
Sharp Contact A
Pervasively thrust faulted (dm
to m displacements) sandstone
Aztec Siltstone

that is internally folded and Small-scale


thrust faulted thrust faults

Internally folded and thrust


faulted with cm displacements
Small-scale
thrust faults
Undeformed Aztec
Siltstone
0
86 Isbell

individual tectonic and subglacial shear zones displays a sym- of the upper Midwestern United States are weakly consolidated
metrical increase in strain away from upper and lower bound- and were locally deformed by glaciogenic activity during the
aries toward a homogenized central region of high finite shear Pleistocene. The interstratification of sands and muds in the
strain (Van der Wateren, 1987, 2002). The distribution of strain Aztec Siltstone likely inhibited subglacial drainage into underly-
in the shear zone at Mount Metschel is asymmetrical, with ing aquifers, possibly resulting in high pore-water pressures and
changes in structures indicating that strain increased upward the formation of a subglacial deforming bed during overriding of
from undeformed rock in the Aztec Siltstone into the overly- the area by a Permian glacier. The orientation of the thrust faults
ing homogenized diamictite at the base of the Metschel Tillite. and shear planes at Mount Metschel indicate local ice move-
However, structures indicating decreasing strain do not occur as ment toward 126°. Barrett and Kyle (1975) reported striations at
the homogenized diamictite passes abruptly upward into unde- Mount Metschel oriented toward 120°. However, these striations
formed stratified diamictite. were likely slickensides contained within the shear zone.
Rocks at the contact between the Aztec Siltstone and the The deformed zone at the top of the Metschel Tillite on the
Metschel Tillite are similar to sediment deformed and deposited east-facing slope of Mount Ritchie is also interpreted to have
subglacially. Under certain conditions, shear between glacial ice formed subglacially as a glaciotectonite and deformation till. The
and its unconsolidated or weakly consolidated substrate results in occurrence of weakly stratified and stratified diamictite (see sec-
formation of a deforming bed beneath the ice-sediment interface tion below on Diamictite Facies Association) directly below this
(Alley, 1991; Benn and Evans, 1998). For ice streams and surg- unit, coupled with the orientation of the structural fabric within
ing glaciers, much of the forward motion of the glacier may be the deformed zone, suggests that grounded ice advanced north-
accounted for within such a unit. Formation of a deforming bed is eastward (063°) into a glaciomarine setting.
favored by the occurrence of (1) subglacial waters, (2) unfrozen
unconsolidated or weakly consolidated (sedimentary) substrates, Large-Scale Thrust Sheet Facies Association
and (3) substrate materials that inhibit water from draining away
into the underlying sediments (Alley, 1991; Boulton, 1996). Description of Strata at Mount Richie
Under these conditions, glacial ice forms the “mirrored top” At Mount Ritchie (Fig. 2), thrust sheets in the Metschel
of the idealized symmetrical shear zone with strain increasing Tillite, up to 15 m thick, occur at and just above the contact with
upward within the substrate toward the ice-sediment interface undeformed strata of the Devonian Aztec Siltstone (Fig. 5). The
and then decreasing away from the interface into the overlying sheets, which are imbricately stacked, make up a 600+-m-long
glacier (Boulton, 1979; Boulton and Jones, 1979; Boulton and duplex along a north-northwest– to south-southeast–trending
Hindmarsh, 1987; Van der Wateren, 2002). A vertical strain pro- ridge near the summit of the mountain. These strata are exposed
file within a subglacial deforming bed consists of (1) undeformed on the west-facing slope of Mount Richie. Along much of the
sediment at depth, (2) materials deformed by brittle failure and ridge the contact between the thrust sheets and underlying strata
faulting, (3) pervasive ductile shearing of materials, and (4) plow- is covered by snow and scree. Where exposed, the substratum
ing and sliding of ice and debris along the ice-sediment contact includes (1) diamictite, (2) deformed interstratified diamictite
(Banham, 1977; Boulton, 1987; Alley, 1991; Hart and Boulton, and sandstone, (3) breccia, and (4) strata of the Aztec Siltstone
1991; Benn, 1995; Benn and Evans, 1996; Van der Wateren, (Figs. 5A, 5B, 5C). At its northern end the duplex ramps up onto
2002). A décollement may separate structural zones display- a 4–6-m-thick, fine- to medium-grained massive sandstone (see
ing different degrees of strain (Banham, 1977; Boulton, 1987). the section below on Sandstone Sheet Facies Association) in the
Subglacially sheared deposits are classified by Benn and Evans Metschel Tillite. There, individual thrust sheets are bounded by
(1996) as glaciotectonites if the deposits contain structural char-
acteristics of the original parent material, or as deformation tills if
the material has been homogenized. Recently, van der Meer et al.
(2003) reported that all subglacial tills form from a combination Figure 5. (A) Photo of imbricate thrust sheets exposed on the west-
of both deformation and lodgment. southwest face of Mount Ritchie, interpreted to be part of a subaquatic
At Mount Metschel, deformed Aztec Siltstone has all of the thrust moraine complex. For scale, the distance from the base of the
characteristics of a glaciotectonite, whereas the overlying foli- Weller Coal Measures to the base of the sill is 25 m. (B) Interpretive
map of the thrust sheets on the west-southwest face of Mount Ritchie.
ated diamictites at the base of the Metschel Tillite display char- Stereonet shows the orientation of the thrust sheets, and the rose dia-
acteristics of a deformation till. The two formations are separated gram shows paleo ice-flow directions interpreted from the orientations
by a décollement, which would have served as the structural of the thrust sheets. (C) Brecciated zone marking the location of a shear
boundary between brittle and ductile deformation. The décolle- plane at the base of a thrust sheet on Mount Ritchie; 15-cm-long ruler
ment also occurs at the position of the regional unconformity that for scale. (D) Thrust sheets of the Metschel Tillite, composed of strata
of the Aztec Siltstone at Mount Ritchie. The sheets are bounded by
now separates the two formations. The occurrence of this type of major thrust faults (MTF), whereas recumbent folds and small-scale
deformation indicates that rocks of the Upper Devonian Aztec thrust faults (STF) with meter-scale displacements occur within the
Siltstone were weakly consolidated during glaciation. Such an sheets. Person and Jacob’s staff (1.6 m) for scale. vm—vector mean,
interpretation is tenable, as Cambrian and Ordovician sandstones mvl—mean vector length, n—number of measurements.
A

C D
88 Isbell

listric to sigmoidal-shaped shear zones that consist of either a 1–7 m thick consist of (1) weakly stratified diamictite; (2) inter-
0.05–0.3-m-thick, intensely foliated mudstone containing rare bedded conglomerate and faulted, cross-bedded to massive
granite and gneiss granules, pebbles, and cobbles or locally by sandstone containing dish structures and dewatering pipes; and
a 0.5–3-m-thick brecciated and/or chaotically folded zone. The (3) coarsening-upward siltstone to sandstone successions with
thrust sheets are sigmoidal in shape and consist of massive sand- siltstones containing isolated ripples (lenticular bedding) and
stone, which contains an internal mosaic of annealed fractures sandstones containing cross-laminae, horizontal laminations,
and dewatering structures. The sheets within this complex dip at wave ripple laminae, swaley cross-stratification, load and flame
up to 45° toward 127°. Near the back of the thrust complex, sev- structures, and scattered dropstones.
eral thrust sheets consist of an interbedded succession of cross- The contact with the underlying Aztec Siltstone is cov-
stratified and horizontally laminated sandstones and mudstones ered by scree. However, some thrust sheets cut below the level
of the Aztec Siltstone (Figs. 5A, 5B, 5D). These sheets rest on of the contact (Figs. 7A and 7B). Undeformed conglomerates
the backs of the thrust sheets described above. Internally, the and sandstones of the Weller Coal Measures rest on an erosion
Aztec thrust sheets contain internal deformational structures that surface with a relief of several meters cut into the underlying
include (1) high-angle normal faults, (2) small-scale reverse faults Metschel Tillite.
with displacements of centimeters to a few meters, (3) overturned
folds, and (4) recumbent folds (Fig. 5D). The folds and small- Interpretation
scale thrust faults indicate compression toward the WNW. The Deformational structures in the Metschel Tillite at Mount
thrust sheets are overlain by weakly stratified and stratified dia- Ritchie and Kennar Valley indicate formation caused by horizon-
mictite (see Diamictite Facies Association, below) that lap onto tal compression. These structures include listric-shaped thrust
and drape over the thrust complex. faults, sigmoidal-shaped thrust sheets, fold and thrust nappes,
and recumbent folds. Shear zones at the base of individual thrust
Description of Strata at Kennar Valley sheets are marked by boudin-bearing, foliated mudstones and
At Kennar Valley (Fig. 2), rocks of the Metschel Tillite are decimeter- to meter-thick breccia zones. Substantial evidence
exposed primarily on a ridge that extends northeastward into the indicates that deformation occurred primarily during deposition
center of the valley. Here, highly deformed rocks of the Metschel of the Metschel Tillite. This evidence includes (1) occurrence of
Tillite occur between undeformed strata of the Devonian Aztec the thrust sheets between undeformed strata of the Aztec Silt-
Siltstone below, and undeformed strata of the Weller Coal Mea- stone and glaciogenic deposits of the Metschel Tillite below, and
sures above (Figs. 6 and 7). Within the Metschel Tillite the style undeformed glaciogenic and fluvial strata of the Metschel Tillite
of deformation changes progressively along the ridge toward the and Weller Coal Measures above; (2) truncation and overriding of
southwest from ductile, to brittle, to undeformed (Fig. 6). glaciogenic deposits of the Metschel Tillite by the thrust sheets;
A chaotic zone of ductile deformation is exposed on the east- (3) composition of the thrust sheets, which consist primarily of
ern valley wall and on the northeastern end of the central ridge. glaciogenic deposits; (4) occurrence of both soft sediment defor-
The folding consists of large-scale recumbent and isoclinal folds mational (i.e., massive sandstones, dewatering structures, and
and/or nappe-like structures, with structural displacement toward chaotically folded sandstones and diamictites) and brittle (normal
the west-southwest. The strata consist of interstratified, medium- and reverse faults) structures within the thrust sheets, indicating
to coarse-grained, massive to internally deformed sandstones and that both ductile and brittle deformation and water expulsion from
stratified diamictites (Figs. 6A–6D). Some of the sandstone units the sediments occurred, which is characteristic of deformation of
contain dewatering features (pipes, sheets, and dish structures). unconsolidated and weakly consolidated deposits; and (5) onlap-
Low-angle thrust-fault–bounded packages up to 10 m thick ping and overlapping of the thrust complexes by diamictites,
occur throughout the middle and southern parts of the central which indicate that continued glaciogenic deposition occurred
ridge (Figs. 6 and 7A–7C). These thrust sheets occur in front, and across relief generated by the compressional structures.
on top of, the chaotic ductile zone described above. The faults, Thrust sheets truncating (Kennar Valley), and composed of
marked by 0.01–1-m-thick, boudin-bearing, foliated and homog- strata (Mount Ritchie) derived from the underlying Aztec Silt-
enized mudstones and siltstones (Figs. 7D and 7E), dip at up to stone, indicate that some excavation of bedrock also occurred.
51° toward the east-northeast (toward 72°), indicating structural Because the zone of deformation appears to have occurred in
transport of the sheets toward the west-southwest (252°). These unconsolidated glaciogenic sediment, and because it is confined
faults have high-angle dips along the northeast part of the ridge between undeformed strata of the Aztec Siltstone below and the
but become subhorizontal toward the southwest. Internally within Permian Weller Coal Measures above, the compressional struc-
the thrust-fault–bounded packages, beds are slightly folded and tures are interpreted to be intraformational features associated
contain abundant high-angle normal and listric-shaped reverse with glacial activity rather than being due to tectonic stresses.
faults (Figs. 7C, 7F, 7G). The normal faults are common at In glaciogenic settings, compressional features result from
the bases of the thrust sheets and occur within the concave-up either sliding-slumping or glaciotectonic deformation, both
troughs of small, open synclines (Fig. 7C). On the west-northwest of which produce similar structures. Slides occur on slopes
side of the central ridge (Fig. 7A and 7B), stacked thrust sheets where gravitational failure of the substrate results in rotational
A

B Weller Coal Measures

Brittle zone

Unconformities
Surfaces (bedding Metschel Chaotic zone (Fig. 6C)
and faults) Tillite
C D

Brittle Ductile
Glacier

Figure 6. Folded and thrust-faulted strata of the Metschel Tillite at Kennar Valley that are interpreted to have formed as a subaquatic thrust mo-
raine. (A, B) Highly folded strata (right side of photo), giving way to strata deformed by thrust faults (left side of photo). The strata are exposed
on the east-northeast end of the central ridge in Kennar Valley. The photo shows a slope and cliff face ~50 m high. (C) Closeup of the highly
folded portion of the cliff face shown in Figures 6A and 6B. (D) Highly deformed strata contained within the chaotic zone exposed on the west-
northwest side of the central ridge in Kennar Valley. Jacob’s staff (1.6 m) for scale. (E) Schematic diagram showing formation and orientation of
deformation in a modern thrust moraine (diagram modified from Croot, 1988).
A

C D

Thrust
sheet

Major
thrust fault

Normal faults Thrust fault

E F G Normal faults

Major Major Internal


thrust fault thrust fault thrust fault
Figure 7. Strata exposed on the west-northwest–facing slope of the central ridge in Kennar Valley, interpreted to be part of a thrust moraine
complex. (A, B) Thrust faults and thrust sheets exposed along the northern end of the central ridge. The Weller Coal Measures are ~28 m thick,
for scale. Stereonet shows the orientation of thrust faults, and the rose diagram shows the ice-flow directions interpreted from fault orientations.
(C) Thrust sheet showing basal thrust fault and internal accommodation faults. Jacob’s staff (1.6 m) for scale (white arrow). (D, E) Foliated and
homogenized mudstones mark the position of major thrust faults. (F) Small-scale thrust fault at the base of a thrust sheet. Jacob’s staff (1.6 m)
for scale (white arrow). (G) Normal faults near the base of a thrust sheet; 15-cm-long ruler for scale. VM—vector mean, MVL—mean vector
length, N—number of measurements.
Permian deposits, southern Victoria Land, Antarctica 91

extension of a sediment mass along listric glide planes followed thrust sheets rest on top of younger sheets or where early formed
by downslope sliding of a coherent block away from a headwall. sheets display near-vertical orientations owing to rotation and
Deposition occurs when the block comes to rest at the toe of the piggyback transport on younger sheets that formed along frontal
slide. Slumps form in a similar fashion. However, slumps display thrust faults during continued compression; (5) a decrease in the
internal folding and faulting. The resulting slide-slump structures degree of deformation in the direction of transport owing to the
consist of either a single block or stacked blocks or sheets (mul- formation of frontal thrust faults; and (6) thrust complexes that
tiple slide-slump events), each bounded below by a sheared glide typically contain excavated blocks of bedrock (Croot, 1988; Aber
plane (Allen, 1985; Collinson and Thompson, 1989; Ricci Luc- et al., 1989).
chi, 1995; Benn and Evans, 1998). Slump and/or slide blocks are Small thrust sheets composed of periglacial material also
identified by the following criteria: (1) truncation of underlying occur along both terrestrial and subaqueous ice margins. These
strata from rotational extension in areas adjacent to the slump- thrust blocks or push moraines, which are typically no more than
slide scarp; (2) deposits, which at the head of the slump-slide 5 m high, form during small-scale seasonal advance and retreat
dip away from the headwall scarp; (3) deposits, which at the toe cycles of the ice front. During a seasonal advance, periglacial
of the slump-slide typically rest on slopes that dip in the direc- material is pushed or “bulldozed” into a morainal ridge. Although
tion of sliding; (4) deposits typically composed of only one to these features are commonly composed of supraglacial debris
several sheets with younger sheets resting on the backs of older dumped at the ice margin, some also include proglacial sediment
sheets; (5) the occurrence of slump fold noses; (6) an increase in (Bennett and Glasser, 1996).
the degree of deformation in the direction of transport owing to Compressional deformation in the Metschel Tillite is consis-
compression at the toe of the slump-slide; and (7) deposits that tent with formation by glaciotectonic deformation of periglacial
typically do not contain excavated bedrock blocks (cf. Tucker, deposits. Supraglacial material does not occur within the thrust
1996; Collinson and Thompson, 1989; Ricci Lucchi, 1995; Stow, sheets. At Mount Ritchie, shear zones truncate glaciogenic and/or
2005). Slides and slumps are commonly associated with debris Devonian strata throughout the thrust complex. Near the front of
flows, which are often triggered during formation and movement the complex the basal thrust sheet ramps up and over preexisting
of the slump and slide blocks. deposits, with the orientation of the truncation surface dipping
Structures formed by glaciotectonic compression develop in a direction opposite to that of the direction of transport for the
at ice margins during glacial advance and involve displacement sheets. Because thrust sheets of the Aztec Siltstone rest on these
of subglacial and proglacial sediment, and weak rock by both basal imbricated sheets, the “Aztec” sheets likely formed early,
ductile and brittle deformation (Aber et al., 1989; Hart and Boul- only to be later transported piggyback on younger structures that
ton, 1991). Deformation is facilitated by (1) horizontal stresses developed along frontal thrust faults. Dip directions of the thrust
resulting from high ice overburden pressures up-glacier, (2) weak sheets at Mount Ritchie suggest glaciotectonic transport, and
subglacial and periglacial substrates that produce décollements therefore glacial advance toward 307°.
during failure, (3) high pore-water pressures that reduce the cohe- At Kennar Valley the lateral change from highly contorted
sive and yield strength of ice marginal and subglacial sediments beds, recumbent folds, and thrust nappes to thrust sheets indi-
and coherent substratum, and (4) by the presence of reverse slopes cates a decrease in the intensity of deformation from ductile to
along the glacial margin that increase horizontal stress (Bluemle brittle in the direction of structural transport, which was toward
and Clayton, 1984; Aber et al., 1989; Boulton and Caban, 1995). 252°. The high-angle orientation of the proximal deforma-
Ductile deformation includes formation of open to overturned tion zone suggests that these may have been the first sheets to
folds, whereas brittle failure results in the development of low- have formed and that they were later transported piggyback on
angle overthrusts, listric thrust faults, imbricately stacked thrust younger thrust sheets and rotated into high angles during contin-
sheets, and nappes (Hart, 1990; Van der Wateren, 2002). “Piggy- ued propagation of the thrust front. This progression of structures
back” thrusting is common where early formed proximal sheets is similar to proximal to distal changes in deformation style seen
are carried on the backs of later formed distal blocks. In this sce- within modern ice-marginal thrust moraine complexes (Fig. 6E;
nario, intense folding and high-angle reverse faulting occur along cf. Croot, 1988; Van der Wateren, 2002). Therefore, structures
the ice margin, decreasing to lower angle emplacement of thrust exposed at Kennar Valley are interpreted to have formed as gla-
sheets outward (Eybergen, 1987; Croot, 1988; Van der Wateren, ciotectonic features along an ice margin. During emplacement,
2002). Thrust moraines are the surface expression of this defor- flexure of the deforming mass would have produced both exten-
mation (Aber et al., 1989). sional and compressional stresses within individual thrust sheets
Several criteria can be used to identify thrust sheet complexes where the thrust sheets were flexed during transport into broad
produced by glaciotectonic deformation. These include (1) trun- synclines and anticlines. Such stresses would have resulted in the
cation of underlying strata throughout the zone of deformation, formation of internal accommodation features (i.e., normal and
especially along frontal thrusts; (2) occurrence of compressional listric-shaped reverse faults), which would have facilitated flex-
features throughout; (3) décollement–thrust fault surfaces that ing of the thrust sheets.
ramp up and over truncated proglacial sediment near the lead- In modern glacial settings, glaciotectonic deformation occurs
ing edge of the frontal thrust; (4) thrust complexes where older along both terrestrial and glaciomarine ice margins (Bennett
92 Isbell

et al., 1999; Van der Wateren, 2002). Massive, weakly stratified, in diameter occur (Fig. 9A). Many of these clasts cut stratifi-
and stratified diamictite facies associated with the thrust sheets cation and have their long axes oriented at high angles to bed-
at Mount Ritchie suggest that deposition occurred in a basinal ding. Faceted and striated clasts also occur. At the base of the
setting (see next section). The lithologies and sedimentary struc- overlying Weller Coal Measures, gravel clasts (pebbles, cobbles,
tures of strata contained within the thrust sheets at Kennar Valley, and boulders) protrude downward into the underlying diamic-
which include stratified to massive diamictite (see next section), tites, as do load structures at the base of coarse-grained sand-
coarsening upward siltstone to sandstone successions containing stones and pebble to cobble conglomerate units. Intrusion of the
wave ripple laminae, and swaley cross-stratification, also suggest Weller by diamictite diapirs also occurs, as does interfingering of
that ice advanced into a glaciomarine environment. Metschel diamictites and Weller sandstones and conglomerates
(Fig. 9B). These coarse-grained Weller sediments are part of a
Diamictite Facies Association 6- to 10-m-thick wedge-shaped coarse- to very coarse grained
trough-cross-bedded sandstone body (Figs. 10A–10D). When
Description traced over several hundred meters, this body displays an inter-
Massive to weakly stratified diamictite occurs at Kennar Val- nal geometry characterized by low-angle downlapping beds and
ley and at Mount Ritchie (Figs. 2 and 8). At Kennar Valley several surfaces. When viewed on exposures perpendicular to paleoflow
of the thrust sheets are composed entirely of this type of diamic- orientations, the downlapping units are cut by numerous small-
tite (see large-scale thrust-sheet-facies association, Figs. 8A and scale (meters to tens of meters wide and up to a few meters deep)
8B). At Mount Ritchie, massive and weakly stratified diamictites cross-bedded, sand-filled channels. Paleocurrent directions are
typically have gradational upper and lower contacts with strati- oriented toward 234° (Figs. 10C and 10D).
fied diamictite. However, weakly stratified diamictite is also trun-
cated by conglomerates and sandstones in the overlying Weller Interpretation
Coal Measures. The diamictite consists of a clay to fine-grained Massive and stratified diamictites are commonly interpreted
sand matrix containing scattered to abundant clasts. Laminae to have formed in both polar and temperate glaciomarine sys-
and thin wisps of silt and sand, which are better sorted than the tems by a number of different processes. In polar settings the
surrounding matrix, define faint stratification. Clasts of granite, temperature and strength of the ice allow for the development of
quartzite, and gneiss up to 1 m in diameter, pierce stratification, floating glacial tongues and ice shelves. Under these conditions,
and the long axes of some clasts are oriented at high angles to ice decouples from the bed and begins to float just seaward of a
bedding. Within massive to weakly stratified diamictite units, grounding line. Here, in this proximal setting, melt-out of basal
isolated centimeter- to meter-thick and decimeter- to meter-wide debris from the underside of the glacier allows both fine- and
pods of contorted and massive sandstone and conglomerate with coarse-grained particles to settle through the water column and to
dewatering structures occur (Fig. 8C and 8D). Injection struc- produce massive diamictites. Sedimentation rates are relatively
tures in the form of diamictite diapirs intrude these pods. high near the grounding line, but they decrease distally owing to
Stratified diamictites are common at Mount Metschel, the loss of debris-rich basal ice (Evans and Pudsey, 2002). Strati-
Mount Ritchie, and within the thrust sheets at Kennar Valley fied diamictites occur where sedimentation rates are lower, and
(Figs. 8E–8G). These diamictites have a similar matrix and clast where bottom currents (e.g., tidal pumping) winnow out finer
(lithology and size) composition as those of the massive and grained particles (Domack et al., 1999). Dropstones are produced
weakly stratified diamictites. Stratification is distinct, however, by either iceberg rafting in open-marine settings or as rain-out
and consists of centimeter- to decimeter-thick semi-continuous from the debris-poor distal portions of the ice shelf or glacial
layers of silty sand–rich layers alternating with mud-rich lay- tongue (Evans and Pudsey, 2002).
ers. In a few places, intercalations of centimeter-scale units of Under temperate conditions, warmer, weaker ice typically
normally graded to cross-laminated, medium-grained sandstones results in tidewater glaciers entering the sea as an actively calv-
delineate stratification (Fig. 8G). The bases of the sandstones ing ice cliff. Under these conditions the glacial terminus is
commonly display load structures. Within the stratified diamic- grounded or partially floating. Here, glacial deposition is domi-
tite, clasts up to 30 cm in diameter are common and cut strati- nated by meltwater and, to an equal or lesser extent, by rafting of
fication. Within the stratified diamictites, decimeter-thick and debris by icebergs. Grounding-line fans form where subglacial
tens-of-meters-long lenses, or lobe-shaped bodies of massive meltwaters, emerging from conduits along the ice front, deposit
sandstone and conglomerate, occur (Fig. 4E). These bodies char- sand and gravel. As velocity in the effluent flow drops, buoyant
acteristically contain sharp bases with abundant centimeter- to forces become dominant, and the inflowing fresh water rises to
decimeter-scale load structures. The bodies are also intruded by the surface to form an overflow plume that transports fine sand,
diapirs of stratified diamictite. Boulders up to 0.7 m in diameter silt, and clay seaward. Sedimentation occurs as particles released
protrude from the top of the lobe-shaped bodies. from the plume settle through the water column (Cowan and
On the north-northwest and west-southwest sides of Mount Powell, 1990). Plume sedimentation, in conjunction with dump-
Ritchie, just below the contact with the overlying Weller Coal ing of coarse debris during calving at the ice front, and by the
Measures, locally abundant accumulations of clasts up to 1 m release of clasts from the melting of icebergs, produces stratified
A B
Weakly stratified diamictite

Massive diamictite
Thrust faults

Massive diamictite

C Sandstone D
lens

Sandstone
lens

E F G

Figure 8. (A) Massive diamictite contained within two different thrust sheets exposed on the west-northwest–facing slope of the central ridge
in Kennar Valley. Person for scale. (B) Massive diamictite grading into weakly stratified diamictite on the west-northwest–facing slope of the
central ridge in Kennar Valley. The diamictite unit is ~4 m thick at its thickest point in the photo. (C) Deformed sandstone lens contained within
weakly stratified diamictite exposed on the eastern side of Mount Ritchie. Person for scale. (D) Deformed pebbly sandstone lens containing
diamictite diapirs near the summit of Mount Ritchie. Trekking pole for scale. (E) Stratified diamictite containing a lens of massive diamictite at
Mount Metschel. Ice axe for scale. (F) Stratified diamictite, Mount Ritchie; 15-cm-long ruler for scale. (G) Interstratified diamictite and beds of
medium- to coarse-grained sandstone, Mount Ritchie. Persons for scale.
A A

B
C

Figure 9. (A)Weakly stratified diamictite at the top of the Metschel


D
Tillite at Mount Ritchie. Note the abundance of clasts and the high dip
angle of the long axes of many of the clasts. Clasts in the overlying
Weller Coal Measures also project down into the underlying diamic-
tite. Hammer for scale. (B) Interfingering of weakly stratified dia-
mictite of the Metschel Tillite, and sandstone and conglomerate of the
Weller Coal Measures. Trekking pole for scale.

Figure 10. (A, B) Sandstone sheet containing diamictite lenses and


a channel body in the Metschel Tillite. A wedge-shaped sandstone
at the base of the Weller Coal Measures containing downlapping
surfaces is also shown. For scale, the distance from the base of the
Weller Coal Measures to the base of the dolerite sill (Jurassic) is 25 m.
(C, D) Wedge-shaped sandstone body at the base of the Weller Coal
Measures, containing broad, channel-like scours. For scale, the dis-
tance from the base of the Weller Coal Measures to the base of the
dolerite sill is 25 m. (E) Dewatering pipes within massive sandstone of
the sandstone sheet facies association. Mechanical pencil for scale.
Permian deposits, southern Victoria Land, Antarctica 95

diamictites in proximal glaciomarine positions (Cowan and and (3) Metschel diamictites intruding Weller strata as diapirs.
Powell, 1991; Smith and Andrews, 2000). Massive diamictites The occurrence of load structures and diapirs suggests failure of
occur in ice distal settings owing to iceberg rafting and scour- a water-saturated Metschel substrate owing to deposition of the
ing (Dowdeswell et al., 1994). However, massive and stratified overlying Weller sandstones and conglomerates. Downlapping
deposits and/or mudstones lacking clasts can occur in either surfaces within the basal Weller strata indicate progradation of
proximal or distal glaciomarine positions because of changes in a wedge-shaped body across proximal Metschel glaciomarine
sea ice and/or oceanographic conditions (Smith and Andrews, deposits (massive diamictite and abundant dropstones). The basal
2000; Dowdeswell et al., 2000). In both polar and temperate set- sandstone and conglomerate body has a geometry and internal
tings, stratification is also produced by periodic remobilization of features that are similar to grounding-line fans described by Pow-
the deposits by sediment gravity flows (Evans and Pudsey, 2002; ell (1990) and Powell and Alley (1997).
Powell and Domack, 2002).
In the Metschel Tillite, the occurrence of massive, weakly Sandstone Sheet Facies Association
stratified, and stratified diamictites with gradational upper and
lower contacts is suggestive of deposition in proximal to distal Description
glacial basinal settings during fluctuations in the location of the On the northwest side of Mount Ritchie, strata near the
ice front. Although it is unknown whether strata in southern Vic- middle of the Metschel Tillite consist of a 16.5-m-thick succes-
toria Land were deposited under glaciolacustrine or glaciomarine sion of fine- to medium-grained cross-bedded sandstone, fine- to
conditions, a glaciomarine setting seems likely owing to an abun- medium-grained massive sandstone containing dewatering pipes
dance of stratified diamictites, which suggests that sedimentation (Fig. 10E), shale, and massive to weakly stratified diamictite.
occurred from settling of particles from buoyant low-density These units are contained within a sheetlike sediment body that
meltwater plumes. Clasts in these deposits were likely introduced is in erosional contact with underlying shales and diamictites
from the melting out of debris from ice fronts or from the release (Figs. 10A–10D). Sandstone within the sheet is laterally con-
of debris rafted by icebergs. The abundance of massive, normally tinuous. However, diamictite and shale are discontinuous within
graded, and cross-bedded sandstone layers and pods within the the sheet and either drape underlying beds or form lens- to pod-
strata is indicative of deposition from subaqueous meltwater as shaped bodies. On the eastern end of the west-northwest face of
tractive flows and as sediment gravity flows, suggesting temper- Mount Ritchie a large channel is incised to a depth of 10 m into
ate glacial thermal conditions (Mackiewicz et al., 1984; Powell the sheet (Figs. 10A and 10B). Laterally, channel margins are
and Domack, 2002). Sandstones were deposited subaqueously concordant with beds in the underlying sandstone. The channel
on water-saturated substrates, which, when loaded, failed, pro- is filled with clast-supported conglomerate, cross-bedded sand-
ducing load structures, dewatering structures, and disruption of stone, massive sandstone containing dewatering pipes, and by
the sandstone bodies by intrusion of diamictite diapirs. Such massive to weakly stratified diamictite. The sheet and channel
unstable substrates are the result of high sedimentation rates in are overlain by, and interfinger with, massive, weakly stratified,
ice proximal zones (Boulton, 1990). The occurrence of lens- and and stratified diamictite (see section on Diamictite Facies Asso-
lobe-shaped sandy diamictite beds with protruding boulders is ciation, above). On the west-southwest side of Mount Ritchie
suggestive of deposition from debris flows with the boulders the sandstone sheet is overridden by thrust sheets (see section on
introduced as ice-rafted debris. However, small lenses of sand Large-Scale Thrust Sheet Facies Association, above; Figs. 10C
and conglomerate could have formed as iceberg dump structures and 10D).
(Thomas and Connell, 1985). Ice rafting of debris was likely an
important component of sedimentation in both proximal and dis- Interpretation
tal locations as indicated by the occurrence of clasts penetrat- The interfingering of sandstone and massive to stratified
ing stratification (Thomas and Connell, 1985). Locally abundant diamictite (see section on Diamictite Facies Association, above)
boulders at the top of the Metschel Tillite at Mount Ritchie may suggest that the sandstone sheet was deposited in a glacioma-
indicate either iceberg dump structures or dumping of clasts at rine setting. Features within the sandstone sheet, which include
the ice front during calving events (Thomas and Connell, 1985; channeling, conglomerates, and cross-bedded sandstone, indicate
Powell and Domack, 2002). Further evidence of a glacial origin deposition from high-energy tractive currents. Massive sandstone
for these strata is provided by the occurrence of striated and fac- with dewatering pipes suggests rapid sedimentation rates possibly
eted dropstones. from either suspension or sediment gravity flows. These condi-
Strata at the base of the overlying Weller Coal Measures tions are common near the grounding lines of temperate tidewa-
show evidence of deposition contemporaneous with or shortly ter glaciers (Powell, 1990). In this setting, fresh-water effluent
following deposition of the Metschel Tillite. Evidence includes flow, emanating from the base of the glacier, deposits wedge-
(1) interfingering of Metschel diamictites with Weller sandstones shaped bodies of sand and gravel known as grounding-line fans.
and conglomerates, (2) boulders and cobbles in the Weller con- These bodies are laterally continuous for hundreds of meters and
glomerates protruding downward into the underlying diamictite, are commonly cut by channels as the effluent flow cuts into early
96 Isbell

formed deposits (Powell, 1990). Away from the glacial front, cur- of thrust sheets, orientation of foliation, and analyses of folds.
rent velocity drops, allowing the low-density fresh-water flow to Interpretation of these data suggests that glacial ice converged on
detach from the bottom and rise to the surface to form a buoyant southern Victoria Land off the East Antarctic craton to the west
overflow plume. Rapid sedimentation from the rain-out of sand, (e.g., glaciotectonite and deformation tills at Mount Metschel
silt, and clay from the plume results in deposits with high initial and Mount Ritchie) and off an area in the direction of the pres-
water contents (Cowan and Powell, 1990). These ice proximal ent Ross Sea to the east (Fig. 11; e.g., thrust sheets at Kennar
deposits are unstable and are highly susceptible to dewatering Valley and Mount Ritchie). If these directions are correct, then
and/or to remobilization as sediment gravity flows. Addition of ice advanced from glacial centers on opposite sides of the depo-
coarse debris from ice rafting results in a complex interfingering sitional basin. Expansion of ice from these centers then allowed
of sands, conglomerates, muds, and diamictites (Powell, 1990; advance of ice margins into a glaciomarine setting in southern
Powell and Domack, 2002). The sandstone sheet facies associa- Victoria Land.
tion is here interpreted as a grounding-line fan. This interpreta-
tion is consistent with the observed sedimentary features and DISCUSSION
with the overriding of the sandstone sheet on the west-southwest
side of Mount Ritchie by ice proximal thrust sheets, which also Models for the Late Paleozoic Ice Age place Victoria Land
are common near the grounding line of some tidewater glaciers beneath the center of an immense Gondwanan Ice Sheet that
(Bennett et al., 1999). waxed and waned throughout the Mississippian, Pennsylvanian,
and Permian (e.g., Scotese et al., 1999). These models predict
OVERALL DEPOSITIONAL SETTING that terrestrial ice flowed southward across southern Victoria
Land out of a major glacial spreading center (Lindsay, 1970; Bar-
Deformational and depositional lithofacies associations in rett, 1991; Veevers, 2001). However, lithofacies and paleocurrent
the Metschel Tillite suggest that sedimentation occurred in ice data presented here do not support such a conclusion. Instead, the
marginal and glaciomarine settings. Deformational features are results of this study suggest that expansion of temperate glaciers,
interpreted as glaciotectonites, deformation tills, and thrust flowing out of smaller glacial centers, converged on southern Vic-
duplexes. Glaciotectonites and deformation tills resulted from toria Land and extended into and retreated out of a glaciomarine
subglacial deforming beds, which, in modern settings, typically setting during the Early Permian. No evidence for Carboniferous
form beneath ice streams, outlet glaciers, tidewater glaciers, and glaciation occurs. Therefore, glaciogenic strata of the Metschel
surging glacial lobes where high pore-water pressures facilitate Tillite are inconsistent with deposition in this area from a single,
deformation of unconsolidated or weakly consolidated substrates. massive, long-duration Gondwanan Ice Sheet.
Thrust duplexes formed as periglacial thrust moraines along ice Recent work in other areas of Antarctica and Gondwana
margins. The occurrence of these strata and their glaciotectonic are also challenging the prevailing view of the extreme size and
structures suggests that deposition in southern Victoria Land duration of late Paleozoic glaciation. In the central Transantarc-
occurred at or near glacial termini. The occurrence of massive tic Mountains, ongoing facies and paleocurrent analyses do not
and stratified diamictites, lonestone-bearing deposits, sheet sand- support the traditional view of terrestrial ice flowing radially out
stones, and sandstones with swaley cross-stratification and wave of Victoria Land (Isbell et al., 1997, 2005, 2008; Isbell, 1999).
ripple laminations suggests that Permian glaciers in southern Instead, results show that ice converged on an elongate basin
Victoria Land advanced into, and retreated from glaciomarine whose long axis was oriented parallel to the present trend of the
settings. Owing to the occurrence of meltwater plume deposits mountain range (Figs. 1 and 11). In the central Transantarctic
(stratified diamictites), the depositional setting was most likely Mountains, glaciers, grounded along basin margins in the direc-
glaciomarine. Evidence in the form of glaciotectonite, defor- tion of the present polar plateau and in the direction of the Ross
mation till, and cross-bedded sandstone indicates that abundant Sea–Marie Byrd Land, flowed transversely off the margins into
meltwater was present at the time of deposition, and therefore a glaciomarine setting. This scenario is similar to that proposed
that temperate thermal conditions characterized the depositing for southern Victoria Land, thus suggesting the occurrence of
ice. Such thermal conditions are characteristic of tidewater gla- multiple Permian glacial centers in Antarctica (Isbell et al.,
ciers where deposition is dominated by subglacial deforming bed 1997, 2005, 2008; Isbell, 1999). Elsewhere in Antarctica, inter-
conditions, grounding line processes, meltwater outflow, meltwa- pretations of earlier work suggest that a glacial center on the
ter plumes, and iceberg rafting of debris. Ellsworth Mountains crustal block supplied ice to glaciomarine
Only a few ice-flow directions have been reported from settings in the Ellsworth and Pensacola Mountains (Fig. 11C;
upper Paleozoic glaciogenic rocks in southern Victoria Land Frakes et al., 1971; Matsch and Ojakangas, 1991; Collinson
(Barrett and Kohn, 1975; McKelvey et al., 1977). Flow directions et al., 1994).
from these data are ambiguous. Although deformation structures In Queensland and New South Wales, Australia, Jones and
reported in this paper represent local glaciotectonic displacement, Fielding (2004), Birgenheier et al. (2005, 2009), and Fielding
these features also provide a record of paleo-ice-flow directions. et al. (2005, 2008) reported the occurrence of short, discrete
Transport directions were derived from displacement directions intervals of mountain-valley glaciations, ice caps, and/or small
Permian deposits, southern Victoria Land, Antarctica 97

A C B
°
b

Figure 11. Tectonic transport direction for rocks of the Metschel Tillite as indicated by the orientation of (A) thrust faults and foliation in glacio-
tectonite and deformation till (Mount Metschel and Mount Ritchie) and the orientation of (B) thrust faults associated with thrust sheets (Kennar
Valley, Mount Ritchie). (C) Map showing interpreted flow directions in southern Victoria Land, plotted with data from glaciogenic strata else-
where in the Transantarctic and Ellsworth Mountains (EM), and in South Africa. Data from Frakes et al. (1971), Barrett (1981), Collinson et al.
(1994), Visser (1997), Isbell (1999), Lenaker (2002). v.m.—vector mean.

ice sheets. Their findings are in marked contrast with earlier Earth systems during the late Paleozoic. Data obtained from the
reports that suggested that much of Australia was covered by a Metschel Tillite clearly show that multiple glaciers were active in
continental-scale polar ice sheet. southern Victoria Land during the Permian. These results, cou-
The hypothesis that numerous ice centers occurred in Gond- pled with recently reported data in Australia and South America,
wana during the late Paleozoic is not a new concept. Work by strongly suggest that Gondwana glaciation was characterized by
Crowell and Frakes (1970), Caputo and Crowell (1985), Eyles numerous glacial centers and alpine glaciers rather than by a sin-
(1993), López Gamundí (1997), Limarino et al. (2002), Isbell gle massive ice sheet. Because the geographic area of ice cover,
et al. (2003), Henry et al. (2008), and Isbell et al. (2008) showed ice volume, and the number of ice sheets are directly related,
that multiple ice sheets, ice caps, and alpine glaciers diachro- multiple glaciers, for a given land area, contain considerably less
nously waxed and waned as Gondwana drifted across the late ice volume than a single massive ice sheet (Crowley and Baum,
Paleozoic South Pole. The glacial record in southern Victoria 1991; Isbell et al., 2003). Therefore, multiple Gondwana glaciers
Land is consistent with the concept of multiple ice centers within would have had a completely different impact on Earth’s natural
Gondwana, and interpretation of the record disproves that Ant- systems than that of a massive ice sheet. For example, the waxing
arctica was covered by a single, massive ice sheet during the late and waning of multiple ice sheets would have resulted in consid-
Paleozoic. erably smaller changes in eustatic sea level than those produced
Identification of the size and duration of Gondwana glacia- by growth and decay of a single glacier covering the same geo-
tion is of great importance in developing an understanding of graphical area.
98 Isbell

CONCLUSIONS Tingey, R.J., ed., The Geology of Antarctica: Oxford, UK, Oxford Uni-
versity Press, p. 120–152.
Barrett, P.J., and Kohn, B.P., 1975, Changing sediment transport directions
Glaciotectonic deformation and glaciomarine deposits with- from Devonian to Triassic in the Beacon Supergroup of South Victoria
in the Metschel Tillite indicate that late Paleozoic glaciation was Land, Antarctica, in Campbell, K.S.W., ed., Gondwana Geology: Can-
characterized by temperate glacial conditions along Permian ice berra, Australian National University Press, p. 15–35.
Barrett, P.J., and Kyle, R.A., 1975, The early Permian glacial beds of south-
margins in southern Victoria Land rather than by polar glacial ern Victoria Land and the Darwin Mountains, Antarctica, in Campbell,
conditions associated with a glacial spreading center as previ- K.S.W., ed., Gondwana Geology: Canberra, Australian National Univer-
ously hypothesized. Deposition occurred beneath glaciers under- sity Press, p. 333–346.
Barrett, P.J., and McKelvey, B.C., 1981, Permian tillites of southern Victoria
going deforming bed conditions and in a glaciomarine setting Land, Antarctica, in Hambrey, M.J., and Harland, W.B., eds., Earth’s
in front of either alpine glaciers, outlet glaciers, ice streams, or Pre-Pleistocene Glacial Record: Cambridge, UK, Cambridge University
Press, p. 233–236.
surging glacial lobes. The direction of glaciotectonic transport Barrett, P.J., Elliot, D.H., and Lindsay, J.F., 1986, The Beacon Supergroup
indicates that ice converged on southern Victoria Land from two (Devonian-Triassic) and Ferrar Group (Jurassic) in the Beardmore Glacier
directions. One source of ice was from the direction of the pres- area, Antarctica, in Turner, M.D., and Splettstoesser, J.F., eds., Geology of
the Central Transantarctic Mountains: Washington, D.C., American Geo-
ent East Antarctic craton. The other source of ice was from the physical Union, Antarctic Research Series, p. 339–428.
opposite direction, that of the present Ross Sea. These findings Benn, D.I., 1995, Fabric signature of subglacial till deformation, Brei-
imply that multiple glaciers were active in this region. Because damerkurjökull, Iceland: Sedimentology, v. 42, p. 735–747, doi: 10.1111/
j.1365-3091.1995.tb00406.x.
multiple ice sheets covering a given area contain less ice than a Benn, D.I., and Evans, D.J.A., 1996, The interpretation and classification of
single ice sheet, multiple glaciers would have influenced Earth subglacially deformed materials: Quaternary Science Reviews, v. 15,
systems (e.g., eustatic sea level, climate) differently than the mas- p. 23–52, doi: 10.1016/0277-3791(95)00082-8.
Benn, D.I., and Evans, D.J.A., 1998, Glaciers and Glaciation: London, Edward
sive-ice-sheet models predict. Arnold, 734 p.
Bennett, M.R., and Glasser, N.F., 1996, Glacial Geology: Ice Sheets and Land-
ACKNOWLEDGMENTS forms: New York, Wiley & Sons, 364 p.
Bennett, M.R., Glasser, N.F., Crawford, K., Hambrey, M.J., and Huddart, D.,
1999, The landform and sediment assemblage produced by a tidewater
Discussions with Rosemary Askin, Jim Collinson, Ellen glacier surge in Kongsfjorden, Svalbard: Quaternary Science Reviews,
Cowan, Dyana Czeck, Pete Flaig, Tom Hooyer, Mark Johnson, v. 18, p. 1213–1246.
Birgenheier, L.P., Fielding, C.R., Frank, T.D., and Roberts, J., 2005, Strati-
Carlos Oscar Limarino, and Molly Miller are greatly appreciated. graphic record of late Paleozoic Gondwanan ice age in New South Wales,
I also thank Paul Lenaker, Rosemary Askin, Tim Cully, Molly Australia: A review and revision of the Carboniferous System: Geological
Miller, and Keri Wolfe for their help in the field. Peter Barrett, Society of America Abstracts with Programs, v. 37, no. 7, p. 256.
Birgenheier, L.P., Fielding, C.R., Rygel, M.C., Frank, T.J., and Roberts, J.,
Luis Buatois, Jim Collinson, Chris Fielding, and Antonio Carlos 2009, Evidence for dynamic climate change on sub-106-year scales from
Rocha-Campos provided valuable comments on earlier drafts of the late Paleozoic Glacial record, Tamworth Belt, New South Wales, Aus-
this paper. The U.S. National Science Foundation, Raytheon Polar tralia: Journal of Sedimentary Research, v. 79, p. 56–82.
Bluemle, J.P., and Clayton, L., 1984, Large-scale glacial thrusting and related
Services, the New York Air National Guard, Kenn Borek Air LTD, processes in North Dakota: Boreas, v. 13, p. 279–299.
Trans World Logistics, and Petroleum Helicopters Incorporated Boulton, G.S., 1979, Processes of glacier erosion on different substrata: Journal
provided logistic support for fieldwork in Antarctica. National of Glaciology, v. 23, p. 15–38.
Boulton, G.S., 1987, A theory of drumlin formation by subglacial sediment
Science Foundation grants OPP-9909637, ANT-0440919, ANT- deformation, in Menzies, J., and Rose, J., eds., Drumlin Symposium: Rot-
0635537, and OISE-0825617 supported this work. terdam, A.A. Balkema, p. 25–80.
Boulton, G.S., 1990, Sedimentary and sea level changes during glacial cycles
and their control on glacimarine facies architecture, in Dowdeswell, J.A.,
REFERENCES CITED and Scourse, J.D., eds., Glacimarine Environments: Processes and Sedi-
ments: Geological Society of London Special Publication 53, p. 15–52.
Boulton, G.S., 1996, Theory of glacial erosion, transport and deposition as a
Aber, J.S., Croot, D.G., and Fenton, M.M., 1989, Glaciotectonic Landforms consequence of subglacial sediment deformation: Journal of Glaciology,
and Structures: Dordrecht, Kluwer Academic Publishers, 200 p. v. 42, p. 43–62.
Allen, J.R.L., 1985, Principles of Physical Sedimentology: London, George Boulton, G.S., and Caban, P., 1995, Groundwater flow beneath ice sheets:
Allen & Unwin, 272 p. Part II—Its impact on glacier tectonic structures and moraine forma-
Alley, R.B., 1991, Deforming-bed origin for southern Laurentide till sheets?: tion: Quaternary Science Reviews, v. 14, p. 563–587, doi: 10.1016/0277
Journal of Glaciology, v. 37, p. 67–76. -3791(95)00058-W.
Askin, R.A., 1998, Floral trends in the Gondwana high latitudes: Palynologi- Boulton, G.S., and Hindmarsh, R.C.A., 1987, Sediment deformation beneath
cal evidence from the Transantarctic Mountains: Journal of African Earth glaciers: Rheology and sedimentological consequences: Journal of Geo-
Sciences, v. 27, p. 12–13. physical Research, v. 92, p. 9059–9082, doi: 10.1029/JB092iB09p09059.
Banham, P.H., 1977, Glacitectonics in till stratigraphy: Boreas, v. 6, p. 101– Boulton, G.S., and Jones, A.S., 1979, Stability of temperate ice caps and ice
105. sheets resting on beds of deformable sediment: Journal of Glaciology,
Barrett, P.J., 1972, Late Paleozoic glacial valley at Alligator Peak, southern Vic- v. 24, p. 29–43.
toria Land, Antarctica: New Zealand Journal of Geology and Geophysics, Collinson, J.D., and Thompson, D.B., 1989, Sedimentary Structures: London,
v. 15, p. 262–268. Chapman & Hall, 207 p.
Barrett, P.J., 1981, History of the Ross Sea region during the deposition of the Collinson, J.W., Isbell, J.L., Elliot, D.H., Miller, M.F., and Miller, J.M.G., 1994,
Beacon Supergroup 400–180 million years ago: Journal of the Royal Permian-Triassic Transantarctic basin, in Veevers, J.J., and Powell, C.M.,
Society of New Zealand, v. 11, p. 447–458. eds., Permian-Triassic Pangean Basins and Foldbelts along the Panthalas-
Barrett, P.J., 1991, The Devonian to Jurassic Beacon Supergroup of the Trans- san Margin of Gondwanaland: Geological Society of America Memoir
antarctic Mountains and correlatives in other parts of Antarctica, in 184, p. 173–222.
Permian deposits, southern Victoria Land, Antarctica 99

Cowan, E.A., and Powell, R.D., 1990, Suspended sediment transport and depo- Henry, L.C., Isbell, J.L., and Limarino, C.O., 2008, Carboniferous glacigenic
sition of cyclically interlaminated sediment in a temperate glacial fjord, deposits of the proto-Precordillera of west-central Argentina, in Fielding,
Alaska, U.S.A., in Dowdeswell, J.A., and Scourse, J.D., eds., Glacimarine C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice
Environments: Processes and Sediments: Geological Society of London Age in Time and Space: Geological Society of America Special Paper
Special Publication 53, p. 75–89. 441, p. 131–142.
Cowan, E.A., and Powell, R.D., 1991, Ice-proximal sediment accumulation Hyde, W.T., Crowley, T.J., Tarasov, L., and Paltier, W.R., 1999, The Pangean ice
rates in a temperate glacial fjord, south-eastern Alaska, in Anderson, J.B., age: Studies with a coupled climate–ice sheet model: Climate Dynamics,
and Ashley, G.M., eds., Glacial Marine Sedimentation: Paleoclimatic Sig- v. 15, p. 619–629, doi: 10.1007/s003820050305.
nificance: Geological Society of America Special Paper 261, p. 61–73. Isbell, J.L., 1999, The Kukri Erosion Surface; a reassessment of its relation-
Croot, D.G., 1988, Morphological, structural and mechanical analysis of ship to rocks of the Beacon Supergroup in the central Transantarctic
neoglacial ice-pushed ridges in Iceland, in Croot, D.G., ed., Glaciotecton- Mountains, Antarctica: Antarctic Science, v. 11, p. 228–238, doi: 10.1017/
ics: Forms and Processes: Rotterdam, A.A. Balkema, p. 33–47. S0954102099000292.
Crowell, J.C., and Frakes, L.A., 1970, Ancient Gondwana glaciations, in Isbell, J.L., Gelhar, G.A., and Seegers, G.M., 1997, Reconstruction of pre-
Haughton, S.H., ed., Proceedings and Papers of the Second Gondwana glacial topography using a post-glacial flooding surface: Upper Paleozoic
Symposium, South Africa: Pretoria, CSIR, p. 469–476. deposits, central Transantarctic Mountains, Antarctica: Journal of Sedi-
Crowell, J.C., and Frakes, L.A., 1971, Late Paleozoic glaciation: Part IV, Aus- mentary Research, v. 67, p. 264–272.
tralia: Geological Society of America Bulletin, v. 82, p. 2515–2540, doi: Isbell, J.L., Miller, M.F., Wolfe, K.L., and Lenaker, P.A., 2003, Timing of late
10.1130/0016-7606(1971)82[2515:LPGPIA]2.0.CO;2. Paleozoic glaciation in Gondwana: Was glaciation responsible for the
Crowley, T.J., and Baum, S.K., 1991, Estimating Carboniferous sea-level fluc- development of northern hemisphere cyclothems?, in Chan, M.A., and
tuations from Gondwana ice extent: Geology, v. 19, p. 975–977, doi: 10 Archer, A.W., eds., Extreme Depositional Environments: Mega End
.1130/0091-7613(1991)019<0975:ECSLFF>2.3.CO;2. Members in Geologic Time: Geological Society of America Special Paper
Cúneo, N.R., Isbell, J.L., Taylor, T.N., and Taylor, E.L., 1993, The Glossopteris 370, p. 5–24.
Flora in Antarctica: Taphonomy and paleoecology, C.R.: Buenos Aires, Isbell, J.L., Miller, M.F., Askin, R.A., Lenaker, P.A., and Koch, Z.J., 2005, Late
International Congress of Carboniferous and Permian Stratigraphic Geol- Paleozoic glaciation in Antarctica: Are models depicting an immense ice
ogy, 12th, p. 13–40. sheet correct?: Geological Society of America Abstracts with Programs,
Domack, E.W., Jacobson, E.A., Shipp, S., and Anderson, J.B., 1999, Late v. 37, no. 7, p. 257.
Pleistocene–Holocene retreat of the West Antarctic Ice-Sheet system in Isbell, J.L., Koch, Z.J., Szablewski, G.M., and Lenaker, P.A., 2008, Permian
the Ross Sea: Part 2—Sedimentologic and stratigraphic signature: Geo- glacigenic deposits in the Transantarctic Mountains, Antarctica, in Field-
logical Society of America Bulletin, v. 111, p. 1517–1536, doi: 10.1130/ ing, C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic
0016-7606(1999)111<1517:LPHROT>2.3.CO;2. Ice Age in Time and Space: Geological Society of America Special Paper
Dowdeswell, J.A., Whittington, R.J., and Marienfeld, P., 1994, The origin of 441, p. 59–70.
massive diamicton facies by iceberg rafting and scouring, Scoresby Sund, Jones, A.T., and Fielding, C.R., 2004, Sedimentological record of the late Paleo-
East Greenland: Sedimentology, v. 41, p. 21–35, doi: 10.1111/j.1365-3091 zoic glaciation in Queensland, Australia: Geology, v. 32, p. 153–156, doi:
.1994.tb01390.x. 10.1130/G20112.1.
Dowdeswell, J.A., Mackensen, A., Marienfeld, P., Whittington, R.J., Jennings, Kyle, R.A., 1976, Palaeobotanical studies of the Permian and Triassic Victoria
A.E., and Andrews, J.T., 2000, An origin for laminated glacimarine sedi- Group (Beacon Supergroup) of south Victoria Land, Antarctica [Ph.D.
ments through sea-ice build-up and suppressed iceberg rafting: Sedimen- thesis]: Wellington, New Zealand, Victoria University of Wellington,
tology, v. 47, p. 557–576, doi: 10.1046/j.1365-3091.2000.00306.x. 306 p.
Evans, J., and Pudsey, C.J., 2002, Sedimentation associated with Antarctic Pen- Kyle, R.A., 1977, Palynostratigraphy of the Victoria Group of south Victoria
insula ice shelves; implications for palaeoenvironmental reconstructions Land, Antarctica: New Zealand Journal of Geology and Geophysics,
of glacimarine sediments: Journal of the Geological Society [London], v. 20, p. 1081–1102.
v. 159, p. 233–237, doi: 10.1144/0016-764901-125. Kyle, R.A., and Schopf, J.M., 1982, Permian and Triassic palynostratigraphy
Evans, P.R., 1969, Upper Carboniferous and Permian palynological stages and of the Victoria Group, Transantarctic Mountains, in Craddock, C., ed.,
their distribution in eastern Australia, in Amos, A.J., ed., Gondwana Stra- Antarctic Geosciences: Madison, University of Wisconsin Press, Interna-
tigraphy: Paris, UNESCO, p. 41–54. tional Union of Geological Sciences, p. 649–659.
Eybergen, F.A., 1987, Glacier snout dynamics and contemporary push moraine Lenaker, P.A., 2002, Sedimentology of Permian glacial deposits in the Dar-
formation at the Turtmannglacier, Wallis, Switzerland, in Van Der Meer, win Glacier region, Antarctica [M.S. thesis]: Milwaukee, University of
J.J.M., ed., Tills and Glaciotectonics: Proceedings of the International Wisconsin–Milwaukee, 173 p.
Union for Quaternary Research (INQUA) Symposium, Amsterdam, 1986: Limarino, C.O., Césari, S.N., Net, L.I., Marenssi, S.A., Gutierrez, P.R., and
Rotterdam, A.A. Balkema, p. 217–231. Tripaldi, A., 2002, The Upper Carboniferous postglacial transgression
Eyles, N., 1993, Earth’s Glacial Record and Its Tectonic Setting: Earth-Science in the Paganzo and Río Blanco basins (northwestern Argentina): Facies
Reviews, v. 35, 248 p., doi: 10.1016/0012-8252(93)90002-O. and stratigraphic significance: Journal of South American Earth Sciences,
Fielding, C., Frank, T., Birgenheier, L., Thomas, S., Rygel, M., and Jones, A., v. 15, p. 445–460, doi: 10.1016/S0895-9811(02)00048-2.
2005, Revised Permian glacial record of eastern Australia: Geological Lindsay, J.F., 1970, Depositional environment of Paleozoic glacial rocks in the
Society of America Abstracts with Programs, v. 37, no. 7, p. 256. central Transantarctic Mountains: Geological Society of America Bulletin,
Fielding, C.R., Frank, T.D., Birgenheier, L.P., Rygel, M.C., Jones, A.T., and v. 81, p. 1149–1172, doi: 10.1130/0016-7606(1970)81[1149:DEOPGR
Roberts, J., 2008, Stratigraphic imprint of the late Palaeozoic ice age in ]2.0.CO;2.
eastern Australia: A record of alternating glacial and nonglacial climate Lindström, S., 1995, Early Permian palynostratigraphy of the northern Heime-
regime: Journal of the Geological Society, London, v. 165, p. 129–140. frontfjella mountain-range, Dronning Maud Land, Antarctica: Review
Frakes, L.A., Matthews, J.L., and Crowell, J.C., 1971, Late Paleozoic gla- of Palaeobotany and Palynology, v. 89, p. 359–415, doi: 10.1016/0034
ciation: Part III, Antarctica: Geological Society of America Bulletin, -6667(95)00058-3.
v. 82, p. 1581–1604, doi: 10.1130/0016-7606(1971)82[1581:LPGPIA López-Gamundí, O.R., 1997, Glacial-postglacial transition in the Late Paleozoic
]2.0.CO;2. basins of southern South America, in Martini, I.P., ed., Late Glacial and
Hart, J.K., 1990, Proglacial glaciotectonic deformation and the origin of the Postglacial Environmental Changes: Quaternary, Carboniferous–Permian,
Cromer Ridge push moraine complex, North Norfolk, England: Boreas, and Proterozoic: Oxford, UK, Oxford University Press, p. 147–168.
v. 19, p. 165–180. Mackiewicz, N.E., Powell, R.D., Carlson, P.R., and Molina, B.F., 1984, Interlam-
Hart, J.K., and Boulton, G.S., 1991, The interrelation of glaciotectonic and gla- inated ice-proximal glacimarine sediments in Muir Inlet, Alaska: Marine
ciodepositional processes within the glacial environment: Quaternary Sci- Geology, v. 57, p. 113–147, doi: 10.1016/0025-3227(84)90197-X.
ence Reviews, v. 10, p. 335–350, doi: 10.1016/0277-3791(91)90035-S. Matsch, C.L., and Ojakangas, R.W., 1991, Comparison in depositional style of
Helby, R.J., and McElroy, C.T., 1969, Microfloras from the Devonian and Trias- “polar” and “temperate” glacial ice; late Paleozoic Whiteout Conglomer-
sic of the Beacon Supergroup, Antarctica: New Zealand Journal of Geol- ate (West Antarctica) and late Proterozoic Mineral Fork Formation (Utah),
ogy and Geophysics, v. 12, p. 376–383. in Anderson, J.B., and Ashley, G.M., eds., Glacial Marine Sedimentation;
100 Isbell

Paleoclimatic Significance: Geological Society of America Special Paper Spörli, K.B., 1992, Stratigraphy of the Crashsite Group, Ellsworth Mountains,
261, p. 191–206. West Antarctica, in Webers, G.F., Craddock, G.F., and Splettstoesser, J.F.,
McElroy, C.T., and Rose, G., 1987, Geology of the Beacon Heights area, south- eds., Geology of the Ellsworth Mountains, Antarctica: Geological Society
ern Victoria Land, Antarctica: New Zealand Geological Survey Miscel- of America Memoir 170, p. 21–35.
laneous Series Map 15 and Notes, 47 p., scale 1:50,000. Stow, D.A.V., 2005, Sedimentary Rocks in the Field: A Colour Guide: London,
McKelvey, B.C., Webb, P.N., and Kohn, B.P., 1972, Stratigraphy of the Bea- Elsevier Academic Press, 320 p.
con Supergroup between the Olympus and Boomerang Ranges, Victoria Thomas, G.S.P., and Connell, R.J., 1985, Iceberg drop, dump, and grounding
Land, in Adie, R.J., ed., Antarctic Geology and Geophysics: Oslo, Univer- structures from Pleistocene glacio-lacustrine sediments, Scotland: Journal
sitetsforlaget, p. 345–352. of Sedimentary Petrology, v. 55, p. 243–249.
McKelvey, B.C., Webb, P.N., and Kohn, B.P., 1977, Stratigraphy of the Tay- Tucker, M.E., 1996, Sedimentary Rocks in the Field: New York, Wiley & Sons,
lor and lower Victoria Groups (Beacon Supergroup) between the Mackay 133 p.
Glacier and Boomerang Range, Antarctica: New Zealand Journal of Geol- Turner, S., and Young, G.C., 1992, Thelodont scales from the middle–late
ogy and Geophysics, v. 20, p. 813–863. Devonian Aztec Siltstone, southern Victoria Land, Antarctica: Antarctic
McPherson, J.G., 1978, Stratigraphy and sedimentology of the Upper Devonian Science, v. 4, p. 89–105, doi: 10.1017/S0954102092000142.
Aztec Siltstone, southern Victoria Land, Antarctica: New Zealand Journal van der Meer, J.J.M., Menzies, J., and Rose, J., 2003, Subglacial till; the
of Geology and Geophysics, v. 21, p. 667–683. deforming glacier bed: Quaternary Science Reviews, v. 22, p. 1659–1685,
McPherson, J.G., 1979, Calcrete (caliche) paleosols in fluvial red-beds of the doi: 10.1016/S0277-3791(03)00141-0.
Aztec Siltstone (Upper Devonian), southern Victoria Land, Antarctica: Van der Wateren, F.M., 1987, Structural geology and sedimentology of the
Sedimentary Geology, v. 22, p. 267–285, doi: 10.1016/0037-0738(79 Dammer Berge push moraine, FGR, in van Der Meer, J.J.M., ed., Tills
)90056-3. and Glaciotectonics: Rotterdam, A.A. Balkema, p. 157–182.
Powell, C.M., and Li, Z.X., 1994, Reconstruction of the Panthalassan margin of Van der Wateren, F.M., 2002, Processes of glaciotectonism, in Menzies, J.,
Gondwanaland, in Veevers, J.J., and Powell, C.M., eds., Permian-Triassic ed., Modern and Past Glacial Environments: Oxford, UK, Butterworth-
Pangean Basins and Foldbelts along the Panthalassan Margin of Gondwa- Heinemann, p. 417–443.
naland: Geological Society of America Memoir 184, p. 5–9. Veevers, J.J., 1994, Case for the Gamburtsev Subglacial Mountains of East Ant-
Powell, R.D., 1990, Glacimarine processes at grounding-line fans and their arctica originating by mid-Carboniferous shortening of an intracratonic
growth to ice contact deltas, in Dowdeswell, J.A., and Scourse, J.D., eds., basin: Geology, v. 22, p. 593–596, doi: 10.1130/0091-7613(1994)022
Glacimarine Environments: Processes and Sediments: Geological Society <0593:CFTGSM>2.3.CO;2.
[London] Special Publication 53, p. 53–73. Veevers, J.J., 2001, Atlas of Billion-Year Earth History of Australia and Neigh-
Powell, R.D., and Alley, R.B., 1997, Grounding-line systems: Processes, gla- bours in Gondwanaland: Sydney, GEMOC Press, 76 p.
ciological inferences and the stratigraphic record, in Barker, P.F., and Visser, J.N.J., 1997, A review of the Permo-Carboniferous glaciation in Africa,
Cooper, A.C., eds., Geology and Seismic Stratigraphy of the Antarctic in Martini, I.P., ed., Late Glacial and Postglacial Environmental Changes:
Margin, 2: Washington, D.C., American Geophysical Union, Antarctic Quaternary, Carboniferous–Permian, and Proterozoic: Oxford, UK, Oxford
Research Series, v. 71, p. 169–187. University Press, p. 169–191.
Powell, R., and Domack, E., 2002, Modern glaciomarine environments, in Visser, J.N.J., and Loock, J.C., 1982, An investigation of the basal Dwyka
Menzies, J., ed., Modern and Past Glacial Environments: Oxford, UK, Tillite in the southern part of the Karoo Basin, South Africa: Transac-
Butterworth-Heinemann, p. 361–389. tions—Geological Society of South Africa, v. 85, p. 179–187.
Pyne, A.R., 1984, Geology of the Mt. Fleming area, South Victoria Land, Young, G.C., 1988, Antiarchs (placoderm fishes) from the Devonian Aztec
Antarctica: New Zealand Journal of Geology and Geophysics, v. 27, Siltstone, southern Victoria Land, Antarctica: Palaeontographica, v. 202,
p. 505–512. 125 p.
Ricci Lucchi, F., 1995, Sedimentographica: A Photographic Atlas of Sedimen- Young, G.C., 1989, The Aztec fish fauna (Devonian) of southern Victoria Land:
tary Structures: New York, Columbia University Press, 255 p. Evolutionary and biogeographic significance, in Crame, J.A., ed., Origins
Ritchie, A.R., 1975, Groenlandaspis in Antarctica, Australia and Europe: and Evolution of the Antarctic Biota: Geological Society [London] Spe-
Nature, v. 254, p. 569–573, doi: 10.1038/254569a0. cial Publication 47, p. 43–63.
Rocha-Campos, A.C., Canuto, J.R., and dos Santos, P.R., 2000, Late Paleozoic Young, G.C., 1991, Fossil fishes from Antarctica, in Tingey, R.J., ed., The Geol-
glaciotectonic structures in northern Paraná Basin, Brazil: Sedimentary ogy of Antarctica: Oxford, UK, Oxford University Press, p. 538–567.
Geology, v. 130, p. 131–143, doi: 10.1016/S0037-0738(99)00110-4. Ziegler, A.M., Hulver, M.L., and Rowley, D.B., 1997, Permian world topog-
Scotese, C.R., Boucot, A.J., and McKerrow, W.S., 1999, Gondwanan palaeo- raphy and climate, in Martini, I.P., ed., Late Glacial and Postglacial
geography and palaeoclimatology: Journal of African Earth Sciences, Environmental Changes: Quaternary, Carboniferous-Permian, and Pro-
v. 28, p. 99–114, doi: 10.1016/S0899-5362(98)00084-0. terozoic: Oxford, UK, Oxford University Press, p. 111–146.
Smith, L.M., and Andrews, J.T., 2000, Sediment characteristics in iceberg dom-
inated fjords, Kangerlussuaq region, East Greenland: Sedimentary Geol-
ogy, v. 130, p. 11–25, doi: 10.1016/S0037-0738(99)00088-3. MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

Formation of euxinic lakes during the deglaciation phase


in the Early Permian of East Africa

Thomas Kreuser
Gebretinsae Woldu
Geology Department, University of Asmara, P.O. Box 1220, Eritrea

ABSTRACT

The continental glaciation of Gondwanaland in the Late Carboniferous–Early


Permian left traces in many places in southern and eastern Africa. This paper focuses
on the last glacial advance and consecutive deglaciation leading to the formation of
large euxinic lakes with high concentrations of organic matter. The Idusi Formation
in the Tanzanian Ruhuhu Basin (initiating the Karoo cycle, which extends into the
Triassic) provides the type section for this depositional sequence. It is subdivided into
a lower Lisimba Member, the basal unit of glacial origin, and an upper Lilangu Mem-
ber, characterized by postglacial black shale and rhythmites as evidence of a climatic
amelioration on a large regional scale in Africa. Thickness and facies variations are
attributed to a pronounced paleotopography as the result of scouring glaciers and local
tectonic events. There is a gradual change between the members, reflecting a continu-
ous climatic amelioration and change of sediment supply. The lacustrine environment
was terminated by the onset of braided stream deposition (Mpera Sandstone Mem-
ber); an erosional unconformity between the units marks the start of initial rifting in
the Early Permian. This is followed by the development of extensive coal swamps in a
temperate climate, where organic matter predominated over clastic supply.
Periglacial deposits with tillites and rhythmites, containing dropstones, are over-
lain by glaciolacustrine laminites intercalated with glaciofluvial marginal deltaic
sediments. Deglaciation provided water and accommodation space for the evolution
of extensive anaerobic stratified lakes, which were the focus of prolific deposition of
organic matter. This black shale may contain up to 11% TOC (total organic carbon)
content. Eventually, the lake became shallower and was succeeded by alluvial fan
deposition. The duration of the glaciation and deglaciation was ~20–25 m.y., and the
lacustrine phase lasted ~4–5 m.y. These ages have been verified by palynology (Gran-
ulatisporites confluens Oppel zone).
The hydrocarbon potential of the black shale was estimated by Rock-Eval pyrol-
yses. Hydrogen index, maximum temperature (Tmax), and vitrinite reflection were
used to determine kerogen type, maturity stage, and subsidence history. A promising
potential with respect to gaseous hydrocarbon generation was detected from both
the euxinic black shale and the overlying coals. A comparison with other Tanzanian

Kreuser, T., and Woldu, G., 2010, Formation of euxinic lakes during the deglaciation phase in the Early Permian of East Africa, in López-Gamundí, O.R., and
Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana: Geological Society of America Special Paper 468, p. 101–112,
doi: 10.1130/2010.2468(04). For permission to copy, contact [email protected]. ©2010 The Geological Society of America. All rights reserved.

101
102 Kreuser and Woldu

Karoo basins reveals similar conditions in TOC values and temperature history. The
wide regional extent of the anaerobic lacustrine black shale of the deglaciation event
in several eastern and southern African basins evinces a similar climatic and regional
tectonic framework in the pre-breakup phase of Gondwanaland during the Early
Permian. This period of time may be of some importance in the future when the
economic potential with respect to hydrocarbon generation of the Permian basins is
scrutinized in more detail.

INTRODUCTION thermal history of Karoo source rocks was performed by Kreuser


et al. (1988), Dypvik et al. (1990), Diekmann (1993), and Kreuser
The present work is a general review of the results of a work- (1995b). The present paper summarizes the most important fea-
ing group from the University of Cologne, Germany, and the Uni- tures and incorporates a model of the depositional environment
versity of Dar es Salaam, Tanzania, during a research period of and thermal history of these potential source rocks for hydrocar-
more than 15 yr. Therefore, most of the findings and interpreta- bon generation.
tions given in the following chapters have already been published Within the Karoo succession, several phases of lacustrine
elsewhere. The focus of this paper was mainly to collect data development occurred in which organic matter accumulated
from different working groups and combine them into a general (Kreuser, 1995b); however, only the periglacial to postglacial
description of the late Paleozoic glacial deposition, the onset of development is highlighted in this paper. Similar sedimen-
Permian rifting, and the associated history of lacustrine deposi- tary successions were recorded from neighboring countries
tion which initiated the onset of the Karoo Supercycle in southern in Africa: South Africa (Rust, 1975), Madagascar (Besairie,
and eastern Africa. 1972), Congo (Boutakoff, 1948), Ethiopia (Worku and Astin,
Continental glaciation in Late Carboniferous to Early Perm- 1992), and Oman (Qidwai, 1988), which are mentioned here
ian times occurred at numerous locations on the African conti- for comparative purposes. The Tanzanian succession serves
nent and other former Gondwanan continents. In Africa, glacial as a reference section, which is comparatively well described
and postglacial deposition represents the onset of the Karoo dep- and analyzed.
ositional cycle, which continued until the Early Triassic. In South
Africa the glacial deposits are referred to as the Dwyka Forma- STRATIGRAPHY
tion, which has been studied intensively by numerous authors
(Rust 1975; Martin, 1981; Visser, 1989); in Tanzania the suc- The southwestern Tanzanian Karoo basins comprise diverse
cession was described by Wopfner and Kreuser (1986), Kreuser facies successions of glacial and postglacial deposits (Fig. 1):
(1987), and Wopfner and Diekmann (1996). In East Africa the 1. Ruhuhu Basin (subdivided into a western Mchuchuma
glaciogene sediments are best exposed in southwestern Tanzania, sub-basin and an eastern Ngaka sub-basin) along the
where the early Karoo deposits are exposed along the uplifted rift eastern fault scarp of the Malawi Rift. Post-Karoo move-
shoulders of Lake Malawi and Lake Rukwa. ments have dissected the Ruhuhu Basin into several half
This paper summarizes the lithological and depositional grabens, i.e., the Lumecha sub-basin. Owing to the south-
history of a number of Tanzanian basins: the Ruhuhu, Songwe eastern tilt of the half grabens, outcrops of the Idusi For-
Kiwira, and Galula Basins, with respect to their glacial and post- mation lie mainly on the northern and western margins of
glacial depositional history. Special emphasis is placed on the these blocks.
accumulation of organic matter in large periglacial and postgla- 2. Songwe-Kiwira Basin, situated northwest of the Ruhuhu
cial lakes and the thermal history of these organic source rocks. Basin.
Additionally, a stratigraphic approach is presented to compare 3. Galula Basin, along the southern boundary of the Rukwa
the Tanzanian sections with other southern and eastern African Rift (Dypvik et al., 1990; Mliga, 1994).
localities of comparable age and focus on the ubiquitous onset of The type section of the glaciogenic sequence in Tanzania is
glacial and postglacial deposits in the region. in the Ruhuhu Basin in the gorge of the Idusi River in the north-
The glacial nature of these deposits was first established western Mchuchuma sub-basin and is known as the Idusi Forma-
by Spence (1957). Later, Wopfner and Kreuser (1986), Kreuser tion (Kreuser et al., 1990; Fig. 2). Comparable stratotypes were
(1987), Diekmann (1993), Wopfner and Diekmann (1992), and detected along the Ketewaka River in the northeastern Mchu-
Wopfner and Diekmann (1996) described these sequences in chuma sub-basin and along the Nyamangami River of the western
detail and established a modern nomenclature and a depositional Ngaka sub-basin (Fig. 2). Additional sections of the Idusi Forma-
model. The most detailed lithologic investigation with respect tion were obtained from logs of exploration boreholes drilled by
to facies distribution and regional differentiation was performed CDC (Colonial Development Corp.) and MADINI (Geological
by Diekmann (1993), when the formal lithostratigraphic termi- Survey of Tanzania) and were described by Wopfner and Diek-
nology was established. Geochemical analysis of the nature and mann (1996).
Formation of euxinic lakes, East Africa 103

30˚ 34˚ 38˚E


UGANDA AFRICA

O
Tanzania

NG
CO
Lake
Victoria Nairobi
RWANDA
2˚S

Kenya
BURUNDI basin
DEM. REP

INDIAN
Tanga
OCEAN
basin

TANZANIA
Lake
Tanganyika Mikumi basin ZANZIBAR 6˚

Namwele-Mkomolo Dares Salaam

Rukwa basin Ruvu


Valley
Muse
Lake Rufiji valley
Rukwa

LEGEND Galula
Karoo basins Songwe-Kiwira tro egu
h
ug
w
Ruhuhu
Lu

Post-Karoo basins Ruvuma 10˚


basin
basin
Lake Nyasa
WI

Faults
IA

MALA

Border Njuga
ZAMB

Mhukuru
Maniamba
0 250 500 km Mbamba basin
Bay MOZAMBIQUE

Figure 1. Regional distribution of continental Karoo (Permian-Triassic) rocks in Tanzania and neighboring countries.

At the type locality in the Idusi River gorge the Idusi For- cambrian basement (Kreuser, 1987). No palynological evidence
mation reaches a thickness of 240 m. Because of facies varia- has been found in the lower Lisimba Member, and owing to the
tions, Diekmann (1993) subdivided the Idusi Formation into two stratigraphic position of the younger Lilangu Member, the lower
lithostratigraphic units, a lower Lisimba Member and an upper member may have been deposited in Late Carboniferous times
Lilangu Member. (Diekmann, 1996).
The basal unit of the Lisimba Member consists of massive
Lisimba Member diamictite, grading into faintly bedded diamictite. Higher in the
section, siltstone with dropstones up to 60 cm in diameter is pres-
This member is named after a tributary of the Idusi River ent. The middle portion of the section consists of siltstone and
(thickness, 170 m, Fig. 3). The lower succession consists of dif- mudstone with slumping structures. The upper part consists of
ferent lithotypes of diamictite, conglomerate, and sandstone as olive green laminites with a few dropstones and sandstones.
well as silty mudstone, exhibiting large dropstones and lone-
stones (Diekmann, 1996). Higher in the succession, laminites Lilangu Member
are dominant in a mudstone matrix with abundant lonestones.
The Lisimba Member exhibits a typical olive green color caused Named after a tributary of the Idusi River (thickness, 70 m,
by abundant chlorite and has an arkosic composition. The thick- Fig. 3), the Lilangu Member developed gradually from the
ness ranges up to 420 m owing to a strong paleorelief of the Pre- glaciogenic succession beneath and is separated locally by an
104 Kreuser and Woldu

Figure 2. Distribution of sub-basins and geology of glacial and postglacial deposits in the Ruhuhu Basin. Note the
fault-controlled basin margins along the northeast trending blocks.

unconformity from the overlying Mchuchuma Formation (Sem- blages assigned to the Granulatisporites confluens Oppel zone,
kiwa, 1992). The boundary between the Lisimba and Lilangu indicating a late Asselian to early Sakmarian age (Wopfner and
Members is characterized by the first black fissile shale that is Kreuser, 1986).
highly pyritic, containing calcareous concretions (Wopfner and
Diekmann, 1996). Above this are black organic-rich siltstones, Regional Lithological Variations
locally bituminous, intercalated with sandstone lobes and brec-
cias. The amount of organic matter is high (kerogen types III The reference section along the Ketewaka River of the Idusi
and IV, from 3% to 6%). Palynological analyses indicate assem- Formation measures ~100 m in thickness and exhibits faceted
Formation of euxinic lakes, East Africa 105

Figure 3. Composite log of the Idusi Formation at the type locality in


the Idusi gorge and the northwestern Mchuchuma sub-basin. Legend
is attached (for location, see Fig. 2). Based on Wopfner and Diekmann
(1996).

by sandstone with cross-bedding, plus siltstone and conglomer-


ate that contain cobble-size lonestones. Diekmann (1993) termed
this section the Ndongosi facies after the escarpment it forms, and
interpreted it as glaciofluvial in origin. This is overlain by 160 m
of massive deltaic sandstones, referred to as the Nanderuka facies
(Diekmann, 1993). The Lilangu Member reaches a maximum
thickness of ~300 m; common intercalations of mudstone are
present, and a generally higher sand to shale ratio is observed.
Both south and north of the “long section” the Idusi For-
mation is characterized by a considerably condensed succession
with a thickness of up to 40 m. For that reason, other authors
have used the term “short section” (McKinlay, 1954). The three-
fold lithologic subdivision and most of the described lithofacies
are recognizable (Fig. 5). However, a section north of the Mkapa
River exhibits a massive basal diamictite with faceted and stri-
and striated diamictites at the base and continues in a similar way ated clasts of various diameters resting on basement, interpreted
along the Idusi River into laminites farther upward (Fig. 4). The as a lodgment till (Diekmann, 1996). It is followed by reworked
Lilangu Member, however, exhibits a different facies develop- diamictite and green bedded sandstones. The Lilangu Member
ment. Although the dominant black, organic-rich shale is pres- has the typical black shales at the base, but the upper part is usu-
ent, intercalations with sandy lobes, several of which attain 10 m ally absent, probably having been removed by erosion prior to
in thickness, are present, with an overall increased thickness of deposition of the Mchuchuma Formation (Diekmann, 1996).
120 m. Diekmann (1993) termed these sediments the Muhimbi
facies, named after a tributary of the Ketewaka River. Songwe Kiwira Basin
In the Ngaka sub-basin (Fig. 1) the “long section” of the
Idusi Formation reaches a thickness of >700 m. In contrast to cor- Approximately 50 km west of the northern end of Lake
relative lithologic successions, the lower 140 m is characterized Malawi (Fig. 1), another small basin appears, named after the
106 Kreuser and Woldu

paleovalley filled with basal diamictites interfingering with sand-


stone and conglomerate, capped by a few meters of laminites.

Galula Basin

This small basin, along the southwestern flank of the Rukwa


Rift (Fig. 1), is an erosional remnant of Karoo strata. Here, Karoo
strata are preserved along the depressed side of tilted fault blocks.
Mbede (1993) noted seismic surveys indicating that in the deep-
est part of the Rukwa graben, some 7 km below ground level,
5 km of Karoo strata is present. There is only one outcrop of
the Idusi Formation along the Chizi River, of ~13 m thickness
(Wopfner and Diekmann, 1996). The succession begins with a
basal diamictite, overlain by laminites of the Kipololo facies,
topped by laminated mudstones with microclasts. The Lilangu
Member is not developed here.

Mhukuru Basin

Located 65 km south-southwest of Songea near the Mozam-


bique border (Fig. 1), this basin contains a well-developed succes-
sion of the Idusi Formation, with a maximum thickness of ~120 m.
The Lisimba Member exhibits basal diamictites with dropstones
and sandstones and green laminites. The overlying Lilangu Mem-
ber consists of 20 m of black shales (Diekmann, 1993).

DEPOSITIONAL ENVIRONMENT

Lisimba Member

The basal diamictites consist of massive to faintly bedded,


poorly sorted mixtures of clasts resting in a sandy-silty matrix.
Clasts, derived from local basement rocks, are faceted and locally
striated, which is characteristic for glacial transport (Wopfner
and Kreuser, 1986). In this paper, diamictites are interpreted as
subaqueous melt-out and flow tillites, intercalated with meltwa-
ter deposits. Lodgment tills rarely occur. Depositional evidence
Figure 4. Log of the Ketewaka section in the northwestern Mchuchuma for this interpretation is given in detail in Wopfner and Kreuser
sub-basin. See Figure 3 for legend. Based on Wopfner and Diekmann (1986). In the long section they are replaced by an upwardly
(1996). fining succession, starting with conglomerate and ending with
mudstone. This is interpreted as a transition from braided-river
to subaquatic conditions in a proglacial environment (Diekmann,
1993; Fig. 6).
main rivers Kiwira and Songwe, which cross it. The southern part The dominant lithology of the Lisimba Member is charac-
of the basin continues into Malawi, exhibiting an originally larger terized by silty and sandy mudstones containing rare lonestones
areal extent of Karoo deposits. The northern part of the basin is with diameters up to 80 cm. The lonestones are interpreted as
covered by Quaternary basalts of the Rungwe volcanics. The ice-rafted dropstones. The fine clastics were settling from melt-
Idusi Formation is restricted to parts of the Lisimba Member, water plumes as suspended particles. A few ripple laminations
starting with a basal diamictite with cobbles that locally reach and sandy layers are current indicators that were formed during
30 cm in diameter. This unit is interpreted as a lodgment tillite subaqueous flows (Diekmann, 1996).
(Dypvik et al., 1990). Facies variations can be studied in borehole Laminites consist of alternating silt-clay couplets (8–15 cm)
sections drilled by the Coalfield Exploration Team of the Peoples with common deformation structures, and some microclasts and
Republic of China (Anonymous, 1979). Wopfner and Diekmann rare dropstones are also present, which were deposited as perigla-
(1996) presented a depositional model that shows a 60-m-deep cial varves during seasonal changes from a thermally stratified
Formation of euxinic lakes, East Africa 107

Figure 5. Correlation chart of three sections of the Idusi Formation in the Ruhuhu Basin, displaying a typical threefold subdivision. In addition
to the depositional environments, the successive climatic changes are seen at the right. Modified from Diekmann (1993).

water column (Diekmann, 1993). Diekmann interpreted the detected. These siltstones contain fossil-wood remains and rem-
laminites as sandy layers from proximal positions that developed nants of Gangamopteris. Rhythmites are intercalated with the
during meltwater discharge during times of episodically slightly massive siltstones and consist of couplets of silt and medium
higher temperatures (Fig. 6). sand laminae. This succession is interpreted as sapropel-rich
deposits of a deep-lake environment with restricted circulation
Lilangu Member in a thermally stratified water column (Diekmann, 1993). Rare
dropstones probably signal seasonal ice rafts on the lake, with
No sharp boundary could be discerned between the Lisimba common graded bedding. Some lenticular sandstone bodies are
and Lilangu Members. A transitional zone exists where the silt intercalated into the black siltstones, which exhibit sole marks
gradually disappears and is replaced by black to olive gray shale (flute and groove casts) and current lineations. These sandstones
with a high TOC (total organic carbon) content of up to 0.5%. are interpreted as turbidites accumulating at the apex of sub-
Early diagenetic concretions are commonly observed wherein aquatic channels on delta fronts (Diekmann, 1993; Fig. 6).
pyrite replaces plant remains. These basal shales represent a deep- Toward the top of the member an increase in sand-size
water lacustrine environment during a postglacial stage (Diek- grains was detected, which developed parallel with a decrease
mann, 1993). in TOC. Wopfner and Diekmann (1996) interpreted these silty
Overlying the basal shales are black, organic-rich siltstones, sandstones as part of a shallow lake environment, which alter-
reflecting a high amount of terrestrial organic debris that con- nated with upper-flow-regime sandstone deposits accompanied
sists of kerogen types III and IV. TOC values of up to 12% were by lag deposits with common mud clasts.
108 Kreuser and Woldu

Figure 6. Paleogeographic model of the depositional environment during deglaciation time. (A) Periglacial setting with
ice rafts (1), valley glaciers (2), glacial outwash (3), reworked glaciofluvial residue (4). (B) Postglacial lake model with
laminites, rhythmites, lakeshore flora, and associated marginal deltaic lobes.

PALEOGEOGRAPHY ciation. This climatic amelioration is evinced in the lithologic


development of the Idusi Formation irrespective of local facies
Records of the Permo-Carboniferous glaciation in Africa are or thickness variations. The correlation between lithology and
in most cases incomplete, and commonly only the last glacial depositional environment is shown in Figure 5.
advance was spared from erosion. Deeply incised paleovalleys The Lisimba Member, characterized by massive diamictite,
around the centers of glaciation are known from several locali- records the last glacial advance and subsequent retreat. These are
ties, i.e., from South Africa (Visser, 1989, 1997), Namibia (Mar- end-glacial or periglacial deposits that are overlain by a proximal
tin, 1981), and Congo (Boutakoff, 1948). From other Gondwanan glaciolacustrine succession intercalated with glaciofluvial and
continents, comparable features have been recorded—e.g., Cam- marginal deltaic deposits. The middle part of the Lisimba Mem-
pos (1994) from Brazil, and Woolfe (1994) from Antarctica. In ber is characterized by mudstones with dropstones that record a
Tanzania a number of such tectonically controlled paleoval- distal facies of a proglacial lake environment.
leys have preserved much of the sedimentary record, the most The final deglaciation is characterized by the deposition of
impressive of these being the Ruhuhu Basin, with almost 900 m fine-grained, organic-rich sediments of the Lilangu Member,
of sediments. This is the only region where such a thick succes- which took place in a large, shallow lake that developed as an
sion was preserved in Tanzania, attributed probably to thermal expansion of formerly proglacial to periglacial lake systems.
subsidence during the early phase of rifting in the Early Permian Plant remains and sporadic desiccation cracks exhibit times of
of East Africa. exposure when the lake was reduced in extent and depth. This
The Idusi Formation appears to record the last glacial large lake not only filled the paleorelief but overstepped onto
advance and the succeeding stages of glacial retreat and degla- higher areas where it was directly in contact with Precambrian
Formation of euxinic lakes, East Africa 109

basement. During successive deglaciation and exposure of the an erosional hiatus separates Precambrian to Permian strata in
lowland areas, which were increasingly vegetated, deposition the Majunga Basin (Besairie, 1972). For the Congo, basal dia-
switched to plant debris and silty-clayey sediments that culmi- mictites are recorded and interpreted as being of glacial origin
nated in the accumulation of organic carbon–rich black shale. (Boutakoff, 1948).
These deposits took place in an anaerobic environment at the lake The black shale above the diamictite is not of such a ubiqui-
bottom, interchanging at marginal positions with deltaic lobes tous distribution. In Zambia the Mukumba Member shows simi-
and subaqueous fans (Fig. 5). larities to the Lilangu Member as being part of the Luwumba
The top of the Lilangu Member records the filling of the Formation in the Luangwa Valley, both separated by an uncon-
lake and subsequent alluvial fan deposition. Wopfner (1996) formity (Kreuser et al., 1990). In Malawi the black shales are
interpreted the angular unconformity between the Idusi Forma- present in the Nkana and Livingstonia Basins, separated from
tion and the overlying Mchuchuma Formation (Mpera Sandstone the underlying tillites by an unconformity (Kreuser et al., 1990).
Member) as a basinwide tectonic event, which in the Galula and Also in Madagascar, black shales follow the tillites in the Moron-
Songwe-Kiwira Basins led to erosion and removal of the sedi- dava Basin (Besairie, 1972). In the Congo the tillites are over-
ments of the Lilangu Member. lain by two separate black shale zones that locally interfinger
with fluvial sandstones, both separated by local unconformities
REGIONAL TECTONIC CONTROL AND (Kreuser, 1995a).
APPROXIMATE DATING The Lilangu Member is partly developed in those basins
which survived later tectonic exhumation. Locally it was not
The deposition of black shales during the euxinic stages of deposited at all, but deposition evolved from a glacial to a peat
lake development of the Ruhuhu Basin and other early Karoo swamp environment, marked by either a rapid transition or a
basins was a significant event that can be traced regionally across regional unconformity between the glacial and postglacial units.
East Africa. The abundance of organic matter led to anaerobic However, the viability of much of the data presented in the lit-
conditions in those regionally extensive deposits. Alternatively, erature is not convincing, as much of the evidence has not been
anaerobic conditions may have led to extensive preservation of verified by detailed lithological descriptions; thus some of the
organic matter. Postglacial conditions were characterized by a stratigraphic nomenclature is not up to date with lithostratigraphic
climatic amelioration and represented a transition from cold and correlations. Additionally, dating by palynology or paleontology
arid to temperate and humid conditions. Shortly thereafter, coal- is poor for these continental deposits, and much more work is
rich cyclothems prevailed, which recorded seasonal variations needed in order to establish a reliable litho-chronological frame-
during deposition of the Mchuchuma Formation (Kreuser, 1991). work. Apparently the Karoo basins farther east in the vicinity of
Coal deposition commenced as an almost simultaneous event the coastal area were never reached by the Paleozoic glaciation,
throughout entire Gondwanaland, which likely contributed to a or at least are not exposed at the surface (Kreuser et al., 1990).
lowering of atmospheric CO2 (Wopfner and Diekmann, 1996). No basal diamictites and succeeding euxinic black shales are
The unconformity between the underlying Lilangu Mem- known from the coastal basins of Tanzania, Kenya, and Somalia
ber and the overlying Mpera Sandstone is evidence for a basin- (Mbede, 1997).
wide tectonic event that probably led to the complete erosion of Wopfner and Diekmann (1996) suggested a duration for the
the black shale in the Galula and Songwe-Kiwira Basins. The deposition of the Lilangu Member of ~5–6 m.y. In comparison
Mpera Sandstone is interpreted as braided stream deposits that with the approximate duration of the glaciation, which lasted
initiated the second depositional sequence of the Karoo succes- ~20–25 m.y., deglaciation would have taken one quarter of the
sion (Kreuser et al., 1990). This second unit, the Mchuchuma overall time. The revised estimate is based on palynological anal-
Formation (Fig. 5), was dominated by coal swamps and is also yses that indicate a late Asselian to early Sakmarian age (Granu-
important for the source rock assessment of the Karoo succession latisporites confluens Oppel zone; Wopfner and Kreuser, 1986)
in Tanzania (Kreuser et al., 1988). for the Lilangu Member of the Idusi Formation.
On a regional scale in Africa south of the Sahara, glacio-
genic and postglacial deposits are known from South Africa, ORGANIC GEOCHEMISTRY OF EUXINIC
Zimbabwe, Zambia, Malawi, Kenya, Madagascar, Congo, and LACUSTRINE BLACK SHALES OF THE LILANGU
Mozambique. The basal diamictite is present in Zimbabwe MEMBER
(Mid-Zambezi and Sabi-Lunde Basins; Kreuser et al., 1990). In
Zambia glacial deposits are known from the Gwembe, Luano, Several authors analyzed samples with high organic contents
Luangwa, and Barotse Basins, respectively (Kreuser et al., 1990). from Lower Permian lacustrine strata from East Africa in order to
In Malawi there is some evidence for the glacial character of the estimate their hydrocarbon potential (Kreuser et al., 1988; Dypvik
basal beds. In Kenya some indicators point to a glacial origin of et al., 1990; Kagya, 1991; Diekmann, 1993). Only a few samples
the basal beds (Martin, 1981), and a similar record was noted for were collected from the postglacial deposits, however; most of the
Mozambique (Kreuser, 1995b). There is no doubt that a glacial samples came from coal seams in the Mchuchuma Formation or
character exists in Madagascar in the Morondava Basin, whereas from Upper Permian deposits in rift lake environments (Mpanju,
110 Kreuser and Woldu

1999). For the sake of comparison, some of these results are men- (0–224 mg HC/g TOC). Tmax is in the range of 430 °C to 458 °C
tioned here as well, but our focus is on the black shale of the (Fig. 7C). Unfortunately, nothing is known about the maceral
Lilangu Member. composition and maturity derived from these vitrinite reflection
The hydrocarbon potential of the Lilangu Member black measurements. No specific information about the stratigraphic
shale was evaluated by Rock-Eval pyrolyses, and the hydrogen position is available, so it is not known if the black shale was
index (HI in mg HC/g TOC) was measured. We also analyzed the sampled or is the age equivalent of the Lilangu Member in the
Tmax (°C) values during successive heating of samples, and this Ruhuhu Basin. Also Mpanju (1999) analyzed the hydrocarbon
was matched with vitrinite reflection data. potential of the Karoo coals in the southwestern rifts of Tanza-
Temperature ranges (Rock-Eval pyrolyses) for most of the nia, and his results are generally in accordance with the analyses
samples were from 425 °C to 460 °C, with the black shale from described above.
the Lilangu Member showing the highest temperatures. These Kreuser (1995b) calculated a subsidence history for the
samples exhibit a fair hydrocarbon generation potential, ranging Ruhuhu Basin using vitrinite reflection, palynological dating,
from 65 to 150 mg HC/g TOC. Some TOC values reach 11%, but and lithostratigraphic correlation with an inferred geothermal
most are <5% (Fig. 7A). The kerogen of the dispersed organic gradient of 25 °C km–1. Maximum subsidence of the basal glacial
matter in the black shale is dominantly type III. Vitrinite reflec- units of 2.6 km was attained during the Early Triassic. Afterward
tion values of the black shale average 0.97%, which is a little there might have been a slight uplift of several hundred meters
lower than expected from pyrolysis data (Kreuser, 1995b). Mac- during the breakup unconformity of Gondwanaland in Middle
erals are dominated by inertinite and vitrinite (Semkiwa et al., Jurassic time. A strong uplift occurred only during the late Mio-
2003). In comparison, coals of the Mchuchuma Formation show cene, which brought basal rocks to the surface along the uplifted
HI values up to 300 mg HC/g TOC with slightly lower Tmax val- basin margins. Several other Karoo rifts in Tanzania, and prob-
ues (Fig. 7B). The generation potential from the black shale ably in neighboring areas, were exposed to temperatures condu-
is mainly gas-prone, depending on the maceral composition. cive to the formation of light hydrocarbons. This has not been
However, lamalginite (the most oil-prone liptinite maceral) and proven for many rifts, however, and it might be possible that
exsudatinite (remobilized bitumen) indicate the onset of bitumi- lower geothermal gradients and subsidence values for the rift
nization and liberation of mobile hydrocarbons, both present in centers were more favorable locally for the generation of liquid
the black shale below the coals (Semkiwa et al., 2003). hydrocarbons. In turn, these hydrocarbons might have migrated
Sediments and coals from the Ilima colliery in the Songwe- into overlying porous rock that acted as suitable reservoirs and/or
Kiwira Basin and in the Rukwa Basin, measured by Dypvik et al. traps for hydrocarbon accumulation. More research is needed to
(1990), show fair HI values, although they exhibit high variations shed more light on this matter.

Figure 7. Composite HI (hydrogen index) versus Tmax (maximum temperature) diagrams for interpretation of kerogen type and maturity level
of organic-rich source rocks. (A) Lilangu Member of the Ruhuhu Basin in Tanzania (modified from Diekmann, 1993). (B) Sediments and coals
from the Permian Mchuchuma Formation. Open circles, Ketewaka sub-basin; full circles, Ngaka sub-basin; full squares, Mchuchuma sub-basin.
(C) Sediments and coals from the Songwe-Kiwira and Rukwa Basins (adapted from Dypvik et al., 1990). TOC—total organic carbon.
Formation of euxinic lakes, East Africa 111

CONCLUSIONS Dypvik, H., Nesteby, H., Ruden, F., Aagard, P., Johansson, T., Msindai, J.,
and Massay, C., 1990, Upper Paleozoic and Mesozoic sedimentation in
the Rukwa-Tukuyu region, Tanzania: Journal of African Earth Sciences,
The glacial and postglacial deposits of Tanzania and other v. 11, p. 437–456, doi: 10.1016/0899-5362(90)90022-7.
countries of eastern and southern Africa show a remarkable Kagya, M.J., 1991, The source rock potential of the Nyasa rift basin, oil shows
homogeneity in lithology and distribution. The lower part of the in Tanzania: Journal of Southeast Asian Earth Sciences, v. 5, p. 407–419,
doi: 10.1016/0743-9547(91)90055-3.
sequence was deposited in many basins of the region, mainly in Kreuser, T., 1987, Late Paleozoic glacial sediments and transition to coal bear-
incised paleovalleys, and later was exposed owing to uplift and ing Lower Permian in Tanzania: Fazies, v. 17, p. 149–158.
tilting during Tertiary rift development. The twofold subdivision Kreuser, T., 1991, Facies evolution and cyclicity of alluvial coal deposits in the
Lower Permian of East Africa (Tanzania): Geologische Rundschau, v. 80,
into a lower diamictite (Lisimba Member) and an upper lacus- p. 19–48, doi: 10.1007/BF01828766.
trine euxinic postglacial unit (Lilangu Member) has been well Kreuser, T., 1995a, Tectonics and climatic controls of lacustrine sedimentation
established in Tanzania and can be correlated with many neigh- in pre-rift and rift settings in the Permian-Triassic of East Africa: Journal
of Paleolimnology, v. 13, p. 3–19, doi: 10.1007/BF00678108.
boring successions in the region. Kreuser, T., 1995b, Rift to drift evolution in Permian-Jurassic basins of East
During early Permian times, a high TOC content in the Africa, in Lambiase, J., ed., Hydrocarbon Habitat in Rift Basins: Geologi-
black shale of the Lilangu Member records a euxinic lake envi- cal Society [London] Special Publication 80, p. 297–315.
Kreuser, T., Schramedei, R., and Rullkötter, J., 1988, Gas prone source rocks
ronment in a postglacial setting in most of southern and eastern from cratogene Karoo basins in Tanzania: Journal of Petroleum Geology,
Africa. These black shales represent source rocks for the genera- v. 11, p. 169–184, doi: 10.1111/j.1747-5457.1988.tb00811.x.
tion of natural gas and some liquid oil and are within a favorable Kreuser, T., Wopfner, H., Kaaya, C.Z., Markwort, S., Semkiwa, P.M., and
Aslanidis, P., 1990, Depositional evolution of Permo-Triassic basins in
temperature window. Even though they probably were deposited Tanzania with reference to the economic potential: Journal of African
ubiquitously in most of the Permian basins of the region, distribu- Earth Sciences, v. 10, p. 151–167, doi: 10.1016/0899-5362(90)90052-G.
tion was also controlled by the regional and local tectonic frame- Martin, H., 1981, The Late Paleozoic Gondwana glaciation: Geologische Rund-
schau, v. 70, p. 480–496, doi: 10.1007/BF01822128.
work. The transition from a euxinic to lacustrine environment of Mbede, E., 1993, Tectonic Development of the Rukwa Rift Basin in SW Tanza-
a cool climate to a moderately warm climate, where coal swamps nia: Berliner geowissenschaftliche Abhandlungen, Reihe A, 152, 92 p.
Mbede, E., 1997, The coastal basins of Somalia, Kenya and Tanzania, in Selley,
developed, is often disguised by an erosional hiatus. R.C., ed., Sedimentary Basins of the World, v. 3: Amsterdam, Elsevier,
In combination with the overlying coal deposits, these p. 211–233.
organic-rich deposits represent a potential source rock for the McKinlay, A.C.M., 1954, Geology of the Ketewaka-Mchuchuma Coalfield,
Njombe District: Geological Survey of Tanganyika Bulletin 21, 46 p.
entire East and South African region, which could be of some Mliga, N.R., 1994, Depositional environment, stratigraphy and hydrocarbon
future economic potential. potential of the Rukwa rift basin, SW Tanzania [Ph.D. thesis]: Durham,
North Carolina, Duke University, 172 p.
Mpanju, F.J., 1999, Hydrocarbon potential of Karoo coals and associated rocks
ACKNOWLEDGMENTS of the SW rift basins of Tanzania: Journal of African Earth Sciences, v. 28,
p. 52–53.
The authors are indebted to the sponsoring institutions of Qidwai, H.A., 1988, Evidence of Permian–Carboniferous glaciation in the basal
Murbat Sandstone Formation, southern region, Sultanate of Oman: Jour-
Deutsche Forschungsgemeinschaft and Deutscher Akademis- nal of Petroleum Geology, v. 11, p. 81–88, doi: 10.1111/j.1747-5457.1988
cher Austauschdienst and are grateful for the collaboration with .tb00802.x.
the former working group members Heli Wopfner, Berni Diek- Rust, I.C., 1975, Tectonic and sedimentary framework of Gondwana basins in
southern Africa, in Campbell, C.S.W., ed., Gondwana Geology: Sympo-
mann, Pascal Semkiwa, Charles Kaaya, and Stefan Markwort. sium, 3rd, Canberra, Australia, p. 537–564.
The first author is grateful for the convener of International Geo- Semkiwa, P.M., 1992, Depositional Environment and Coal Petrography of
logical Congress 34 in Florence, Italy, Oscar López-Gamundí, Permian Coal Deposits in Karoo Basins of SW Tanzania: Köln, Geologis-
ches Institut der Universität Köln, Sonderveröffentlichungen 84, 184 p.
for the invitation to publish in this GSA Special Paper. Semkiwa, P.M., Kalkreuth, W., Utting, J., Mpanju, F., and Hagemann, H., 2003,
The geology, petrology, palynology and geochemistry of Permian coal
REFERENCES CITED basins in Tanzania: 2. Songwe-Kiwira Coalfield: International Journal of
Coal Geology, v. 55, p. 157–186, doi: 10.1016/S0166-5162(03)00108-3.
Spence, J., 1957, The Geology of Part of the Eastern Province of Tanganyika:
Anonymous, 1979, Coalfield Geological Exploration Team of the P.R. of China, Geological Survey of Tanganyika Bulletin 84, 126 p.
1979, report on the geological exploration of the Songwe-Kiwira coal- Visser, J.N.J., 1989, The Permo-Carboniferous Dwyka Formation of Southern
field: Beijing, United Republic of Tanzania, unpublished report, 223 p. Africa: Deposition by a predominantly subpolar marine ice sheet: Palaeo-
Besairie, H., 1972, Géologie de Madagascar. I. Les terrains sédimentaires: geography, Palaeoclimatology, Palaeoecology, v. 70, p. 377–391, doi:
Annales Géologiques de Madagascar, Fascule 35, Tananarive. 10.1016/0031-0182(89)90115-6.
Boutakoff, N., 1948, Les Formations Glacières et Post-Glacières Fossilifères Visser, J.N.J., 1997, A review of the Permo-Carboniferous glaciation in Africa,
d’Age Carbonifère de la Région Walikale (Kivu, Congo Belge): Louvain, in Martini, I.P., ed., Late Glacial and Postglacial Environmental Changes,
Belgium, Institut Géologie Université Louvain, Memoire 9, 124 p. Quaternary, Carboniferous-Permian and Proterozoic: Oxford, UK, Oxford
Campos, J.E., 1994, A glaciaçao Neopalozoica na porçao meridional da Bacia University Press, p. 169–191.
Sanfranciscana: Revista Brasileira de Geociencias, v. 24, p. 65–76. Wopfner, H., 1996, The late Paleozoic Idusi Formation of SW Tanzania. Record
Diekmann, B., 1993, Paläoklima und glazigene Karoo-Sedimente des späten of change from glacial to post-glacial conditions: Journal of African Earth
Paläozoikums in SW-Tansania: Köln, Geologisches Institut Universität Sciences, v. 22, p. 575–595, doi: 10.1016/0899-5362(96)00038-3.
Köln, Sonderveröffentlichungen, 90, 193 p. Wopfner, H., and Diekmann, B., 1992, Neue Ergebnisse aus der spätpaläozois-
Diekmann, B., 1996, Petrographic and diagenetic signatures of climatic changes chen glazigenen Abfolge in der Karoo Tansanias, Zentralblatt Geologie
in peri- to post-glacial Karoo sediments of SW-Tanzania: Palaeogeogra- Paläontologie: Teil 1, p. 2689–2700.
phy, Palaeoclimatology, Palaeoecology, v. 125, p. 5–25, doi: 10.1016/0031 Wopfner, H., and Diekmann, B., 1996, The Late Paleozoic Idusi Formation of
-0182(95)00084-4. SW Tanzania: A record of change from glacial to postglacial conditions:
112 Kreuser and Woldu

Journal of African Earth Sciences, v. 22, p. 575–595, doi: 10.1016/0899 Worku, T., and Astin, T.R., 1992, The Karoo sediments (Late Paleozoic to
-5362(96)00038-3. Early Jurassic) of the Ogaden Basin, Ethiopia: Sedimentary Geology,
Woolfe, K.J., 1994, Cycles of erosion and deposition during Permo-Carbonif- v. 76, p. 7–21, doi: 10.1016/0037-0738(92)90136-F.
erous glaciation in the Transantarctic Mountains: Antarctic Science, v. 6,
p. 93–104, doi: 10.1017/S095410209400012X.
Wopfner, H., and Kreuser, T., 1986, Evidence for Late Palaeozoic glaciation in
southern Tanzania: Palaeogeography, Palaeoclimatology, Palaeoecology,
v. 56, p. 259–275, doi: 10.1016/0031-0182(86)90098-2. MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

Stratigraphic and paleofloristic record of the Lower Permian


postglacial succession in the southern Brazilian Paraná Basin

Roberto Iannuzzi*
Paulo A. Souza*
Michael Holz*
Departamento de Paleontologia e Estratigrafia, Instituto de Geociências, Universidade Federal do Rio Grande do Sul,
Cx.P. 15.001, CEP. 91.501-970, Porto Alegre, RS, Brazil

ABSTRACT

A correlation between plant zones and palynozones within a sequence-


stratigraphic framework for the upper Paleozoic rocks of the Paraná Basin, Brazil,
is attempted for the first time. Our study indicates that (1) the stratigraphic bound-
aries (lithostratigraphic boundaries and sequence boundaries) are not coincident
with most of the biostratigraphic horizons; (2) the boundaries between palynozones
and plant zones are also not coincident with each other; and (3) the boundaries of
palynozones lie near the maximum flooding surfaces through the interval analyzed.
The emerging pattern suggests that plant zones are mostly controlled by depositional
processes and palynozones by climate-driven changes. Therefore, the plant zones
correspond to distinct ecofacies, and are better regarded as ecozones rather than as
biozones. On the other hand, the climatic changes that affected the palynofloras were
related directly to the most significant transgressive events, suggesting a link with eus-
tatic oscillations caused by Early Permian Glacial-Interglacial Phases. Aims of fur-
ther work may include (a) evaluation of taphonomic controls in plant-bearing beds,
(b) better understanding of the relation between the plant-bearing strata and their
equivalent palynologic zones, (c) correlation between palynologic and paleobotanic
data and the sequence-stratigraphic framework already established in other areas,
and (d) improvement of the chronostratigraphic chart of the basin through the dis-
covery of layers suitable for radiometric dating.

INTRODUCTION ning of the twentieth century. Most of these studies were moti-
vated by the presence of economically important coal reserves in
Stratigraphic and paleontologic studies of sedimentary suc- the area. In the last decades, increased knowledge on the stratig-
cessions from the southernmost Brazilian Paraná Basin, compris- raphy and paleontology of the deposits from Rio Grande do Sul
ing deposits in the states of Santa Catarina and Rio Grande do resulted in the establishment of sequence-stratigraphic and bio-
Sul, have been carried out by many researchers since the begin- stratigraphic frameworks applicable to the southernmost part of

*E-mails: Iannuzzi—[email protected]; Souza—[email protected]; Holz—[email protected].

Iannuzzi, R., Souza, P.A., and Holz, M., 2010, Stratigraphic and paleofloristic record of the Lower Permian postglacial succession in the southern Brazilian Paraná
Basin, in López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana: Geological Society of
America Special Paper 468, p. 113–132, doi: 10.1130/2010.2468(05). For permission to copy, contact [email protected]. ©2010 The Geological Society of
America. All rights reserved.

113
114 Iannuzzi et al.

the Paraná Basin (Guerra-Sommer and Cazzulo-Klepzig, 1993; cycle that began with the deposition of the upper Itararé unit and
Holz, 1999; Souza and Marques-Toigo, 2003). However, no has its maximum flooding surface within the Palermo Formation
attempts have been made to integrate sequence-stratigraphic and (e.g., Milani et al., 1994; Holz, 1999).
biostratigraphic data. This paper is intended as a first approach
to integrating the stratigraphic sequences and biozones described EARLY PERMIAN SUCCESSION IN THE SOUTHERN
in the sections from the Rio Grande do Sul State. It is part of a PARANÁ BASIN
major biostratigraphic study, including the revaluation of the bio-
zonations already proposed. The main objective of this study is to The lowermost unit of the Early Permian succession in Rio
understand the stratigraphic significance and external controls on Grande do Sul, the Itararé Group, was formed in pro- and post-
these biozones and to evaluate the associated paleoecologic fac- glacial environments. The Rio Bonito Formation is composed
tors in order to establish a broad biostratigraphic framework for of fluvio-deltaic and estuarine to shallow-marine deposits, and
the Lower Permian deposits of the Paraná Basin in the future. the uppermost Palermo Formation has a shallow-marine (lower
shoreface to offshore) origin. The Rio Bonito Formation is fluvial
GEOLOGIC SETTING and deltaic at the base, turns into estuarine in the middle portion,
and is coastal to shallow marine in the upper part. Details may be
The southern margin of the Paraná Basin, an Ordovician to seen in Holz (2003) and Holz et al. (2006). Detailed (third-order)
Cretaceous intracratonic basin in the South American platform, sequence-stratigraphic analysis of this stratigraphic interval
is located in southernmost Brazil and north/northwestern Uru- within the study area, carried out by our group (e.g., Holz, 1999,
guay (Fig. 1A). The basin covers a surface area ~1,700,000 km2, 2003; Holz et al., 2000, 2006), revealed that important regional
has a NE-SW elongated shape, and is ~1750 km long and base-level changes generated three distinct regional unconformi-
900 km wide. The sedimentary fill of the basin was controlled by ties, which constitute sequence boundaries roughly correspond-
tectonic-eustatic cycles linked to late Paleozoic orogenic events ing to the lithostratigraphic limits (Fig. 2).
caused by subduction and terrain accretion on the southwestern The lowermost limit is SB 1, an unconformity regionally
margin of the Gondwana continent, and ended with the open- marked by the contact between the Early Permian succession
ing of the South Atlantic during the Mesozoic (e.g., Zalán et al., and Neoproterozoic crystalline rocks of the basement (or locally
1990; Milani et al., 1994). The prevalence of eustatic-tectonic metasedimentary rocks of the early Paleozoic Camaquã Basin),
cycles, which controlled sedimentation in Paraná Basin, has gen- enclosing a hiatus up to 300 Ma. SB 2 is marked by the con-
erated a stratigraphic record that is marked by numerous inter- tact between fluvial and deltaic sandstone belonging to the Rio
ruptions caused by erosion and non-deposition. Milani et al. Bonito Formation and the underlying glacio-marine mudstone
(1994) subdivided basin fill into six second-order depositional of the Itararé Group. This noticeable facies shift is indicative of
sequences, referred to as “Supersequences”: Rio Ivaí (Rio Ivaí an unconformity, marking a relative sea-level drop which is also
Group, Ordovician–Silurian), Paraná (Paraná Group, Devonian), recorded at other locations around the Rio Grande do Sul shield
Gondwana I (Tubarão and Passa Dois Groups, Carboniferous– (e.g., Alves and Ade, 1996; Holz, 1997). SB 3 is marked by an
Permian), Gondwana II (several Triassic formations), Gondwana erosive surface generated by fluvial incisions on underlying estu-
III (São Bento Group, Jurassic-Cretaceous), and Bauru (Bauru arine and shallow-marine mudstone and sandstone (within the
Group, Cretaceous). Figures 1A and 1B show the distribution of upper interval of the Rio Bonito Formation). These unconformi-
these strata within the southernmost Brazilian areas of the Paraná ties define three third-order depositional sequences, which were
Basin (Rio Grande do Sul and Santa Catarina States) as well as mapped throughout the study area on the basis of subsurface data,
the upper Paleozoic basin stratigraphy. using the maximum flooding surface of the second sequence as
The stratigraphic interval discussed in this paper comprises a datum.
Lower Permian rocks (ranging in age from Asselian/Sakmarian
to late Artinskian/early Kungurian), encompassing the following AGE CONSTRAINTS IN THE PERMIAN OF
lithostratigraphic units (in stratigraphic order): Itararé Group, THE PARANÁ BASIN
Rio Bonito Formation, Palermo Formation, and Irati Formation
(Fig. 1B). This succession constitutes the lower portion of the The main biostratigraphic problem of most of Gondwanan
Gondwana I Supersequence of Milani et al. (1994), which forms Carboniferous-Permian deposits, including those in the Paraná
the thickest sedimentary sequence of the basin (up to 2300 m). Basin, is the absence of biostratigraphically significant marine
The lowermost section of this second-order sequence occurs only faunal elements (such as foraminifers or ammonoids), which pre-
in the depocenter of the basin (from north Santa Catarina to the vent correlations with the Late Carboniferous and Early Permian
Mato Grosso State), assigned to the Late Carboniferous and cor- international stages. In addition, radiometric data are still scarce
responding to the basal and middle portions of the Itararé Group and conflictive in the Permian sequences of Gondwana, militat-
(see Petri and Souza, 1993, and Souza, 2006). From the Early ing against accurate age calibration of the available palynostrati-
Permian onwards, glacially related strata onlap basin margins, graphic schemes. Thus, correlation with international stratigraphic
as in Rio Grande do Sul, recording a second-order transgressive stages can be considered as difficult and speculative for the most
A

SAN
BRAZIL TA C
ATA
R IN A
Paraná
Basin
FLORIANÓPOLIS

SUL
DO
DE
AN
GR
O
RI
3
PORTO ALEGRE

2 1
Geologic Key

on
go
Cenozoic coverage

La
s
Mesozoic rocks (Paraná Basin)

to
Pa
Paleozoic rocks (Paraná Basin)
Paleoproterozoic basement

N
Studied locations:
1 - Morro do Papaleo outcrop
2 - Quiteria outcrop
3 - Borehole CA-53
0 50 100km
STATE CAPITAL

B Rio do Rasto Formation:


deltaic, lagoonal
fluvial, eolian
S-SE N-NW
Teresina
Passa Dois

Formation:
shoreface Corumbataí V
Group

Formation:
---
GONDWANA I SUPERSEQUENCE

Serra Alta tidal plain


---- V
-------
(Upper Carboniferous - Permian)

-- Formation:

-. -. -. -.-. -. -. .... ..
shallow marine

. . . . .. .. ..
V
Irati Formation:
marine restricted V

. .. ........ V .......................................... .
-.
-. Palermo Formation:
.. .......................................... . .
Tubarão Group

.....
shallow marine
Dourados Formation: .. ..... .. V .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. . . . .
continental
.. .. .. .. V .................. . . .
Rio Bonito Formation:
. .... V ..............
shore face, deltaic, lagoonal,
fluvial ................ . . .. .
.. .. V ................. . .
Aquidauana Formation: .................
glacio-continental ..................
... ................
...
Itararé Subgbroup:
glacio-marine to
Serra Geral Formation: .............
glacio-continental
magmatic (Jurassic/Cretaceous) ..........
Figure 1. (A) Location and distribution of the Paraná Basin in southern Brazil (small map) and the main stratigraphic
units of the Rio Grande do Sul State (large map) showing the studied outcrops. (B) Chronostratigraphic chart for the
upper Paleozoic succession of the Paraná Basin with an overview on spatial relations of the deposits throughout the basin
and the main depositional environments; the rectangular box refers to the section analyzed in this paper (modified from
Souza, 2006).
116 Iannuzzi et al.

Chrono- Lithostratigraphy 3rd Order Stratigraphic Idealized Stratigraphic Profile LEGEND


Stratigraphy Framework Sta. Catarina Rio Grande do Sul
Interbedded sand-
HST3 stones and coal
seams (deltaic,
Artinskian

MFS-3 lagoonal systems)


Palermo

SEQ-3
Fm. Shallow-marine
TST3 shales and
mudstones
PERMIAN

Coastal (delta,
TS-3 shoreface, barrier,
Siderópolis
LST3 SB-3 lagoon, tidal flat)
Mb. sandstones and
HST2 MFS-2
Paraguassu Mb.

shales
Rio Bonito Fm.

Fluvial sandstones,
SEQ-2

conglomerates and
Sakmarian

TST2 silts

TS-2 Glacio-continental
to glacio-marine
Triunfo Mb. LST2 SB-2 sedimentites
Itararé Gr. SEQ-1 SB-1
Pre-Cambrian
Interval not studied basement

Figure 2. Sequence-stratigraphic framework and lithostratigraphic equivalents of the studied interval. Lithostratigraphy is based on Schneider
et al. (1974), third-order sequence stratigraphy is from Holz et al. (2000), and ages are based on the palynomorph biostratigraphy provided by
Souza and Marques-Toigo (2003, 2005) and the chronostratigraphic scheme established after Césari (2007), Iannuzzi (2007), Stephenson (2008),
and Guerra-Sommer et al. (2008a, 2008b). Abbreviations: SEQ—third-order sequences, SB—sequence boundary, TS—transgressive surface,
MFS—maximum flooding surface, LST—lowstand systems tract, TST—transgressive systems tract, HST—highstand systems tract.

part. For all these reasons, an additional stratigraphic control based liz Zone (according to Souza and Marques-Toigo, 2005). In the
on extra-basinal correlations with Namibian (Aranos Basin) and context of the time scale of Gradstein et al. (2004), this age is
South African (Main Karoo Basin) sequences is proposed herein assigned to the Roadian Stage. However, this same coal seam and
to support a more consistent chronostratigraphic framework for the other one above were recently redated by Guerra-Sommer
the Lower Permian strata of the Paraná Basin. et al. (2008b), giving new ages that varied from 288.4 ± 2.6 to
In the Itararé Group of Rio Grande do Sul, the lowermost 293 ± 3.5 Ma for the four interbedded tonsteins analyzed. These
strata are positioned slightly above the stratigraphic levels con- ages correspond to earliest to mid-Sakmarian interval, according
taining marine shells related to “Eurydesma Fauna,” which are to the above-mentioned time scale.
typically recorded toward the central parts of the basin (Rocha- In addition, Guerra-Sommer et al. (2008a, 2008b) furnished
Campos and Rösler, 1978). In the Main Karoo Basin of South new radiometric zircon ages of 285.4 ± 8.6 Ma and 288.76
Africa, deposits associated with the “Eurydesma Fauna” encom- ± 1.4 Ma geochronologic data from a tonstein interbedded with
pass a time span of ~10 m.y, based on under- and overlying tuffs other coal seams of the Faxinal coalfield, also situated in the mid-
dated through radiometric methods by Bangert et al. (1999). dle Rio Bonito Formation from Rio Grande do Sul. These dates
Hence, the “Eurydesma Fauna” ranges from the earliest Asselian correspond to the Hamiapollenites karroensis Subzone of Vitta-
(ages of 302 ± 3 Ma and 299.2 ± 3.2 Ma from Bangert et al., tina costabiliz Zone (VcZ). According to the time scale of Grad-
1999) up to the middle Sakmarian stage (288 ± 3 Ma and 289.6 stein et al. (2004), this dating corresponds to the earliest mid- to
± 3.8 Ma according to Bangert et al., 1999). This fauna has been latest Sakmarian interval.
used as an important stratigraphic marker throughout Gondwa- These ages have been partially contested by some authors
nan sections, being associated with a widespread transgressive (Césari, 2007; Iannuzzi, 2007; Stephenson, 2008) who disagree
event referred to as the “Eurydesma transgression” by Dickins with the dating postulated by the biostratigraphic framework,
(1984). Thus, this independent calibration suggests that the strati- which is based mostly on palynomorphs but also on plants,
graphic interval studied begins somewhere in the early to mid- marine invertebrates, and terrestrial vertebrates used in the cor-
Sakmarian (Fig. 2). relative sections from the other Gondwanan deposits. We have
Recently, de Matos et al. (2001) obtained a date of 267.1 therefore discarded the Roadian age obtained by de Matos et al.
± 3.4 Ma (U/Pb) from a tonstein interbedded in the lower coal (2001) and the oldest ages proposed by Guerra-Sommer et al.
seam of Candiota Coal, positioned in the middle Rio Bonito (2008a, 2008b).
Formation, from Rio Grande do Sul, which corresponds to the On the other hand, Guerra-Sommer et al. (2008b) established
Protohaploxypinus goraiensis Subzone of the Vittatina costabi- a mean average age of 290.6 ± 1.5 Ma for tonsteins of both the
Southern Brazilian Paraná Basin 117

Candiota and Faxinal coalfields, indicating an age for this coal Vittatina costabilis Interval Zone
succession constrained to middle Sakmarian. This age interval is Bilaterally and radially symmetrical monosaccate pollen
reasonable and matches up approximately with the ages previously grains are the most common palynomorphs of this palynozone,
established by fossil content, and we accept it herein as the middle such as Cannanoropollis, Plicatipollenites, Caheniasaccites, and
portion of the Rio Bonito Formation (Fig. 2). In conclusion, the Potonieisporites, constituting up to 50%–60% of assemblages, as
deposits of the Rio Bonito Formation start in the earliest middle well as cingulizonate spores (e.g., Lundbladispora, Cristatispo-
Sakmarian and extend up to the latest Sakmarian or even the earli- rites). Disaccate pollen grains became dominant only within the
est Artinskian, considering the age obtained for the overlying Irati upper portion of this palynozone, considered to be correlated with
Formation and the relatively precise dating of this unit. the Hamiapollenites karroensis Subzone. Common taxa are Limi-
Finally, the Irati Formation, the uppermost unit studied herein, tisporites, Vittatina, Scheuringipollenites, and Protohaploxypinus.
was considered as Late Artinskian in age (according to Gradstein Spores are common and only locally dominant in the coal beds
et al.’s time scale), based on the dating of 278.4 ± 2.2 Ma obtained representing local parautochthonous/autochthonous flora (Punc-
from tuff beds through the SHRIMP zircon method by Santos et al. tatisporites, Horriditriletes, Lundbladispora, and Cristatispo-
(2006). Presence of the “Mesosaurus Fauna” and of equivalent rites), where they represent up to 80% of the assemblages.
lithofacies indicates that this unit is correlated with the Whitehill The first appearance of polyplicate species of the genus Vit-
Formation. In Namibia (Aranos Basin) and South Africa (Main tatina (V. saccata, V. subsaccata, V. costabiliz, V. vittifera), as
Karoo Basin), the age of the Whitehill Formation is given by a well as species of Protohaploxypinus (P. goraiensis, P. limpidus),
U/Pb age of 270 ± 1 Ma (= Kungurian-Roadian boundary accord- Fusacolpites fusus and Illinites unicus, marks the lower limit of
ing to the time scale of Gradstein et al., 2004) as determined from this zone, recognized within the upper Itararé Group. The upper
tuffs in the Collingham Formation of the Main Karoo Basin in limit is marked by the appearance of diagnostic species of the
South Africa (Turner in Stollhofen et al., 2000). This unit over- Lueckisporites virkkiae Interval Zone, within the uppermost Rio
lies the Whitehill Formation, indicating a probable Artinskian/ Bonito Formation and/or the lowermost Palermo Formation.
Kungurian age for the Whitehill deposits. This dating agrees with The Vittatina costabiliz Zone is divided into two units, the
the late Artinskian age obtained from tuffs in the Irati Formation Protohaploxypinus goraiensis and Hamiapollenites karroensis
of the Paraná Basin (Santos et al., 2006). Consequently, the top of Subzones (in ascending stratigraphic order). The former is char-
the stratigraphic interval under consideration is neither older than acterized by the range of Protohaploxypinus goraiensis, Illinites
Late Artinskian nor younger than Kungurian (Fig. 2). unicus, and Protohaploxypinus limpidus. The latter is defined by
the range of the eponymous species; its base corresponds to the
PALEONTOLOGIC SETTING first appearance of Striatopodocarpites fusus and Staurosaccites
cordubensis. This interval zone correlates with the H3-J intervals
The most complete palynologic and floral zonations for the of Daemon and Quadros (1970) and with the Cannanoropollis
upper Paleozoic strata of the Paraná Basin have been proposed korbaensis Zone of Marques-Toigo (1991).
by Daemon and Quadros (1970) and Rösler (1978), respectively.
Although these schemes are based on extensive sampling, they Lueckisporites virkkiae Interval Zone
are informal and lack detail. Accordingly, these zonations were Taeniate and polyplicate pollen grains, mainly those of the
abandoned in the last decades in favor of more detailed schemes. genus Protohaploxypinus, Striatopodocarpites, Striatoabieites,
The following formal palynozones and plant zones proposed are Lunatisporites, Lueckisporites, Vittatina, Weylandites, and Mar-
restricted to the southernmost Paraná Basin. supipollenites, are dominant in this zone. Common monosaccate
and disaccate pollen grains (Plicatipollenites, Potonieisporites,
Palynostratigraphy Limitisporites) of the underlying zone are less common. In gen-
eral, spores are scarce; among them monolete species are more
Marques-Toigo (1991) formally proposed the first palyno- common (Thymospora, Laeviagatosporites). Two new species
stratigraphic scheme for the upper Paleozoic in the southern por- appear within this zone (Thymospora criciumensis and Convo-
tion of the Paraná Basin. Further refinements were published by lutispora pintoi). Lueckisporites virkkiae seems to have a strati-
Souza and Marques-Toigo (2003, 2005). graphically consistent first appearance in the Paraná Basin, at the
Permian palynozones of the southernmost Paraná Basin base of the K interval of Daemon and Quadros (1970), which is
were established by Souza and Marques-Toigo (2005) and corre- considered a datum throughout the Paraná Basin.
spond to interval biozones, in accordance with the formal criteria The last appearance of Hamiapollenites karroensis and
of the International Subcommission on Stratigraphic Classifica- Stellapollenites talchirensis, and the first appearance of Lueck-
tion (ISSC) of the International Union of Geological Sciences isporites virkkiae, L. densus, L. stenotaeniatus, Pakhapites fasci-
(IUGS) (Murphy and Salvador, 1999). Selected taxa of these olatus, Weylandites lucifer, Protohaploxypinus hartii, P. sewardi,
palynozones, including common and diagnostic species previ- P. microcorpus, Lunatisporites variesectus, Alisporites nuthal-
ously illustrated by Iannuzzi and Souza (2005), are shown in Fig- lensis, and Striatopodocarpites pantii characterize the lower
ure 3. limit of this zone in the uppermost Rio Bonito Formation and the
A B C

D E F

G H I

J K L

Figure 3. Selected pollen taxa of the Vittatina costabiliz Interval Zone (based on Souza and Marques-Toigo, 2005, and
Iannuzzi and Souza, 2005). (A) Protohaploxypinus goraiensis (Potonié and Lele) Hart 1964 (slide MP-P: 324, England
Finder coordinate: R35/1). (B) Illinites unicus Kosanke emend. Jansonius and Hills 1976 (MP-P: 4033, U50). (C) Vit-
tatina costabiliz Wilson 1962 (MP-P: 3573, G36/2). (D) Striatopodocarpites fusus (Balme and Hennelly) Potonié 1958
(MP-P: 312, U27/2). (E) Hamiapollenites karroensis (Hart) Hart 1964 (MP-P: 1534, Q35/2). (F) Fusacolpites fusus Bose
and Kar 1966 (MP-P: 333/287, P44). Selected pollen taxa of the Lueckisporites virkkiae Interval Zone (based on Souza
and Marques-Toigo, 2005, and Iannuzzi and Souza, 2005). (G) Thymospora criciumensis Quadros, Marques-Toigo and
Cazzulo-Klepzig 1995 (MP-P: 1447, J24). (H) Lueckisporites densus Cauduro 1970 (MP-P: 21, N35-4). (I) Vittatina
subsaccata Samoilovich 1963 (MP-P: 2541, L40/2). (J) Marsupipollenites striatus (Balme and Hennelly) Foster 1975
(MP-P: 4033, Q45). (K) Lunatisporites variesectus Archangelsky and Gamerro 1979 (MP-P: 2344, M49). (L) Stauro-
saccites cordubensis Archangelsky and Gamerro 1979 (MP-P: 2345, M32/2). Slides are housed at the Departamento de
Paleontologia e Estratigrafia/UFRGS (under codes “MP-P”). Scale bar—20 µm.
Southern Brazilian Paraná Basin 119

lowermost Palermo Formation. This interval zone is correlated nate whereas Glossopteris-like leaves are generally rare and
with the K+L intervals of Daemon and Quadros (1970). represented by few species (Glossopteris indica, G. commu-
nis). The protoglossopterid elements (e.g., Rubidgea lanceolata,
Reference Wells R. obovata), initially described by Guerra-Sommer and Cazzulo-
The stratigraphic distribution proposed by Marques-Toigo Klepzig (1993) for this zone, have been discarded by Tybusch
(1991) and Souza and Marques-Toigo (2005) was mainly based and Iannuzzi (2008). The latter authors considered all Rubidgea-
on coal mine wells, as follows: 2-TG-69-RS well (drilled in the like leaves as different types of Gangamopteris ones. Sphenopsid
locality of Santa Terezinha Coal), 5-CA-03-RS, 5-CA-41-RS stems (Paracalamites-like) and leaf branches, glossopterid foli-
(Charqueadas Coal), 5-CA-91-RS (Gravataí-Morungava Coal), age and platispermic seeds are abundant, whereas pteridophylls
2-TG-88-RS (Chico Lomã Coal), P7 (Iruí Coal), and N3 (Santa (Botrychiopsis plantiana) and cordaitalean foliage (Cordaites
Rita Coal). Geologic and palynologic data from these wells were hislopi = C. spathulata) are common, and ginkgoalean leaves
published by Marques-Toigo and Pons (1974), Bortoluzzi et al. (Chiropteris sp.), conifer-like leaf shoots, and glossopterid fruc-
(1980), Marques-Toigo et al. (1982, 1984), and Picarelli et al. tifications are relatively rare.
(1987). The lower limit of this zone is defined by the first appearance
of glossopterid elements in the upper Itararé Group, whereas the
Plant Biostratigraphy upper limit is mostly marked by the appearance of diagnostic
species of the overlying Glossopteris/Rhodeopteridium Zone
Guerra-Sommer and Cazzulo-Klepzig (1993) proposed a in the middle-upper Rio Bonito Formation (Fig. 6). This plant
preliminary plant stratigraphic zonation containing two assem- zone was originally divided into two units, the Gangamopteris
blage zones geographically restricted to Rio Grande do Sul. A obovata and Phyllotheca indica Subzones. However, the spe-
new plant stratigraphic scheme was recently proposed by Jasper cies Phyllotheca indica, eponym of the last subzone, has been
et al. (2003), apparently as a replacement for the above-mentioned regarded by some authors as a synonym of Phyllotheca australis
scheme. However, this scheme was not established in accordance (in Rohn and Lages, 2000). In fact, considering the sphenopsid
to the criteria and recommendations indicated by the ISSC (as material used by Guerra-Sommer and Cazzulo-Klepzig (1993)
summarized by Murphy and Salvador, 1999), because it lacks to erect the P. indica Subzone and the information provided by
(a) information on co-occurring diagnostic taxa, (b) stratotypes, these authors, it seems reasonable to conclude that Brazilian
(c) a reference section, and (d) stratigraphic ranges of taxa. In specimens attributed to P. indica must be included in P. austra-
addition, no correlation between previous and recent schemes lis. Consequently, the denomination P. indica Subzone is herein
was established. Therefore, this scheme is discarded herein. formally replaced by the P. australis Subzone. The Gangamo-
The two plant zones correspond to assemblage biozones pteris obovata Subzone seems to be defined only by the ranges of
as established by Guerra-Sommer and Cazzulo-Klepzig (1993) Cornucarpus patagonius and Cordaicarpus truncata (Iannuzzi
based on a few outcrops. All these localities were referred as et al., 2003a, 2003b, 2006; Jasper et al., 2003; Souza and Ian-
“stratotypes” to these zones by Guerra-Sommer and Cazzulo- nuzzi, 2007, 2009; Tybusch and Iannuzzi, 2008). Therefore, this
Klepzig (1993, p. 64 and 66). However, they are better under- subzone is mainly defined by the absence of the diagnostic spe-
stood as type localities containing type sections. In reality, no cies of the overlying Phyllotheca australis Subzone (Fig. 6). The
specific stratotype has been proposed for each zone/subzone. Phyllotheca australis Subzone is a more consistent unit, defined
Apart from this, these plant zones are in accordance with the cri- by the first appearance of the eponymous species as well as by
teria of the International Subcommission on Stratigraphic Clas- the following taxa: Stephanophyllites cf. S. sanpaulensis, Cheiro-
sification of IUGS, as generally applied to paleobotanic units phyllum speculare, Kawizophyllum sp., Glossopteris occidenta-
(Cleal, 1999). Selected taxa are illustrated in Figures 4 and 5, lis, Samaropsis kurtzii, and S. gigas (Fig. 6).
including main diagnostic taxa and accessory elements. Recently, The Gangamopteris obovata Subzone was originally recog-
Iannuzzi et al. (2003a, 2003b, 2006), Jasper et al. (2003, 2005), nized in the upper Itararé Group, and the Phyllotheca australis
Tybusch (2005), Souza and Iannuzzi (2007, 2009), Vieira et al. Subzone has been placed in the lowermost Rio Bonito Forma-
(2007), and Tybusch and Iannuzzi (2008) have included new key tion (Guerra-Sommer and Cazzulo-Klepzig, 1993). However,
taxa, and extended the range of some taxa previously known in Iannuzzi et al. (2003a, 2003b, 2006) suggested that the range of
these zones. The present scheme is outlined below and shown in the Phyllotheca indica Subzone be restricted to the top of the
Figure 6. Itararé Group.

Botrychiopsis plantiana Assemblage Zone Glossopteris/Rhodeopteridium Assemblage Zone


This zone is characterized by the first appearance of glosso- In this zone, the genus Glossopteris becomes dominant with
pterid elements and by the local abundance of the pteridophyll the emergence of new species (G. mosesii, G. browniana), dis-
Botrychiopsis plantiana and/or Phyllotheca-like sphenophytes. placing Gangamopteris as the most common glossopterid ele-
Among the glossopterid plants, the gangamopterid elements ment. In addition, lycophytes (Brasilodendron pedroanum) and
(Gangamopteris buriadica, G. angustifolia, G. obovata) domi- true ferns (Pecopteris sp., Asterotheca sp., Sphenopteris sp.,
A B C

D E F G

Figure 4. Selected plant taxa of the Botrychiopsis plantiana Assemblage Zone. (A) Gangamopteris buriadica. (B) Ganga-
mopteris obovata. (C) Cordaites sp. (D) Phyllotheca indica (= P. australis). (E) Cheirophyllum speculare. (F) Botrychi-
opsis plantiana. (G) Stephanophyllites cf. S. sanpaulensis. Specimens housed at the Departamento de Paleontologia e
Estratigrafia, Universidade Federal do Rio Grande do Sul.
A B C

D E

F G
H

Figure 5. Selected plant taxa of the Glossopteris/Rhodeopteridium Assemblage Zone. (A) Glossopteris occidentalis.
(B) Botrychiopsis valida. (C) Asterotheca sp. (D) Coricladus quiterensis. (E) Rhodeopteridium sp. (F) Lycopodites sp.
(G) Leaves of Brasilodendron sp. (H) Stem of Brasilodendon pedroanum. Specimens housed at the Departamento de
Paleontologia e Estratigrafia, Universidade Federal do Rio Grande do Sul.
122 Iannuzzi et al.

GEOCHRONOLOGY SAKMARIAN /ARTINSKIAN


GUATÁ GROUP
LITHOSTRATIGRAPHY ITARARÉ GROUP
RIO BONITO FM.

BOTRYCHIOPSIS PLANTIANA ZONE


GLOSSOPTERIS /
BIOSTRATIGRAPHY RHODEOPTERIDIUM
GANGAMOPTERIS PHYLLOTHECA ZONE
OBOVATA INDICA
SUBZONE SUBZONE

Cornucarpus patagonicus
Cordaicarpus truncata *
Gangamopteris obovata
Botrychiopsis plantiana §
Chiropteris spp. *
Gangamopteris angustifolia §
Gangamopteris buriadica §
Glossopteris indica § Figure 6. Updated biostratigraphic
Glossopteris communis distribution of plant taxa within the
Cordaites hislopi assemblage biozones of Guerra-Sommer
Buriadia isophylla § and Cazzulo-Klepzig (1993) based on
Samaropsis seixasi new data sets published by Iannuzzi
Cordaicarpus aff. C. famatinensis * et al. (2003a, 2003b, 2006), Jasper et al.
Stephanophyllites spp. * (2003, 2005), Souza and Iannuzzi (2007,
2009), Vieira et al. (2007), and Tybusch
Cheirophyllum speculare *
and Iannuzzi (2008). Taxa marked with
Dicranophyllum sp. * “*” represent new key species added
Samaropsis kurtzii * recently. Taxa marked with “§” represent
Samaropsis cf. S. rigbyi * previous taxa whose stratigraphic range
Samaropsis gigas * has been modified based on the new data
Cordaicarpus cerronegrensis * set. Ages of biozones are based on the
Cordaicarpus aff. C. brasilianus * palynostratigraphy provided by Souza
Phyllotheca australis and Marques-Toigo (2003, 2005) and
Kawizophyllum sp. * the chronostratigraphic scheme estab-
Glossopteris occidentalis § lished after Césari (2007), Iannuzzi
Scutum sp. (2007), Stephenson (2008), and Guerra-
Rhodeopteridium sp. Sommer et al. (2008a, 2008b).
Sphenopteris spp. *
Pecopteris pedrasica *
Neomariopteris sp. *
Asterotheca sp. *
Brasilodendron pedroanum
Cyclodendron sp. *
Ginkgophytopsis sp. *
Botrychiopsis valida *
Glossopteris mosesii
Glossopteris browniana
G. obovata var. major *
Plumsteadia sennes
Arberia minasica
Otokaria sp.
Coricladus quiteriensis *
Samaropsis aff. S. millaniana *

Neomariopteris sp.) appear for the first time in Rio Grande do to the typical “Glossopteris Flora” and typically occur in associa-
Sul. The appearance of a new conifer-like plant (Coricladus quit- tion with coal measures.
eriensis) and Gangamopteris leaves (G. obovata var. major) is The lower limit of this zone is apparently defined by the last
another significant feature. Consequently, the plant assemblages record of Stephanophyllites cf. S. sanpaulensis and Cheirophyl-
become highly diversified. Assemblages of this zone correspond lum speculare, and by the first appearance of Brasilodendron
Southern Brazilian Paraná Basin 123

pedroanum, Pecopteris spp., Asterotheca sp., Neomariopteris sp., occurs from the middle interval of Sequence 3 to at least the
Sphenopteris spp., Botrychiopsis valida, Glossopteris browni- top of Sequence 4 in the Irati Formation (Fig. 9). The boundary
ana, G. mosesii, Arberia minasica, Ottokaria sp., Plumsteadia between the Protohaploxypinus goraiensis and the Hamiapollen-
sennes, and Coricladus quiteriensis within the middle-upper Rio ites karroensis Subzones lies near the maximum flooding surface
Bonito Formation (Fig. 6). Its upper limit is marked by the dis- MFS 2, but its exact position cannot be traced with the avail-
appearance of the most diagnostic species in the overlying Pa- able data.
lermo Formation. The stratigraphic position of plant zones is indicated by
According to available data, the B. valida Subzone of Jas- correlations between the Morro do Papaléo and Quitéria out-
per et al. (2003) and the Glossopteris/Rhodeopteridium Zone of crops and the CA-53 borehole (Fig. 7) (Holz, 1999). This cor-
Guerra-Sommer and Cazzulo-Klepzig (1993) are considered relation was established on the basis of outcrop data. The Morro
equivalent units for stratigraphic purposes. Thus, the new taxa do Papaléo section is the most complete, containing the three
introduced by Jasper et al. (2003) as belonging to the B. vali- plant zones (Fig. 8). The Botrychiopsis plantiana Zone extends
da Subzone were transferred to and listed in the Glossopteris/ throughout Sequence 1, being its upper limit coincident with SB
Rhodeopteridium Zone for the present analysis (Fig. 6). 2, which corresponds to the Itararé/Rio Bonito boundary (Fig. 9).
The precise position of the boundary between the Gangamopteris
Reference Sections obovata and Phyllotheca australis Subzones remains unknown.
The reference sections for the Botrychiopsis plantiana Zone The Glossopteris/Rhodeopteridium Zone passes through SB 3,
are in the localities of Morro do Papaléo (including both Ganga- extending approximately from SB 2 to TS 3 in the upper Rio
mopteris obovata and Phyllotheca australis Subzones) and Qui- Bonito Formation (Fig. 9).
téria (the upper subzone only) (Guerra-Sommer and Cazzulo- With respect to correlations between the plant zones and
Klepzig, 1993). Additional localities were listed for the lower palynozones, the Botrychiopsis plantiana Zone is equivalent to
subzone, G. obovata, including Faxinal, Goulart Farm, Acampa- the Protohaploxypinus goraiensis Subzone of the Vittatina costa-
mento Velho, and Cambaí Grande. The reference sections for the biliz Interval Zone (Fig. 8). The Glossopteris/Rhodeopteridium
Glossopteris/Rhodeopteridium Zone occur in Quitéria and Faxi- Zone corresponds mostly to the Hamiapollenites karroensis Sub-
nal Mine (Guerra-Sommer and Cazzulo-Klepzig, 1993). In addi- zone, being normally recorded in the upper Rio Bonito Forma-
tion, the uppermost part of the section from Morro do Papaléo tion. Nevertheless, the presence of lycopsid stems (assigned with
is regarded as a stratotype for the Glossopteris/Rhodeopteridium doubts to the genus Cyclodendron) in coal seams of the Morro
Zone (Iannuzzi et al., 2003a, 2003b). Papaléo (Fig. 8), and of Brasilodendron pedroanum in coal seams
Consequently, we chose herein the localities of Morro do of the Candiota Mine (Chaloner et al., 1979), indicates that the
Papaléo and Quitéria for a detailed lithologic logging of plant- Glossopteris/Rhodeopteridium Zone extends well into the Proto-
bearing exposures and stratigraphic studies, due to the presence haploxypinus goraiensis Subzone (Fig. 9).
of more than one of the biostratigraphic units erected by Guerra- In contrast to the original proposal of Guerra-Sommer and
Sommer and Cazzulo-Klepzig (1993) (see Piccoli et al., 1991, Cazzulo-Klepzig (1993), the presence of the Phyllotheca aus-
and Iannuzzi et al., 2003a, 2003b, 2006). In Morro do Papaléo, tralis Subzone in the Quitéria outcrop has not been confirmed
all three successive biostratigraphic units previously established herein. Despite the occurrence of some diagnostic taxa related
have been recorded, whereas in Quitéria at least the two upper- to this subzone in the lowermost strata of the Quitéria expo-
most units (Phyllotheca australis Subzone and Glossopteris/ sure, such as Botrychiopsis plantiana and Chiropteris sp., the
Rhodeopteridium Zone) are present. palynologic content indicates a higher stratigraphic position for
this outcrop. Among the palynomophs recovered, an important
DISCUSSION index element of the Hamiapollenites karroensis Subzone is
Striatopodocarpites fusus (Jasper et al., 2006). This indicates
Integration of Biostratigraphic and Sequence- that the Quitéria exposure corresponds to the uppermost Vitta-
Stratigraphic Information tina costabiliz Interval Zone, being stratigraphically close to the
Rio Bonito/Palermo boundary (Figs. 8 and 9). The presence of
Figures 7, 8, and 9 show the results obtained from the inte- typical elements of the P. australis Subzone in Quitéria outcrop
gration of stratigraphic and paleontologic data. Palynozones are therefore should be interpreted as a result of a facies-controlled
more precisely correlated with the stratigraphic scheme of Holz reappearance of those taxa at a higher stratigraphic position.
(1999) than plant zones because palynozones are based on well Consequently, the Glossopteris/Rhodeopteridium Zone is con-
samples. The Vittatina costabiliz Interval Zone occurs from the sidered the only biozone present in the Quitéria outcrop (Fig. 8),
base of Sequence 1, extending through the upper Itararé Group and the above-mentioned taxa correspond to co-occurring ele-
and most of the Rio Bonito Formation. Its upper limit lies near ments present in the lower levels, within the underlying B. plan-
the maximum flooding surface of Sequence 3, encompassing the tiana Zone.
uppermost Rio Bonito Formation and the lowermost Palermo Another important aspect that emerges from this analysis is
Formation (Fig. 9). The Lueckisporites virkkiae Interval Zone the fact that the lithostratigraphic boundaries are not coincident
Towards Platform - Shoreline Towards Inner Valley - Continent

Distal borehole Approx. 30 km QUITÉRIA OUTCROP Approx. 20 km MORRO DO


(CA - 53) PAPALÉO OUTCROP

PALYNOZONES
FLUVIO-DELTAIC
SB 3

FLUVIAL
H. karro-
ensis Sub.

-C- C- C-

-C- C- C-
-C- C- C-

-C- C- C-
HST 2
?
FAN DELTAIC

LAGOONAL
MFS 2
FLUVIAL
Vittatina costabilis Zone

P. goraiensis Sub.

FLUVIO-DELTAIC

SEQUENCE STRATIGRAPHY
2
TST 2 SB

LAGOONAL

SB 1

FLUVIO-DELTAIC
LST 2

GLACIAL CONTINENTAL -
GLACIAL MARINE

Figure 7. Stratigraphic correlation between outcrops analyzed herein and borehole CA-53. Palynologic zonation (right side) was very useful for determining the stratigraphic posi-
tion of the Quitéria section. Transect made along a paleovalley (“Mariana Pimentel Paleovalley”), being the distal facies represented by the well and the proximal ones by outcrops.
Sequence stratigraphy shown at left side from Holz et al. (2000), palynostratigraphy at right side according to Souza and Marques-Toigo (2003, 2005). Stratigraphic abbreviations,
see Figure 2; palynozones, see the text.
No record
SB 3

r-t
Morro do Papaléo

Glossopteris /
1m
n, o p, q
-C- C- C-

V. costabilis I. Zone
-C- C- C-

H. karroensis Subzone
-C- C- C-

-C- C- C-

Rhodeopteridium A. Zone
0m a, d b n, v - y

Quitéria
r, u
No record

m
Figure 8. Stratigraphic range of plant fossils: Quitéria and
Morro do Papaléo (based on Piccoli et al., 1991; Iannuzzi
et al., 2003a, 2003b, 2006; Jasper et al., 2003, 2005). Palyno-
zones and plant zones are plotted side by side for each outcrop.
Glossopteris / Rhodepteridium A. Zone

Taxa: a—Botrychiopsis plantiana; b—Gangamopteris angus-


tifolia; c—Gangamopteris obovata; d—Chiropteris sp.; e— SB 2
Phyllotheca indica (= P. australis); f—Stephanophyllites san-
paulensis; g—Cheirophyllum speculare; h—?Dicranophyllum c-i j-l
sp.; i—Gangamopteris buriadica; j—Kawizophyllum sp.;
k—Glossopteris indica; l—Glossopteris occidentalis; m—
Cyclodendron sp.; n—Gondwanostachis sp.; o—Arberia sp.;
p—Glossopteris browniana; q—Rhodeopteridium sp.; r—
Brasilodendron cf. pedroanum; s—Coricladus quiteriensis;
t—Lycopodites sp.; u—Ginkgophytopsis sp.; v—Asterotheca
sp.; w—Pecopteris sp.; x—Sphenopteris sp.; y—Neomari-
Vittatina costabilis Interval Zone

opteris sp. Abbreviations: SB—erosive surface/sequence


Protohaploxypinus goraiensis Subzone

boundary; B. plantiana A. Zone—Botrychiopsis plantiana


Assemblage Zone; Glossopteris/Rhodeopteridium A. Zone— 5m
B. plantiana A. Zone

Glossopteris/Rhodeopteridium Assemblage Zone; V. costabiliz


I. Zone—Vittatina costabiliz Interval Zone; G. obovata Sub—
Gangamopteris obovata Subzone; P. australis Sub—Phylloth-
G.obovata Sub.. P.australis Sub.

a b
eca australis Subzone. 0m
126 Iannuzzi et al.

Chronostratigraphy Biostratigraphy Stratigraphy Environment


Period Epoch Stages Ages (Ma) Plants Palynomorphs Lithologies Sequences Main Systems Sea level

Kungurian 275.6 + 0.7 ?????????? + -

Sequence 4
Irati
Shallow
No

Marine
Lueckisporites carbonatic
plant Formation marine
virkkiae
278.4 + 2.2 zones Interval Zone
Artinskian
SB 4
P E R M I A N

Sequence 3
Palermo Shallow
Cisuralian

marine
Fm.
MFS 3
Vittatina costabilis Interval Zone H. karrooensis
284.4 + 0.7 Shoreline
Subzone SB 3

Coastal Plain
Rhodeopteridium
Glossopteris /

Assemblage

285.4 + 8.6
Protohaploxypinus goraiensis Estuarine / MFS 2

Sequence 2
Rio Bonito Lagoonal
Zone

Formation
Fluvial /
Delta plain
Subzone

288.4 + 2.6

Shoreline
Sakmarian
SB 2

Sequence 1

glacial glacial
Pro- Post-
B. plantiana Shoreline
Assemblage Z. (upper)
Itararé Prodelta
No record Glacio-
Group marine

Figure 9. Correlation chart of chrono- and biostratigraphic (= based on palynomorphs and plants) data, and paleoenvironmental framework for
the Lower Permian of the Paraná Basin in Rio Grande do Sul State. Note the relations among lithostratigraphic units, stratigraphic sequences,
biozone boundaries, paleoenvironments, and sea-level changes. Differences between palynozones and plant zones seem to be strongly facies
controlled. Chronostratigraphy according to Gradstein et al. (2004). Abbreviations: SB—sequence boundary, MFS—maximum flooding surface.
Legend: Levels with three stars indicate radiometric ages obtained by Guerra-Sommer et al. (2008a, 2008b) from tonsteins of the Rio Bonito
Formation and by Santos et al. (2006) from tuff beds of the Irati Formation. See text for more details.

with most biostratigraphic horizons. The Vittatina costabiliz Correlations of the Biozones
Interval Zone includes the upper Itararé Group and most of the
Rio Bonito Formation (Fig. 9). No significant biostratigraphic Palynozones can apparently be extended throughout the
difference has been recorded within these sections, despite Paraná Basin, taking into account the range of selected spore-
noticeable lithologic changes below and above SB 2, mapped pollen grains and the horizons of their appearance and disap-
regionally throughout the state of Rio Grande do Sul. The bound- pearance. Therefore, a correlation between these units and the
ary between the V. costabiliz and Lueckisporites virkkiae Inter- previous informal proposal of Daemon and Quadros (1970) for
val Zones is recorded in the uppermost Rio Bonito Formation, the whole basin was established by Souza and Marques-Toigo
slightly below the Rio Bonito/Palermo boundary (Fig. 9). The (2003). According to these authors, the V. costabiliz Interval
only coincidence is that of the Botrychiopsis plantiana Zone and Zone correlates with the H3-J intervals of Daemon and Quadros
Sequence 1 (Fig. 9). (1970), and L. virkkiae Interval Zone is correlated with the K and
With respect to the boundaries between palynologic zones L intervals.
and plant stages, the same situation is noted. The transition In terms of extra-basinal correlations, the V. costabiliz Inter-
from the Botrychiopsis plantiana Zone to the Glossopteris/ val Zone is correlated with the Cristatisporites Zone and the
Rhodeopteridium Zone occurs within the V. costabiliz Interval Fusacolpites fusus-Vittatina subsaccata Interval Biozone from
Zone and is not coincident with the Protohaploxypinus goraiensis/ northern Argentina (Césari and Gutiérrez, 2000). Other simi-
Hamiapollenites karroensis boundary (Fig. 9). lar assemblages occur elsewhere in Lower Permian deposits of
Southern Brazilian Paraná Basin 127

Gondwana regions, such as Australia, India, Antarctica, Oman, Iannuzzi and Souza (2005) are locally abundant in the overly-
and Saudi Arabia (e.g., Jones and Truswell, 1992; Stephenson ing Rio Bonito Formation. In Rio Grande do Sul, these assem-
and Filatoff, 2000). Like the Vittatina costabiliz Zone, the L. virk- blages occur in peat-forming and fluvial environments. Iannuzzi
kiae Interval Zone has equivalent units widely known from the and Souza (2005) considered the “Glossopteris–Brasilodendron
Middle to Upper Permian Gondwanan strata (Césari and Gutiér- Flora” as recording the vegetation of various environments in
rez, 2000). coastal lowlands during the postglacial phase, including peat-
In contrast, plant zones are only valid for the Rio Grande forming swamps.
do Sul State, as originally proposed by Guerra-Sommer and Plant-bearing deposits of the Rio Bonito Formation accumu-
Cazzulo-Klepzig (1993). In fact, the other few formal plant lated in widespread coastal plains developed immediately after
zones proposed in the last decades for the Paleozoic deposits of the last maximum sea-level rise caused by the melting of gla-
the Paraná Basin are restricted in terms of both time intervals and ciers in the Paraná Basin (Iannuzzi and Souza, 2005). Ice melt
geographic distribution within the basin (Millan, 1987; Guerra- occurred during deposition of Sequence 1 (Figs. 2 and 9). As a
Sommer and Cazzulo-Klepzig, 1993). For this reason, Iannuzzi result, the higher diversity of the Glossopteris/Rhodeopteridium
and Souza (2005) recently proposed three successive informal Zone or “Glossopteris–Brasilodendron Flora” was previously
floras, representing developmental stages of the whole flora, to associated with a postglacial climatic amelioration (Guerra-
be adopted for the Lower Permian strata of the Paraná Basin. A Sommer and Cazzulo-Klepzig, 1993; Jasper et al., 2003). How-
tentative correlation between these floras and plant zones can be ever, the palynologic record indicates that such amelioration took
attempted on the basis of similarities in taxonomic lists and rela- place much earlier, with the appearance of assemblages of the
tive abundance of floral elements. Vittatina costabiliz Interval Zone near the base of Sequence 1
As with the other Permian floral records in Gondwana, the (Fig. 9). The lower limit of the V. costabiliz Zone is character-
predominance of gangamopterid leaves over glossopterid ones ized by the appearance of different types of disaccate striate pol-
and the abundance of sphenophytes are not unique to the B. plan- len (Vittatina spp. and Protohaploxypinus spp., among others),
tiana Zone but are also typical features shared among the ear- which has classically been considered as indicative of warmer
liest Permian assemblages of the “Phyllotheca–Gangamopteris conditions by palynologists from Gondwana.
Flora” from the Paraná Basin, as defined by Iannuzzi and Souza Despite changes in plants (from G. obovata–P. australis
(2005). Similar assemblages occur elsewhere in the lowermost Subzones to Glossopteris/Rhodeopteridium Zone) and lithology
lower Gondwana strata (= Sakmarian), as for instance in Austra- (from subglacial-proglacial to paralic facies) through the upper-
lia, India, southern Africa, and southern South America (Retal- most Itararé Group and lower-middle Rio Bonito Formation
lack, 1980; Anderson and Anderson, 1985; Archangelsky et al., (Figs. 2 and 9), no significant variation has been recorded within
1996; Maheshwari and Bajpai, 2001). Glossopteris-dominated the V. costabiliz Interval Zone up to the P. goraiensis/H. kar-
assemblages, on the other hand, similar to those of the younger roensis boundary, except for the appearance of facies-controlled
Glossopteris/Rhodeopteridium Zone, are typically in coal-bearing species locally restricted to coal seams (i.e., the Caheniasaccites
strata throughout Gondwana. In the Paraná Basin, they correspond ovatus Ecozone of Souza and Marques-Toigo, 2005). Although
to the “Glossopteris–Brasilodendron Flora” (Iannuzzi and Souza, evidence of proglacial and subglacial environments is recorded at
2005). Closely related Lower Permian (= Sakmarian–Artinskian) the base of Sequence 1 (e.g., dropstones and striated pavements
assemblages occur in Argentina, South Africa, India, and Austra- corresponding to the lowermost Itararé Group in Rio Grande do
lia (Retallack, 1980; Anderson and Anderson, 1985; Archangel- Sul), the spore/pollen-producing plants are essentially the same as
sky et al., 1996; Maheshwari and Bajpai, 2001). An integrative occur in overlying stratigraphic levels. These data are interpreted
correlation chart of chrono-, litho-, and biostratigraphic data for herein as reflecting the absence of significant large-scale climatic
the Lower Permian of the Paraná Basin is shown in Figure 10. changes throughout Sequence 1 and much of Sequence 2 (Fig. 9).
In conclusion, glacial evidence at the base of the interval under
Significance of the Biozones consideration appears to represent local rather than regional ice
influence, perhaps reflecting the last ice tongues surrounded by
General Controls relatively rich floras. The dominance of pollen species with coni-
According to Iannuzzi and Souza (2005), the “Phyllotheca– fer and pteridosperm affinities (e.g., monosaccate pollen) in the
Gangamopteris Flora” records the transition between degla- V. costabiliz Interval Zone suggests a temperate forest coexisting
ciation and postglacial depositional sequences in the uppermost with the locally distributed ice tongues. Consequently, the whole
Itararé Group. This flora was therefore preserved in nearshore interval analyzed herein is understood as having been generated
deposits after the retraction of the ice sheet. Coeval assemblages during a “postglacial climatic phase,” representing the ending
belonging to the Botrychiopsis plantiana Zone occur only in the of the Gondwanan Ice Age, at least in the Paraná Basin. Thus,
uppermost Itararé Group in postglacial deposits representing the rise of diversity in the Glossopteris/Rhodeopteridium Zone
proximal settings within lacustrine to lagoonal environments. seems to have been driven by changes in the depositional settings
The assemblages of the Glossopteris/Rhodeopteridium Zone preserved rather than by climatic changes. The characteristic ele-
that correspond to the “Glossopteris–Brasilodendron Flora” of ments of this zone are interpreted as colonizers of new habitats
128 Iannuzzi et al.

Palynology Paleobotany Stratigraphic


Chrono- Litho- control
stratigraphy stratigraphy
Daemon and Souza and Iannuzzi Guerra-Sommer
Quadros Marques-Toigo and Souza and (fauna and
(1970) (2003, 2005) (2005) Cazzulo-Klepzig radiometric ages)
(1993)

Roadian Serra Alta / Teresina Zones from Rohn & Rösler (2000)
Kungurian / Rio do Rasto L1 270 + 1 Ma
Fms. L2
L Lueckisporites virkkiae
(Cisuralian - Guadalupian)

L3 Interval Zone No Plant Zones Mesosaurus Fauna


Early to Middle Permian

Irati Fm. Polysolenoxylon - Glossopteris Flora


278 + 1 Ma
Palermo Fm. K Sterile Interval
Artinskian
Hamiapollenites
J karroensis Subzone Glossopteris - Brasilodendron
Rio Bonito Glossopteris / 285.4 + 8.6 Ma
Vittatina Flora Rhodeopteridium
Formation I2 - I4 costabilis Zone
Protohaploxypinus 288.4 + 2.6 Ma
I Interval
goraiensis Phyllotheca - Gangamopteris Flora B. plantiana Zone
I1 Zone 289.6 + 3.8 / 288.0+ 3 Ma
Sakmarian Subzone
H3 Eurydesma-type Fauna
Asselian
Sterile Interval
H 302 + 3.0 / 299.2+ 3.2 Ma
Itararé
Late Carboniferous

H1
(Pennsylvanian)

Subgroup H2 Crucisaccites monoletus


Interval Zone Pre-Glossopteris
Westphalian Flora
No Record
Radiometric ages:
Ahrensisporites cristatus
G Interval Zone Brazilian rocks
African rocks

Figure 10. Correlation chart of chrono- and biostratigraphic (= based on palynomorphs and plants) data, and paleoenvironmental framework
for the Lower Permian of the Paraná Basin and Rio Grande do Sul State with correlative radiometric dated horizons from Brazil (Irati Forma-
tion), southern Namibia (Aranos Basin), and South Africa (Main Karoo Basin). Absolute ages according to Stollhofen et al. (2000), Santos et al.
(2006), and Guerra-Sommer et al. (2008a, 2008b). See text for more details. Chronostratigraphy according to Gradstein et al. (2004). Modified
from Iannuzzi and Souza (2005).

associated with paralic environments recorded during deposition during the Artinskian-Kungurian transition (Ziegler et al., 1997).
of Sequence 2. Accordingly, this flooding event could be considered as a datum
Climatically driven changes in palynofloras appear to have throughout Gondwana, permitting an intra-continental correlation
occurred only during the Vittatina costabiliz/Lueckisporites virk- as good as the “Eurydesma Transgression” delineated by Dickins
kiae transition. Most of the new pollen grains that arose from (1984) for the Early Permian (= early Sakmarian). In fact, some
the Lueckisporites virkkiae Interval Zone have been considered authors have also addressed the use of other Carboniferous–
to be derived from xerophyllic plants, based on their probable Permian postglacial transgressions as tentative chronostrati-
botanical affinities and intrinsic morphologies adapted to dry graphic horizons in Gondwana deposits (López-Gamundí, 1989;
conditions (e.g., striate bodies). Furthermore, the co-appearance López-Gamundí et al., 1994; Veevers and Powell, 1987).
of abundant carbonate deposits within these sections indicates a
general rise in average temperatures and a decrease in humidity. Plant Zones
In addition, the analysis of growth rings of logs recovered from From their analysis of assemblages from the Morro do
the Irati Formation indicates a wet-dry seasonal climate of Medi- Papaléo outcrop, Iannuzzi et al. (2003a, 2003b, 2006, 2007) sug-
terranean type (Alves, 1999). This climatic change associated gested that the subzones included in the B. plantiana Zone are
with maximum flooding surfaces present in the Palermo and Irati probably facies controlled. In this context, the lowermost assem-
Formations seems to indicate the establishment of warmer cli- blage in the Morro do Papaléo outcrop that corresponds to the
mates in the Paraná Basin with increasingly dry conditions. This G. obovata Subzone (Fig. 8) consists of transported plant material
event may correspond to a late Artinskian/early Kungurian warm derived from nearby vegetation and preserved in shallow lake or
peak that followed the cold to cool temperate conditions recorded lagoon fills. Up-section in the Morro do Papaléo, the next assem-
in Western Australia (Nicoll and Metcalfe, 1997). Therefore, a blage (which is related to the P. australis Subzone, Fig. 8) has
rise in sea level resulted from deglaciation in Eastern Gondwana been interpreted as riparian or marginal-lake to lagoon vegetation
Southern Brazilian Paraná Basin 129

(Iannuzzi et al., 2003a, 2003b, 2006, 2007). Consequently, the deposits, and an upper autochthonous assemblage that is domi-
assemblages of the B. plantiana Zone were classified by these nated by lycophytes in living position, corresponding to vegeta-
authors as (1) allochthonous, consisting of drift plant remains tion associated with peat-forming environments (i.e., “roofshale
associated with lower G. obovata Subzone, (2) parautochtho- floras”) (Fig. 8).
nous, composed of short-distance transported material, and Despite differentiation of all these assemblages, the tapho-
(3) autochthonous, dominated exclusively by portions of sphe- nomic processes involved in the generation of these plant assem-
nophytes in growth position, with (2) and (3) being assigned to blages have not yet been adequately studied, preventing the
the same horizon and belonging to the upper P. australis Subzone establishment of a strong relation between microhabitats and
(see Fig. 8). Furthermore, no significant sedimentologic break associated floral elements. In addition, we need a better under-
or environmental change was detected in the interval (5–15 m standing of the ecologies of individual species and/or the associ-
thick) that separates the fossiliferous horizons containing typical ated communities.
elements of the G. obovata Subzone from those of the P. austra-
lis Subzone (Fig. 8). Therefore, taking into account the Morro CONCLUDING REMARKS
do Papaléo section, the plant subzones of the B. plantiana Zone
seem to reflect primarily changes in the depositional settings, Until recently a proper understanding of the stratigraphic
apparently corresponding to distinct ecofacies (Iannuzzi et al., significance of established plant zones and palynozones in the
2003a, 2003b). Thus, the G. obovata and P. australis Subzones Paraná Basin was obscured by the lack of an adequate strati-
are better understood as ecozones rather than as biozones. graphic control. Stratigraphic information from Rio Grande do
During recent fieldwork at the Morro do Papaléo local- Sul has accumulated during the last decades, however, result-
ity, additional recovered plant material demonstrates the pres- ing in the establishment of a sequence-stratigraphic framework.
ence of almost the same taxa as in the levels attributed to both Thus, we have been able to attempt, for the first time, a correla-
the G. obovata and P. australis Subzones, except for the occur- tion of the plant zones with palynozones and stratigraphic frame-
rence of B. plantiana, which still remains restricted to the lower work, with the following results: (1) the stratigraphic boundaries
G. obovata Subzone in this section (Fig. 8). However, the range (lithostratigraphic boundaries and sequence boundaries) are not
of B. plantiana is known to extend stratigraphically upward coincident with most of the biostratigraphic horizons; (2) the
from the base of the Quitéria outcrop, which lies within the boundaries between palynozones and plant zones are also not
Glossopteris/Rhodeopteridium Zone (Fig. 8). Therefore, the coincident with each other; and (3) the boundaries of palyno-
only exclusive taxon left in the G. obovata Subzone is Cornu- zones lie near the maximum flooding surfaces through the inter-
carpus patagonicus, a small seed rarely found in the floral record val analyzed.
of Rio Grande do Sul (Fig. 6). On the other hand, among all These results suggest that plant zones are controlled mostly
the fossiliferous localities mentioned in Rio Grande do Sul by by depositional processes and palynozones by climate-driven
Guerra-Sommer and Cazzulo-Klepzig (1993) (Faxinal, Goulart changes. Consequently, the plant zones correspond to distinct
Farm, Acampamento Velho, Cambaí Grande, Faxinal Mine, and ecofacies, and are better regarded as ecozones rather than as bio-
Quitéria), the successive and simultaneous presence of both the zones. On the other hand, the climatic changes that affected the
G. obovata and P. australis Subzones in the same exposure has palynofloras were related directly to the most significant trans-
been recorded only in the Morro do Papaléo section until now. gressive events recorded throughout the Paraná Basin. This sug-
Furthermore, the exact stratigraphic position of the other out- gests that the primary mechanism driving the changes in climatic
crops is not well determined yet. For all these reasons, it is not conditions was linked to eustatic oscillations caused by Early
clear whether the subdivision of B. plantiana Zone could be use- Permian Glacial-Interglacial Phases recorded in Gondwana.
ful for stratigraphic purposes. The persistence of the same palynozone (i.e., V. costabiliz
Given that the record of the Glossopteris/Rhodeopteridium Zone) from the upper Itararé Group to the upper Rio Bonito
Zone reflects distinct plant-bearing depositional settings across Formation, despite the noticeable facies changes from pro- and
the coastal wetlands, at least four types of assemblages can be subglacial to coastal-plain environments, is underscored in this
distinguished according to Jasper and Guerra-Sommer (1999), study. This pattern appears to reveal the absence of significant
Jasper (2004), and Iannuzzi et al. (2003a, 2003b, 2007). Near the large-scale climatic changes during this interval, because the
top of the Morro do Papaléo section, there are two distinct assem- spore/pollen-producing plants present from the bottom to the top
blages: (1) one is parautochthonous and consists of a sorting of can be considered basically the same. In addition, palynologic
distinct parts of lycophytes and glossopterid and cordaitalean data suggest that the Permian glacial record of the Paraná Basin
plants; and (2) another is hypo-autochthonous and dominated by represents in reality a “postglacial phase,” equivalent to an inter-
ferns and glossopterid foliage. Both are preserved in floodplain glacial phase. If this is the case, it is reasonable to conclude that
deposits (Fig. 8). Two additional assemblage types occur in the the peak of the last glacial event (i.e., the glacial phase) did not
lower and middle intervals of the Quitéria outcrop: lower hypo- get preserved in the sedimentary record.
parautochthonous assemblages composed of plant remains trans- Thus, the plant zones proposed by Guerra-Sommer and
ported a short distance and preserved in lake- or lagoon-margin Cazzulo-Klepzig (1993) seem to reflect primarily changes in the
130 Iannuzzi et al.

physical environment. The reappearance in the Glossopteris/ M., Bossi, G.E., and Andreis, R.S., 1980, Estudo geológico da bacia car-
Rhodeopteridium Zone of taxa characteristic of the lower B. plan- bonífera de Gravataí-Morungava, RS, in Anais, Congresso Brasileiro de
Geologia, 31st, Balneário de Camboriú, Volume 1: Santa Catarina, Brazil,
tiana Zone indicates a strong facies control. In this context, the Sociedade Brasileira de Geologia, p. 266–282.
increase in floral diversity of the Glossopteris/Rhodeopteridium Césari, S.N., 2007, Palynological biozones and radiometric data at the
Zone can be directly linked with the record of new lowland envi- Carboniferous-Permian boundary in western Gondwana: Gondwana
Research, v. 11, no. 4, p. 529–536, doi: 10.1016/j.gr.2006.07.002.
ronments within the Rio Bonito Formation rather than with cli- Césari, S.N., and Gutiérrez, P.R., 2000, Palynostratigraphy of Upper Paleozoic
matic changes. The more physically complex coastal landscape sequences in Central-Western Argentina: Palynology, v. 24, p. 113–146,
that resulted apparently included both new and ancient wetland doi: 10.2113/0240113.
Chaloner, W., Leistikow, K., and Hill, A., 1979, Brasilodendron gen. nov. and
habitats and associated assemblages. However, the nature of the B. pedroanum (Carr.) comb. nov., Permian lycopod from Brazil: Review
plant assemblages (= paleoecologic and taphonomic controls) is of Palaeobotany and Palynology, v. 28, no. 2, p. 117–136.
Cleal, C.J., 1999, Plant macrofossil biostratigraphy, in Jones, T.P., and Rowe,
not yet well elucidated. N.P., eds., Fossil Plants and Spores: Modern Techniques: Geological
Testing the hypothesis delineated in this paper and expand- Society of London, p. 220–224.
ing our conclusions will require (1) an evaluation of taphonomic Daemon, R.F., and Quadros, L.P., 1970, Bioestratigrafia do Neopaleozóico da
Bacia do Paraná, in Anais, Congresso Brasileiro de Geologia, 24th, Brasí-
controls in plant-bearing beds, (2) a better understanding of the lia: Brasília, Sociedade Brasileira de Geologia, p. 359–412.
relation between the plant-bearing strata and their equivalent de Matos, S.L.F., Yamamoto, J.K., Riccomini, C., Hachiro, J., and Tassinari,
palynologic zones, (3) correlation between palynologic and paleo- C.C.G., 2001, Absolute dating of Permian ash-fall in the Rio Bonito
Formation, Paraná Basin: Gondwana Research, v. 4, p. 421–426, doi:
botanic data and the sequence-stratigraphic framework already 10.1016/S1342-937X(05)70341-5.
established in other areas, and (4) improvement of the chrono- Dickins, J.M., 1984, Late Palaeozoic glaciation: BMR Journal of Australian
stratigraphic chart of the Paraná Basin through the discovery of Geology and Geophysics, v. 9, p. 163–169.
Gradstein, F.M., et al. (plus 38 authors), 2004, A geologic time scale 2004:
layers suitable for radiometric dating. Geological Survey of Canada Miscellaneous Report 86, 1 chart.
Guerra-Sommer, M., and Cazzulo-Klepzig, M., 1993, Biostratigraphy of the
ACKNOWLEDGMENTS Southern Brazilian Neopaleozoic Gondwana Sequence: A preliminary
paleobotanical approach, in Compte Rendu, International Congrès de
La Stratigraphie et Géologie du Carbonifère et Permien, 12th, Buenos
This research was partly supported by grants from Aires, 1991, v. 2: Ciudad de Buenos Aires, Argentina, Consejo Nacional
CNPq (RI PQ309322/2007-3, PAS PQ305265/2007-5 and de Investigaciones Científicas y Técnicas, p. 61–72.
Guerra-Sommer, M., Cazzulo-Klepzig, M., Menegat, R., Formoso, M.L.L.,
MH PQ300866/2008-9), and FAPERGS (RI 02/1755-2 and Basei, M.A.S., Barboza, E.G., and Simas, M.W., 2008a, Geochrono-
PROAPP04/1066.0). Many thanks to two anonymous reviewers logical data from Faxinal Coal succession in Southern Paraná Basin: A
who improved the manuscript with many suggestions and helpful preliminary approach combining radiometric U/Pb age and palynostratig-
raphy: Journal of South American Earth Sciences, v. 25, p. 246–256, doi:
comments, and to the editors of this special paper, Luis Buatois 10.1016/j.jsames.2007.06.007.
and Oscar R. López-Gamundí, for kindly inviting the partici- Guerra-Sommer, M., Cazzulo-Klepzig, M., Santos, J.O.S., Hartmann, L.A.,
pation of IGCP Project 471 and for their considerable editorial Ketzer, J.M.M., and Formoso, M.L.L., 2008b, Radiometric age determi-
nation of tonsteins and stratigraphic constraints for the Lower Permian
effort. This is a contribution of the Research Center of Gond- coal succession in Southern Paraná Basin, Brazil: International Journal of
wana (CIGO) to the CNPq (process 483463/2007-8) and IGCP Coal Geology, v. 74, p. 13–27, doi: 10.1016/j.coal.2007.09.005.
Project 471. Holz, M., 1997, Early Permian sequence stratigraphy and paleophysiography of
the Paraná Basin in northeastern Rio Grande do Sul State, Brazil: Anais
da Academia Brasileira de Ciências, v. 69, no. 4, p. 521–543.
REFERENCES CITED Holz, M., 1999, Early Permian sequence stratigraphy and the palaeophysi-
ographic evolution of the Paraná Basin in southernmost Brazil: Journal of
African Earth Sciences, v. 29, no. 1, p. 51–61, doi: 10.1016/S0899-5362
Anderson, J.M., and Anderson, H.M., 1985, Paleoflora of Southern Africa: (99)00079-2.
Prodromus of South African megafloras, Devonian to Lower Cretaceous: Holz, M., 2003, Sequence stratigraphy of a lagoonal estuarine system: An
Rotterdam, A. A. Balkema, 423 p. example from the Lower Permian Rio Bonito Formation, Paraná Basin,
Alves, L.S.R., 1999, Anéis de crescimento em lenhos fósseis como indicadores Brazil: Sedimentary Geology, v. 162, no. 3–4, p. 305–331, doi: 10.1016/
paleoclimáticos no Neopermiano da Bacia do Paraná (Formação Irati e S0037-0738(03)00156-8.
Formação Serra Alta) [Ph.D. thesis]: Porto Alegre, Instituto de Geociên- Holz, M., Vieira, P.E., and Kalkreuth, W., 2000, The Early Permian coal-bearing
cias da Universidade Federal do Rio Grande do Sul. succession of the Paraná Basin in southernmost Brazil: Depositional
Alves, R.G., and Ade, M.V.B., 1996, Sequence stratigraphy and organic petro- model and sequence stratigraphy: Revista Brasileira de Geociências, v. 30,
graphy applied to the study of Candiota Coalfield, RS, South Brazil: Inter- no. 3, p. 420–422.
national Journal of Coal Geology, v. 30, p. 231–248, doi: 10.1016/0166 Holz, M., Küchle, J., Philipp, R.P., Bischoff, A.P., and Arima, N., 2006, Hier-
-5162(95)00041-0. archy of tectonic control on stratigraphic signatures: Base-level changes
Archangelsky, S., González, C.R., Cúneo, N.R., Sabattini, N., Césari, S.N., during the Early Permian in the Paraná Basin, southernmost Brazil: Jour-
Aceñolaza, F.G., Garcia, G.B., Buatois, L.A., Ottone, E., Mazzoni, A.F., nal of South American Earth Sciences, v. 22, p. 185–204, doi: 10.1016/
Hünicken, M.A., and Gutiérrez, P.R., eds., 1996, El Sistema Permico en j.jsames.2006.09.007.
la Republica Argentina y en la Republica Oriental del Uruguay: Córdoba, Iannuzzi, R., 2007, Biostratigraphic versus geochronologic frameworks in the
Academia Nacional de Ciencias, 417 p., lam. I–V. Early Permian from Paraná Basin: Looking forward a possible consen-
Bangert, B., Stollhofen, H., Lorenz, V., and Armstrong, R.L., 1999, The geo- sus, in Iannuzzi, R., and Boardman, D.R., eds., Extended Abstracts, First
chronology and significance of ash-fall tuffs in glacigenic, Carboniferous- Workshop on Problems in Western Gondwana Geology: South America–
Permian Dwyka Group of Namibia and South Africa: Journal of African Africa Correlations: Du Toit Revisited, Gramado, Brazil: Porto Alegre,
Earth Sciences, v. 29, p. 33–49, doi: 10.1016/S0899-5362(99)00078-0. Centro de Investigações do Gondwana, p. 72–77.
Bortoluzzi, C.A., Piccoli, A.E.M., Corrêa da Silva, Z.C., Cazzulo-Klepzig, M., Iannuzzi, R., and Souza, P.A., 2005, Floral succession in the Lower Perm-
Dias-Fabrício, M.E., Silva Fo, B.C., Guerra-Sommer, M., Marques-Toigo, ian deposits of the Brazilian Paraná Basin: An up-to-date overview, in
Southern Brazilian Paraná Basin 131

Lucas, S.G., and Zigler, K.E., eds., The Nonmarine Permian: New Mexico Marques-Toigo, M., Dias-Fabrício, M.E., and Cazzulo-Klepzig, M., 1982,
Museum of Natural History and Science Bulletin 30, p. 144–149. Palynological and paleoecological characterization of Santa Rita Coal-
Iannuzzi, R., Marques-Toigo, M., Scherer, M.S.C., Caravaca, G., Vieira, field, Rio Grande do Sul, Paraná Basin, Lower Permian of Southern Bra-
E.L.C., and Pereira, L.S., 2003a, Phytobiostratigraphical revaluation of zil: Acta Geológica Leopoldense, v. 16, p. 55–74.
the Southern Brazilian Gondwana sequence (Paraná Basin, Lower Perm- Marques-Toigo, M., Dias-Fabrício, M.E., and Cazzulo-Klepzig, M., 1984, A
ian), in Abstracts, International Congress on Carboniferous and Permian sucessão da microflora nas camadas de carvão da bacia carbonífera de
Stratigraphy, 15th: Utrecht, Utrecht University, p. 240–242. Charqueadas–Formação Rio Bonito, RS, Brasil: Boletim IG-USP, v. 15,
Iannuzzi, R., Marques-Toigo, M., Scherer, C.M.S., Caravaca, G., Vieira, C.E.L., p. 65–72.
and Pereira, L.S., 2003b, Reavaliação da fitobioestratigrafia da Seqüência Milani, E.J., França, A.B., and Schneider, R.L., 1994, Bacia do Paraná, in
Gondwanica Sul-Riograndense: Estudo de caso do Afloramento Morro do Feijó, F.J., ed., Cartas estratigráficas das bacias sedimentares brasileiras:
Papaléo (Bacia do Paraná, Permiano Inferior), in Anais, Encontro sobre Rio de Janeiro, Petrobras, Boletim Geociências da Petrobrás, v. 8, no. 1,
estratigrafia do Rio Grande do Sul: Escudos e Bacias, 1st, Porto Alegre: p. 69–82.
Porto Alegre, Universidade Federal do Rio Grande do Sul, p. 182–185. Millan, J.H., 1987, Os pisos florísticos do carvão do Subgrupo Itararé do Estado
Iannuzzi, R., Scherer, C.M.S., Souza, P.A., Holz, M., Caravaca, G., Adami- de São Paulo e suas implicações, in Anais, Congresso Brasileiro de Pale-
Rodrigues, K., Tybusch, G.P., Souza, J.M., Smaniotto, L.P., Fischer, T.V., ontologia, 10th, Rio de Janeiro, Volume 2: Rio de Janeiro, Sociedade Bra-
Silveira, A.S., Lykawka, R., Boardman, D.R., and Barboza, E.G., 2006, sileira de Paleontologia, p. 832–857.
Afloramento Morro do Papaléo, Mariana Pimentel, R.S. Registro ímpar Murphy, M.A., and Salvador, A., 1999, International Stratigraphic Guide: An
da sucessão pós-glacial do Paleozóico da Bacia do Paraná, in Winge, Abridged Version: Episodes, v. 22, p. 255–271.
M., Schobbenhaus, C., Souza, C.R.G., Fernandes, A.C.S., Queiroz, E.T., Nicoll, R.S., and Metcalfe, I., 1997, Early and Middle Permian conodonts from
Berbert-Born, M.L.C., and Campos, D.A., eds., Sítios Geológicos e Pale- the Canning and southern Carnarvon basins, in Shi, G.R., Archbold, N.W.,
ontológicos do Brasil, v. 2, p. 1–13, http://www.unb.br/ig/sigep/sitio101/ and Grover, M., eds., The Permian System: Stratigraphy, Palaeogeography
Sitio101_Morro_do_Papaleo_MarianaPimentelRS.pdf. and Resources: Victoria, Australia, Deakin University and Royal Society
Iannuzzi, R., Scherer, M.S.C., and Caravaca, G., 2007, Taphonomy and paleo- of Victoria, p. 323–343.
ecology of the southern Brazilian Glossopteris Flora (Paraná, Basin, Petri, S., and Souza, P.A., 1993, Síntese dos conhecimentos e novas concep-
Lower Permian), in Díaz-Martínez, E., and Rábano, I., eds., European ções sobre a bioestratigrafia do Subgrupo Itararé, Bacia do Paraná, Brasil:
Meeting on the Palaeontology and Stratigraphy of Latin America, 4th, Revista do Instituto Geológico, v. 14, no. 2, p. 7–18.
Madrid: Publicaciones del Instituto Geológico y Minero de España, Picarelli, A.T., Dias-Fabrício, M.E., and Cazzulo-Klepzig, M., 1987, Consi-
Cuadernos del Museo Geominero, no. 8, p. 201–206. derações sobre a paleoecologia e a palinologia da jazida carbonífera de
Jasper, A., 2004, O modelo deposicional do afloramento Quitéria e a evolu- Santa Terezinha, RS, Brasil: Permiano da Bacia do Paraná, in Actas, Sim-
ção dos biomas úmidos no Permiano Inferior do Sul da Bacia do Paraná pósio Sul-Brasileiro de Geologia, 3rd, Curitiba, Volume 1: Paraná, Brazil,
[Ph.D. dissertation]: Instituto de Geociências da Universidade Federal do Sociedade Brasileira de Geologia, p. 351–372.
Rio Grande do Sul, 248 p. Piccoli, A.E.M., Menegat, R., Guerra-Sommer, M., Marques-Toigo, M., and
Jasper, A., and Guerra-Sommer, M., 1999, Licófitas arborescentes in situ como Porcher, C.C., 1991, Faciologia da seqüência sedimentar nas folhas de
elementos importantes na definição de modelos deposicionais (Formação Quitéria e Várzea do Capivarita, Rio Grande do Sul: Pesquisas, v. 18,
Rio Bonito–Bacia do Paraná–Brasil): Pesquisas, v. 26, p. 49–58. no. 1, p. 31–43.
Jasper, A., Guerra-Sommer, M., Cazzulo-Klepzig, M., and Menegat, R., 2003, Retallack, G.J., 1980, Late Carboniferous to Middle Triassic megafossil floras
The Botrychiopsis genus and its biostratigraphic implications in South- from the Sydney Basin, in Herbert, C., and Helby, R., eds., A Guide to the
ern Paraná Basin: Anais da Academia Brasileira de Ciências, v. 75, no. 4, Sydney Basin: Bulletin of the Geological Survey of New South Wales,
p. 513–535. v. 26, p. 384–430.
Jasper, A., Ricardi-Branco, F., and Guerra-Sommer, M., 2005, Coricladus Rocha-Campos, A.C., and Rösler, O., 1978, Late Paleozoic faunal and floral
quiteriensis gen. et sp. nov., a new conifer in southern-Brazil Gondwana successions in the Paraná Basin, southeastern Brazil: Boletim IG-USP,
(Lower Permian, Paraná Basin): Anais da Academia Brasileira de Ciên- v. 9, p. 1–16.
cias, v. 77, no. 1, p. 157–168. Rohn, R., and Lages, L.C., 2000, Lower Permian Sphenopsids from Cerquil-
Jasper, A., Menegat, R., Guerra-Sommer, M., Cazzulo-Klepzig, M., and Souza, ho, northeastern Paraná Basin, Brazil: Revue Paléobiologie, v. 19, no. 2,
P.A. de, 2006, Depositional cyclicity and paleoecological variability in p. 359–379.
an outcrop of Rio Bonito formation, Paraná Basin, Rio Grande do Sul, Rösler, O., 1978, The Brazilian eogondwanic floral succession: Boletim IG-
Brazil: Journal of South American Earth Sciences, v. 21, p. 276–293, doi: USP, v. 9, p. 85–90.
10.1016/j.jsames.2006.05.002. Santos, R.V., Souza, P.A., Alvarenga, C.J.S., Dantas, E.L., Pimentel, M.M.,
Jones, M.J., and Truswell, E.M., 1992, Late Carboniferous and Early Permian Oliveira, C.G., and Araújo, L.M., 2006, SHRIMP U-Pb Zircon dating
palynostratigraphy of the Joe Joe Group, southern Galilee Basin, Queens- and palynology of bentonitic layers from the Permian Irati Formation,
land, and implications for Gondwana stratigraphy: BMR Journal of Aus- Paraná Basin, Brazil: Gondwana Research, v. 9, p. 456–463, doi: 10.1016/
tralian Geology and Geophysics, v. 13, p. 143–185. j.gr.2005.12.001.
López-Gamundí, O.R., 1989, Postglacial transgressions in Late Paleozoic Schneider, R.L., Mühlmann, H., Tommasi, E., Medeiros, R.A., Daemon, R.F.,
basins of western Argentina: A record of glacioeustatic sea level rise: and Nogueira, A.A., 1974. Revisão estratigráfica da Bacia do Paraná, in
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 71, p. 257–270, Anais, Congresso Brasileiro de Geologia, 28th, Porto Alegre, Volume 1:
doi: 10.1016/0031-0182(89)90054-0. Porto Alegre, Sociedade Brasileira de Geologia, p. 41–66.
López-Gamundí, O.R., Espejo, I.S., Conaghan, P.J., Powell, C.McA., and Souza, J.M., and Iannuzzi, R., 2007, Sementes do gênero Samaropsis Goeppert
Veevers, J.J., 1994, Southern South America, in Veevers, J.J., and Powell, no Permiano Inferior da Bacia do Paraná, sul do Brasil: Revista Brasileira
C.McA., eds., Permian-Triassic Pangean Basins and Foldbelts along the de Paleontologia, v. 10, no. 2, p. 95–106, doi: 10.4072/rbp.2007.2.03.
Panthalassan Margin of Gondwanaland: Geological Society of America Souza, J.M., and Iannuzzi, R., 2009, The genus Cordaicarpus Geinitz in the
Memoir 184, p. 281–329. Lower Permian of the Paraná Basin, Rio Grande do Sul, Brazil: Revista
Maheshwari, H., and Bajpai, U., 2001, Phytostratigraphical succession in the Brasileira de Paleontologia, v. 12, no. 1, p. 5–16, doi: 10.4072/rbp.2009
Glossopteris flora of India: Revista Universidade Guarulhos: Geociên- .1.01.
cias, v. 4, no. 6, p. 22–34. Souza, P.A., 2006, Late Carboniferous palynostratigraphy of the Itararé Sub-
Marques-Toigo, M., 1991, Palynobiostratigraphy of the Southern Brazilian group, northeastern Paraná Basin, Brazil: Review of Palaeobotany and
Neopaleozoic Gondwana sequence, in Proceedings, International Gond- Palynology, v. 138, p. 9–29, doi: 10.1016/j.revpalbo.2005.09.004.
wana Symposium, 7th, São Paulo, 1988: São Paulo, Universidade de São Souza, P.A., and Marques-Toigo, M., 2003, An overview on the palynostratigra-
Paulo, p. 503–515. phy of the Upper Paleozoic strata of the Brazilian Paraná Basin: Revista
Marques-Toigo, M., and Pons, M.E., 1974, Estudo palinológico do furo de del Museo Argentino de Ciências Naturales, n.s., v. 5, no. 2, p. 205–214.
sondagem P7 Malha Oeste da Bacia carbonífera de Iruí, RS, Brasil, in Souza, P.A., and Marques-Toigo, M., 2005, Progress on the palynostratigraphy
Anais, Congresso Brasileiro de Geologia, 28th, Porto Alegre: Porto of the Paraná strata in Rio Grande do Sul State, Paraná Basin, Brazil:
Alegre, Sociedade Brasileira de Geologia, p. 277–288. Anais da Academia Brasileira de Ciências, v. 77, no. 2, p. 353–365.
132 Iannuzzi et al.

Stephenson, M.H., 2008, A review of the palynostratigraphy of Gondwanan Veevers, J.J., and Powell, C.McA., 1987, Late Paleozoic glacial episodes
Late Carboniferous to Early Permian glacigene successions, in Fielding, in Gondwanaland reflected in transgressive-regressive depositional
C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice sequences in Euramerica: Geological Society of America Bulletin, v. 98,
Age in Time and Space: Geological Society of America Special Paper p. 475–487, doi: 10.1130/0016-7606(1987)98<475:LPGEIG>2.0.CO;2.
441, p. 115–129. Vieira, C.E.L., Iannuzzi, R., and Guerra-Sommer, M., 2007, Revisão de Pecop-
Stephenson, M.H., and Filatoff, J., 2000, Correlation of Carboniferous-Permian terídeas Polimórficas do Neopaleozóico da América do Sul: Revista
assemblages from Oman and Saudi Arabia, in Al-Hajri, S., and Owens, Brasileira de Paleontologia, v. 10, no. 2, p. 107–116, doi: 10.4072/rbp
B., eds., Stratigraphic Palynology of the Palaeozoic of Saudi Arabia: Gulf .2007.2.04.
Petrolink, GeoArabia, Special Publication 1, p. 168–91. Zalán, P.V., Wolff, S., Conceição, J.C., Marques, A., Astolfi, M.A., Vieira, I.S.,
Stollhofen, H., Stanistreet, I.G., Bangert, B., and Grill, H., 2000, Tuffs, tec- Appi, V.T., and Zanotto, O.A., 1990, Bacia do Paraná, in Raja Gabaglia,
tonism and glacially related sea-level changes, Carboniferous-Permian, G.P., and Milani, E.J., eds., Origem e evolução das bacias sedimentares:
southern Namibia: Palaeogeography, Palaeoclimatology, Palaeoecology, Rio de Janeiro, Petrobrás, p. 135–168.
v. 161, p. 127–150, doi: 10.1016/S0031-0182(00)00120-6. Ziegler, A.M., Gibbs, M.T., and Huber, M.L., 1997, A mini-atlas of oceanic
Tybusch, G.P., 2005, Análise taxonômica de tipos foliares de glossopterídeas water masses in the Permian period, in Shi, G.R., Archbold, N.W., and
em depósitos do Permiano Inferior da Bacia do Paraná, Rio Grande do Grover, M., eds., The Permian System: Stratigraphy, Palaeogeography
Sul: Rubidgea spp., Gangamopteris spp., Glossopteris occidentalis, and Resources: Victoria, Australia, Deakin University and Royal Society
G. browniana [Master’s thesis]: Porto Alegre, Instituto de Geociências da of Victoria, p. 419–461.
Universidade Federal do Rio Grande do Sul, 102 p.
Tybusch, G.P., and Iannuzzi, R., 2008, Reavaliação taxonômica dos gêneros
Gangamopteris e Rubidgea, Permiano Inferior da Bacia do Paraná, Brasil:
Revista Brasileira de Paleontologia, v. 11, no. 2, p. 73–86, doi: 10.4072/
rbp.2008.2.01. MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

“Levipustula Fauna” in central-western Argentina


and its relationships with the Carboniferous glacial event
in the southwestern Gondwanan margin

Gabriela A. Cisterna*
CONICET, Fundación Miguel Lillo, Instituto de Paleontología, Area Geología. Miguel Lillo 251, 4000 Tucumán, Argentina

Andrea F. Sterren*
CONICET, CIPAL, Centro de Investigaciones Paleobiológicas, Facultad de Ciencias Exactas, Físicas y Naturales,
Universidad Nacional de Córdoba. Av. Vélez Sarsfield 299, X5000JJC Córdoba, Argentina

ABSTRACT

The Levipustula Fauna (included in the Levipustula levis Zone) is a relatively diver-
sified fossil assemblage composed of brachiopods, bivalves, bryozoans, gastropods,
and crinoids. This fauna usually is associated with glaciomarine sequences related
to the Carboniferous glacial event that affected the southwestern Gondwanan mar-
gin. The Levipustula Fauna has been identified in different units (e.g., Hoyada Verde,
La Capilla, Leoncito, and Yalguaraz Formations) exposed in the Calingasta-Uspallata
Basin. The Hoyada Verde Formation, herein proposed as a key section, contains the
most complete record of the Levipustula Fauna. A detailed compositional, taphonomic,
and paleoecological study of this section allows us to propose two associations within
the so-called Levipustula Zone: the “Intraglacial Levipustula Fauna,” present in the
diamictite-dominated lower part, and the “Postglacial Levipustula Fauna,” dominant
in the upper part of section. The fossils of the “Intraglacial Levipustula Fauna” are
scarce and poorly diversified. These two features suggest environmentally stressed
conditions, probably related to low temperatures in areas close to glaciers. In com-
parison, the “Postglacial Levipustula Fauna,” relatively more abundant and diverse,
exhibits compositional variations that could be explained by paleoenvironmental
changes associated with fluctuations in substratum and food supply, such as those
identified in modern ecosystems. The identification of the “Intraglacial Levipustula
Fauna” and the “Postglacial Levipustula Fauna” may constitute a new tool for under-
standing the particular relationship between faunal assemblages and climatic varia-
tions linked to the Gondwanan glaciation in the Calingasta-Uspallata Basin. Also,
the new “Intraglacial Levipustula Fauna” identified in the Hoyada Verde Formation
would have biostratigraphical and paleogeographical implications in intra- and inter-
basinal correlations.

*[email protected]; [email protected]

Cisterna, G.A., and Sterren, A.F., 2010, “Levipustula Fauna” in central-western Argentina and its relationships with the Carboniferous glacial event in the south-
western Gondwanan margin, in López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana:
Geological Society of America Special Paper 468, p. 133–147, doi: 10.1130/2010.2468(06). For permission to copy, contact [email protected]. ©2010
The Geological Society of America. All rights reserved.

133
134 Cisterna and Sterren

INTRODUCTION ren, 2004, 2008; Sterren, 2003, 2005) in the classical localities
with Levipustula Fauna (Hoyada Verde, Leoncito, and La Capilla
The marine invertebrates of the Levipustula Fauna constitute Formations) have recognized significant taxonomic, paleoeco-
a relatively diversified fossil assemblage composed of brachio- logic, and taphonomic variations. These variations, as well as the
pods, bivalves, bryozoans, gastropods, and crinoids. This fauna stratigraphic position of this fauna along the glacial sequence,
can be considered the most conspicuous element in different suggest an important paleoenvironmental control.
localities in the Calingasta-Uspallata Basin (Fig. 1A), where it This paper provides a review of the paleogeographic/
usually is associated with glaciomarine deposits. Previous studies paleoclimatic and biostratigraphical framework of the Levipus-
(Taboada and Cisterna, 1996; Cisterna, 1999; Cisterna and Ster- tula Fauna in the Calingasta-Uspallata Basin. In particular, we

LA RIOJA
INA
NT

SAN JUAN
GE
AR

Cas 69º 30´


Vie taño SAN JUAN
MENDOZA jo R
iver
San Jua
n River
70° W 66° La Capilla
Jagüé Calin
Río del Peñón gasta
Rive
r

RANGE
CALINGASTA
nco

BARREAL Hoyada
C H I L E

Verde
Basin
La Punilla
Río Bla

Villa Unión

DEL TONTAL
er
Riv

La Rioja
tos

Leoncito
Pa
C

30° N Malimán DE RA
VOLCANIC AR

los
in

Patquía
Huaco
L N NG
De
llata Bas

Paganzo 149
AR E

32º
Paganzo Basin
AN
Invernada

GE

JO
RAN
a - Uspa

E
G
AN
RE

Mascasín Yalguaraz
TIG

TA

San Juan
Tontal
t

LA

Barreal
Calingas

DEL

AL
SP

32°
U

MENDOZA USPALLATA
Uspallata
d oza
Mendoza Men r 30 km
A 100 km Positive areas
B Riv e

Figure 1. Location maps showing the paleogeography and geography of the Calingasta-Uspallata Basin with the Levipustula Fauna
occurrences.
“Levipustula Fauna” in central-western Argentina 135

analyze the relationship of this fauna with the Carboniferous gla- The Levipustula Fauna is usually considered to be associated
cial event in the Hoyada Verde Formation, herein proposed as with the postglacial event (López-Gamundí, 1989, 1990; López-
a key section. Because brachiopods and bivalves are the most Gamundí and Espejo, 1993; López-Gamundí and Martínez,
abundant and diversified invertebrates in the Levipustula Fauna, 2000). However, the position of this fauna in the glacial sequence
the present study will be focused on these two groups. of the Hoyada Verde Formation, as well as its taxonomic and
paleoecological/taphonomic variations, suggests the presence of
PALEOGEOGRAPHIC AND PALEOCLIMATIC a recently identified “Intraglacial Levipustula Fauna” (Cisterna
SETTING and Sterren, 2008), besides the typical Levipustula Fauna associ-
ated with the postglacial transgression.
During the late Paleozoic all the Gondwanan subcontinents
were affected by an extensive glaciation (Hambrey and Harland, GEOGRAPHIC AND STRATIGRAPHIC
1981). Paleomagnetic and paleoclimatic data suggest that the DISTRIBUTION OF THE LEVIPUSTULA FAUNA
pole moved across South America, Southern Africa, Antarctica,
and Australia throughout the Carboniferous-Permian transi- The Levipustula Fauna has been identified in different
tion as a result of the apparent path of polar wander (Crowell, diamictite-bearing sections of the Calingasta-Uspallata Basin
1983; Caputo and Crowell, 1985; Scotese and Barret, 1990; (Fig. 1B). Detailed studies on the brachiopods and bivalves of
Scotese and McKerrow, 1990). Three glacial episodes have been this fauna suggest significant taxonomic, taphonomic and paleo-
suggested within this late Paleozoic paleoclimatic mega-event ecological variations (Cisterna and Sterren, 2003, 2004, 2008;
(López-Gamundí, 1997). In western Gondwana, the earliest gla- Sterren, 2005). The material studied in this contribution was col-
cial episode is recorded in sediments of latest Devonian–earliest lected from the classical localities of the Levipustula Fauna of the
Carboniferous exposed in the Solimões and Amazonas Basins Hoyada Verde, La Capilla and Leoncito Formations (Table 1).
of Brazil and the Titicaca Lake region of western Bolivia. The We provide herein a review of the geographic, stratigraphic, and
second glacial episode, identified in the Late Carboniferous sedi- paleontological aspects of these units.
ments (Namurian-Westphalian) along the southwestern South Hoyada Verde Formation. The Hoyada Verde Formation
American basins, is closely associated with the Levipustula (Mésigos, 1953) is exposed in the eroded core of a broad north-
Fauna herein studied. The latest episode has been identified in south anticline (the Hoyada Verde anticline) located 3 km east of
latest Carboniferous to earliest Permian sediments in basins of Barreal village (Fig. 2). The basal contact of the Hoyada Verde
eastern South America (Paraná, Sauce Grande, and Malvinas Formation is unknown and the top is overlain by the Tres Saltos
Basins) and South Africa. Formation with an angular unconformity (Fig. 3). The Hoyada
The late Carboniferous glacial episode is best documented Verde section is characterized by a glaciogenic sequence (dia-
along western Argentina, particularly in the Calingasta-Uspallata mictite and pebbly [dropstone] shale facies) that grades upward
Basin, a back-arc basin formed behind an active magmatic arc into postglacial, dropstone-free shales that contain the typical
(Ramos et al., 1986; Ramos and Palma, 1996). A combination Levipustula Fauna. An intertill boulder pavement, considered
of high latitude and altitude allowed the formation of ice centers of subglacial origin (López Gamundí and Martínez, 2000), is
along the margins of this basin (López-Gamundí, 1997). In this present in the upper part of the Hoyada Verde Formation and is
sense, a local north-trending high called the Proto-Precordillera associated with the diamictic facies. To the uppermost part of
(Amos and Rolleri, 1965), located to the east of the Calingasta- the section (Fig. 3), the sequence is characterized by mudstones
Uspallata Basin (Fig. 1A), might have exerted an important with interbedded fine sandstones arranged in coarsening and
topographic control during this glacial stage (López-Gamundí thickening-upward sequences. This interval has been interpreted
and Rosello, 1995). Striated and boulder pavements within the as the transition from the offshore to the lower shoreface (Buatois
Carboniferous glaciomarine sequences (e.g., Hoyada Verde and and Limarino, 2003), and a succession of fossils traces composed
Leoncito Formations) have been recognized. Strong northward of Psammichnites implexus (Rindsberg) and Psammichnites
components in the paleo-ice flow indicators, as well as paleocur- plummeri (Fenton and Fenton) has been recognized by Mángano
rents from the overlying shallow marine sandstones, suggest a et al. (2003). In the top of the Hoyada Verde Formation the upper-
common regional paleoslope for both glacial and early postgla- most fossiliferous horizon identified is located in a coquina with
cial times (López-Gamundí and Martínez, 2000). gastropods such as Peruvispira reedi Sabattini and Mourlonia sp.
The Carboniferous glacial sequences of the Calingasta- (Sabattini, 1980).
Uspallata Basin are characterized by glaciomarine diamictites The marine invertebrate assemblage associated with the post-
that grade upward to postglacial open-marine fine clastics. The glacial shales is composed of brachiopods, bivalves, bryozoans
presence of the marine fine clastics resting on diamictites has (the fenestellids Fenestella sanjuanesis Sabattini, F. barrealen-
been interpreted as the sedimentary response to a glacio-eustatic sis Sabattini, F. altispinosa Sabattini, and Polypora neerkolensis
sea-level rise, expressed as a transgression in stratigraphic terms, Crockfor, described by Sabattini [1972]) and scarce gastropods
that occurred during the glacial retreat subsequent to a wide- and crinoids. The brachiopods assemblage is dominated by the
spread glaciation (López-Gamundí, 1989, 1990, this volume). Levipustula-Costuloplica-Kitakamithyris association with the
136 Cisterna and Sterren

TABLE 1. TAXONOMIC COMPOSITION OF THE LEVIPUSTULA FAUNA IN THE CALINGASTA-USPALLATA BASIN

HOYADA VERDE FM. LA CAPILLA FM. LEONCITO FM.


Taxon
Lower part Upper part South (LC) North (LJ)

Levipustula levis Maxwell C D D --- M


Costuloplica leoncitensis (Harrington) C D D M D
Kitakamithyris booralensis (Campbell) --- --- D D ---
Kitakamithyris sp. --- D D D ---
Beecheria sp. --- D D M D
Spiriferellina octoplicata (Sowerby) C D M M D
Spiriferellina sp. --- --- D M ---
Septosyringothyris keideli (Harrington) --- --- --- M D
Orthotetoidea indet. --- --- --- M M
Nuculopsis sp. C --- --- --- M
Phestia sp. aff. P. bellistriata (Stevens) --- D --- --- ---
Phestia sp. C --- D --- M
Palaeolima retifera (Shumard) C D --- --- ---
Streblochondria stappenbecki Reed --- D --- --- ---
Streblochondria sanjuanensis Sterren --- D --- --- ---
Aviculopecten barrealensis Reed --- --- D --- ---
Schizodus sp. --- --- D M M
Oriocrassatella ? sp. --- --- D --- ---
Pleurophorella ? sp. --- --- D --- M
Oriocrassatella andina González --- --- --- D ---
Myofossa calingastensis González --- --- --- D ---
Leptodesma (Leiopteria) sp. --- --- --- D ---
Cypricardinia sp. --- --- --- D ---
Promytilus sp. --- --- --- D ---
Pyramus ? sp. --- --- --- D ---
Naiadites sp. --- --- --- --- D
Barrealispira mesigosi Taboada and Sabattini D --- --- --- M
Murlonia striata (Sowerby) D --- --- --- ---
Ptychomphalina sabattinii Taboada --- --- --- D ---
Ptychomphalina turgentis Taboada --- --- --- D ---
Ptychomphalina cf. kuttungensis Taboada --- --- --- D ---
Peruvispira reedi Sabattini --- D --- D ---
Straparollus (Euomphalus) sp. --- --- --- M ---
Leptoptygma sp. --- --- --- M ---
Fenestella sanjuanensis Sabattini --- D --- --- ---
Fenestella barrealensis Sabattini --- D --- --- ---
Fenestella altispinosa Sabattini --- D --- --- ---
Fenestella sp. --- --- --- --- M
Polypora neerkolensis Crockford --- D --- --- ---
Cladochonus harringtoni Sabattini --- D --- --- ---
Sphenotallus stubblefieldi Schmidt and Teichmüller D --- --- --- ---
Bryozoans indet. C --- M M M
Crinoids indet. C C M --- M
Ostracods indet. C --- --- --- M

C - Cited in this work


D - Described in previous work
M - Mentioned in previous work

species Levipustula levis Maxwell, Costuloplica leoncitensis Formation, the gastropods Neilsonia? sp., Neoplatyteichum bar-
(Harrington), and Kitakamithyris sp., usually accompanied by realense (Reed), Barrealispira mesigosi Taboada and Sabattini,
Beecheria sp. and “Spiriferellina” octoplicata (Sowerby). The Ptychomphalina striata (Sowerby), the annelid Sphenotallus
bivalve fauna studied by Sterren (2003) is composed of Phestia stubblefieldi Schmidt and Teichmüller, and the brachiopods Levi-
sp. aff. P. bellistriata (Stevens), Palaeolima retifera (Shumard), pustula levis Maxwell and Kitakamithyris sp. have been identi-
Streblochondria sanjuanensis Sterren, and S. stappenbecki Reed. fied in the lower part of the section (Sabattini, 1980; Taboada
Although this is the main faunal assemblage in the Hoyada Verde and Sabattini, 1987; Taboada, 1997). Taboada (1997) has also
“Levipustula Fauna” in central-western Argentina 137

69°25´ W
De Quaternary
La
Lie
bre Lower Triassic
Cre
ek Pituil Group
El Paso Fm.
Hoyada Verde Fm.
De
Hilario Fm. (Lower Paleozoic)

Un
BARREAL

Sa
Anticline

lto
Cr
Tre Fault

ee
s Sa

k
ltos Strike and dip
Cre
ek

31°40´ N

De los Patos River

y
El
BARREAL

Va l l e
oy
Cr
ee
k
Hoyada Verde

gasta
0 0.5 1
km

Calin
149

Modified from Mésigos (1953) 2 km

Figure 2. Generalized geological map showing the distribution of the Hoyada Verde Formation outcrops. Modified from
Mésigos (1953).

suggested the presence of brachiopods, probably the species is located 1400 m north La Capilla village, along the road that
Levipustula levis Maxwell, associated with the diamictite-rich connects Calingasta and the Castaño Viejo mines, ~300 m west
glacial section. of Las Cambachas. This 39-m-thick section is essentially com-
The Hoyada Verde Formation has been considered chrono- posed of sandstones, gray-green mudstones, coquinoid lenticular
logically equivalent to the El Paso Formation (Amos and Rol- beds, and scarce fine conglomerates (Amos et al., 1963). The
leri, 1965; Amos and López-Gamundi, 1981; González, 1990), Levipustula Fauna occurs in the upper part of the section and
a diamictite-rich section exposed in the southernmost part of is taxonomically close to that present in the Hoyada Verde For-
the Barreal hill (Fig. 2). However, there no physical continu- mation. However, the taxonomic differences among bivalves
ity between the two units, and the brachiopods identified in the (Phestia sp., Aviculopecten barrealensis Reed, Schizodus sp.,
El Paso Formation (Micraphelia indianae Simanauskas and Cis- Oriocrassatella? sp., and Pleurophorella? sp.) are more impor-
terna, Tuberculatella peregrina (Reed), Aseptella aff. A. patri- tant than those of the brachiopods (Cisterna and Sterren, 2003,
ciae Simanauskas, and Rhipidomella? sp.) would exhibit latest 2004; Sterren, 2005).
Carboniferous biostratigraphical affinities (Simanauskas and Other outcrops of the La Capilla Formation are located
Cisterna, 2001). ~5 km north of La Capilla village, in a locality known as
La Capilla Formation. Outcrops of the La Capilla For- Las Juntas. Sessarego and Amos (1987) recognized two mem-
mation (Amos et al., 1963) have been recognized in two sec- bers: a lower member, where the Levipustula Fauna occurs, has
tors in the Calingasta area (Fig. 1B). One of these exposures been interpreted as deposited in a proximal glaciomarine setting
138 Cisterna and Sterren

Tres Saltos Fm.

PG

Shales

Laminated shales with dropstones Figure 3. Stratigraphic section of the Hoyada


Verde Formation (modified from López-Gamundí,
1983) and vertical distribution of the Intraglacial
Fine-grained sandstones (IG) and Postglacial Levipustula Fauna (PG).

Stratified pebbly mudstone


Massive bouldery to pebbly
sandy mudstone
Striated boulder pavement

IG Levipustula fauna

Gastropods
20 m

Ichnofossils

with deltaic influence (Sessarego and Amos, 1987), and an upper Leoncito Formation. Outcrops of the Leoncito Formation
member, consisting of deltaic greenish brown sandstones and (Baldis, 1964) are located ~22 km southeast of Barreal village on
mudstones (Vallecillo and Bercowski, 1998). The Levipustula the western flank of the Precordillera (Fig. 1B), along the south-
Fauna in this section is not well known. The brachiopod assem- ern margin of the Las Cabeceras river. This section is dominated
blage appears to be dominated by the genus Kitakamithyris, with by sandy facies and the diamictite beds appear to the top of the
Kitakamithyris booralensis (Campbell) and Kitakamithyris sp., section, where a striated pavement has been also identified. This
accompanied by Septosyringothyris aff. S. keideli (Harrington) pavement is shaped on bioturbated fine-grained sandstones with
and Orthotetoidea indet. González and Taboada (1987) and plant fragments and the surface is covered by a massive diamic-
González (2002) have also suggested the presence of the brachi- tite (López-Gamundí and Martínez, 2000). In the Leoncito For-
opods Spiriferellina octoplicata (Sowerby), Costuloplica leon- mation section the Levipustula Fauna occurs in sandstone and
citencis (Harrington), Septosyringothyris keideli (Harrington), mudstone horizons, located below the glacial diamictic beds in
Kitakamithyris septata? (Chronic), Lingula sp., and Chonetacea an interval ~11 m thick. The fauna associated with the sandstones
indet.; the bivalves Oriocrassatella andina González, Myofossa appears concentrated in distinct lenses and is composed of the
calingastensis González, Leptodesma (Leiopteria) sp., Schizodus Septosyringothyris keideli–Costuloplica leoncitensis brachiopod
sp., Promytilus sp., Pyramus? sp., and Cypricardinia? sp.; the assemblage, accompanied of “Spiriferellina” octoplicata Sow-
gastropods Peruvispira reedi Sabattini, Peruvispira cf. kuttun- erby and very scarce Beecheria sp. and Levipustula levis Max-
gensis Campbell, Murlonia ssp., Straparollus (Euomphalus) sp., well; bivalves (Phestia sp., Schizodus sp. and Pleurophorella?
Leptoptygma sp.; and the bryozoans Fenestella sp. sp.), bryozoans (Fenestella? sp.); and gastropods (Barrealispira?
“Levipustula Fauna” in central-western Argentina 139

sp.). The fauna from the mudstone-dominated interval is quite levis Zone in Precordillera was discussed by Taboada (1997),
different and composed of brachiopods (Levipustula levis and who recognized a mudstone-rich interval in the lowermost Hoya-
Beecheria sp.), bivalves (Nuculopsis? and bivalvia indet.), ostra- da Verde Formation with Barrealispira mesigosi Taboada and
cods (probably some Aurykirkbya), gastropods, crinoids, and Sabattini, Ptychomphalina striata (Sowerby) and Sphenotallus
bryozoans (Cisterna and Sterren, 2008). stubblefieldi Schmidt and Teichmüller, associated with Levipus-
Although we have not conducted fieldwork in the Yalguaraz tula levis Maxwell. Such an assemblage would indicate an early
Formation outcrops, we are tempted to state that this unit con- Namurian age. A recent biostratigraphic review of the Hoyada
tains some diagnostic elements of the Levipustula Fauna. The Verde Formation (Sterren and Cisterna, 2006) indicates that the
Yalguaraz Formation (Amos and Rolleri, 1965) is exposed on fauna suggested by Taboada (1997) occurs within the diamictic
the west flank of the Cordillera del Tigre, close to the boundary section. There are no diagnostic elements to suggest the precise
between San Juan and Mendoza provinces (Fig. 1B). In the type upper limit of this fauna but, for the moment, its biostratigraphic
section (Arroyo del Tigre Creek), this unit is characterized by relationships would indicate a Westphalian age (González, 1981,
a predominantly diamictic sequence with mudstones and sand- 1990, 1993; Archangelsky et al., 1987).
stones increasing toward the upper part. An intertill striated pave- Brachiopods of the Levipustula levis Zone in the Precordille-
ment on the top of this section has been also suggested (Taboada, ra of Western Argentina appear to be the main tool for biostrati-
1997). The marine fossil assemblage, present in mudstones and graphic correlation, and their Australian faunal affinities would be
sandstones of the middle part of the section, is composed of bra- essentially based on the common species Levipustula levis Max-
chiopods, gastropods, bryozoans and bivalves (Taboada and Car- well, Kitakamithyris booralensis (Campbell), and Kitakamithy-
rizo, 1992; Taboada, 1997). From this assemblage, Taboada and ris immensa (Campbell) (Taboada and Cisterna, 1996). Although
Cisterna (1996) have described the brachiopods Kitakamithyris there are previous systematic studies (Reed, 1927; Keidel and
immensa (Campbell) and Torynifer tigrensis Taboada and Cis- Harrington, 1938; Amos et al., 1963; Lech, 1989; Taboada and
terna, which can be considered conspicuous elements of the Levi- Cisterna, 1996), a taxonomic review of new collections of Levi-
pustula Fauna. pustula Fauna brachiopods from Precordillera started by one the
authors (G.A.C.) can shed further light on some of the current
AGE AND FAUNAL AFFINITIES WITH correlation problems. In this sense, the first problem related to
OTHER BASINS the systematics of this fauna’s brachiopods is the original defini-
tion of Levipustula levis in Australia (Maxwell, 1951). This spe-
The Levipustula Fauna is known from eastern Australia cies was described from different units in eastern Australia, such
where it is also characterized by a low-diversity, cold-water as the Booral Formation in New South Wales (Campbell, 1961)
assemblage dominated by brachiopods referred to the Levipustula and the Poperima Formation and Branch Creek Formation in
levis Zone (Campbell and McKellar, 1969; Jones et al., 1973). It Queensland (Maxwell, 1964). However, specimens from differ-
has not been associated with a specific reference section within ent localities assigned to Levipustula levis appear to have distinct
eastern Australian basins but Roberts et al. (1976) have provided species-diagnostic features.
a summary of its composition, stratigraphic occurrence, and Other South American basins where the Levipustula Fauna
relationships with other zones. The age of this zone was widely have also been identified are the Tepuel-Genoa Basin in south-
discussed by different authors (Campbell, 1961; McKellar, western Argentina and the Tarija Basin in Bolivia. In the
1965; Lindsay, 1969; Jones et al., 1973; Roberts, 1976; Roberts Tepuel-Genoa basin, the Levipustula levis Zone was previously
et al., 1976, 1993) and finally assigned, on the basis of faunal recognized in the Pampa de Tepuel and Las Salinas Forma-
considerations, to the Namurian–Westphalian interval (Roberts tions (Amos et al., 1973). However, the brachiopods from the
et al., 1976). However, sensitive high-resolution ion microprobe Tepuel-Genoa basin originally assigned to the Levipustula levis
(SHRIMP) zircon dating of volcanic horizons in the glaciogenic Maxwell by Amos (1960) were included in the synonymy of
sediments of the Seaham Formation (Southern New England Lanipustula patagoniensis Simanauskas (1996) and Verchoja-
Orogen) appears to confine the age of this zone to early Namu- nia inacayali Taboada (2008). The genus Lanipustula proposed
rian (Roberts et al., 1995). by Klets (1983) is very close to Levipustula Maxwell but it has
In Argentina the Levipustula levis Zone was defined by been differentiated by the disposition of the cardinal ridges and
Amos and Rolleri (1965) in the Calingasta-Uspallata Basin. The the shape of the anterior adductor muscle scars (Simanauskas,
La Capilla Formation has been proposed as the reference section 1996). Recent studies of the genera Levipustula and Lanipustula
of the Levipustula levis Zone, but a detailed study in the Hoyada in Argentina suggest that the diagnostic features proposed for
Verde Formation suggests that the latter unit contains the most distinguishing these genera by Klets (1983) can be considered
complete record of the Levipustula Fauna in the basin. of intraspecific hierarchy (Taboada, 2006). Instead, other main
The age of the Levipustula levis Zone in Argentina was con- differences, such as the abundance of external dorsal spines and
sidered Namurian-Westphalian owing to its Australian affinity variable rugae and/or lamination on both valves developed in
(González, 1981; Archangelsky et al., 1987; Archangelsky et al., Lanipustula, clearly separate it from Levipustula (A.C. Taboada,
1996; Taboada, 1997). The lower limit of age for the Levipustula 2010, personal commun.).
140 Cisterna and Sterren

A multivariate analysis of the Levipustula levis Zone in the mentary bryozoans. The brachiopods recognized are Levipustula
Tepuel-Genoa Basin was conducted by Simanauskas and Sabatti- levis? Maxwell and Levipustulini indet., apparently the dominant
ni (1997), who subdivided the zone into the late Carboniferous species, and very scarce specimens of “Spiriferellina” octopli-
Lanipustula Zone, the Early Permian Pyramus faunule (Asse- cata and Spiriferidae indet. The bivalves distinguished in this
lian), and the Tuberculatella Zone (Sakmarian). The Lanipustula fauna are Nuculopsis sp., Phestia sp. aff. P. bellistriata (Stevens)
Zone was referred to the lower part (“Fenestella and Productus” and Palaeolima retifera (Shumard). A preliminary study of the
horizon) of the Pampa de Tepuel Formation (Freytes, 1971) and fauna associated with the brachiopods and bivalves has allowed
to the lower member of the Las Salinas Formation (González, us to recognize the ostracods Kirkbyidae indet. (probably the
1972); Simanauskas and Sabattini (1997) suggested a Namurian– genus Aurykirkbya Sohn), the annelids Sphenotallus stubblefieldi
Stephanian age for this zone. Schmidt and Teichmüller, the gastropods Barrealispira sp., and
In the Tarija Basin the Levipustula Fauna is not well known the bryozoans Fenestella? sp.
but an invertebrate marine assemblage composed of Levipus- The intraglacial fauna is scarce, poorly diversified, and very
tula levis Maxwell, Cypricardinia? boliviana Rocha-Campos, scattered within thick mudstone packages. The bioclasts show a
Carvalho and Amos, Limipecten cf. L. burnettensis Maxwell, random distribution, degree of fragmentation is low, and delicate
Stuchburia sp., Myonia sp., and Mourlonia balapucense Rocha- details of the fine sculpture are preserved. In addition, individual
Campos, Carvalho and Amos was described from the upper part shells are generally small, exhibiting a relatively wide range of
of the Taiguatí Formation (Bolivia) by Rocha-Campos et al. valve sizes, from 0.3 to 2 cm. Shells of both brachiopods and
(1977). The bivalves Naiadites cf. N. modiolaris (Sowerby) and bivalves are mostly disarticulated; however, specimens belong-
Wilkingia cf. W. elliptica (Phillips) have been identified from the ing to the bivalve family Nuculanidae (Nuculopsis sp., Phestia
Taiguatí Formation as well (Trujillo Ikeda, 1989). sp. aff. P. bellistriata (Stevens)), are exceptionally found with
articulated valves.
LEVIPUSTULA FAUNA IN THE HOYADA VERDE The attributes described for these fossils suggest a fauna
FORMATION buried in situ. Features such as its low diversity and abundance
indicate environmentally stressed conditions probably related to
The typical Levipustula Fauna can be easily identified glacial environment. The presence of articulated bivalves might
associated with mudstone facies, located above of the glacial be related to minor reworking and transport, and conditions of
diamictic sequence, in the upper part of the Hoyada Verde For- sudden burial (Kidwell and Bosence, 1991; Aigner, 1985; Peter-
mation. However, a detailed study along this section indicates son, 1985) when the ligament was still active (Fürsich and Hein-
that a scattered and very poorly diversified faunal assemblage berg, 1983). In addition, low temperatures might have delayed the
can be recognized in the lower part of the section, interbedded decomposition of the soft tissues that unite the valves (Kidwell
with diamictic horizons. These faunas are herein proposed as and Baumiller, 1990).
the “Postglacial Levipustula Fauna” (Fig. 3, PG) and “Intra- The bivalves that characterize this interval, Nuculopsis and
glacial Levipustula Fauna” (Fig. 3, IG), respectively, and their Phestia, are commonly found associated with stress conditions
compositional, taphonomic, and paleoecological features are (Sterren, 2000; Simanauskas and Cisterna, 2000; Lebold and
discussed. Kammer, 2006). The abundance of the eurytopic bivalve Nucu-
lopsis is significant because the deposit-feeding nuculid bivalves
The “Intraglacial Levipustula Fauna” are common component of fossil assemblages in oxygen-deficient
basins (Kammer et al., 1986). The annelid Sphenotallus stubble-
The outcrops that contain the “Intraglacial Levipustula fieldi is another conspicuous element in the “Intraglacial Levipus-
Fauna” can be recognized in the exposed core of the Hoyada tula Fauna.” The annelids are considered stress-tolerant taxa and
Verde anticline. Therefore, these mudstones with fauna have been their presence in glaciomarine sequences indicates a wide range
usually referred to the lowermost part of the section (Mésigos, of temperature, salinity, and oxygen tolerance (Collinson et al.,
1953; Taboada, 1997). Recent field work has allowed us to rec- 1994; Chakraborty and Bhattacharya, 2005).
ognize the correct location of this fossiliferous interval and its The presence of glendonite concretions and dropstones in
stratigraphic relationships (Fig. 3). the stratigraphic interval that contains the “Intraglacial Levipus-
The new fossil assemblage appears in the lower part of the tula Fauna” suggests extreme environmental conditions related to
Hoyada Verde Formation, interbedded with diamictic horizons, in the glacial proximity. The glendonite, a carbonate pseudomorph
a 20-m-thick interval made up of grayish, laminated mudstones that might indicate freezing water (González, 1980; Swainson
characterized by the presence of glendonites and dropstones. and Hammond, 2001; McLachlan et al., 2001) has been also
The “Intraglacial Levipustula Fauna” (Fig. 4) is characterized recorded in similar Carboniferous and Permian sequences of
by a monotonous assemblage of marine invertebrates and stems Argentine Patagonia (Tepuel-Genoa basin, González, 1980), as
of “Dadoxilon.” The fauna is dominated by brachiopods, bivalves, well as in South Africa and Australia (McLachlan et al., 2001;
and annelids, accompanied by gastropods, ostracods, and frag- Thomas et al., 2007; López-Gamundí, this volume).
B
A
E C

D
F
H

G I

J K L
Figure 4. “Intraglacial Levipustula Fauna.” (A–D, G) Levipustula levis? Maxwell. (A, G) internal and external mold
of ventral valve, CEGH-UNC 22171 (×3); (B) internal mold of ventral valve, CEGH-UNC 22172 (×2.5); (C) external
mold of ventral valve, CEGH-UNC 22173 (×3); (D) internal mold of ventral valve, CEGH-UNC 22175 (×2.5); (E) Levi-
pustululini indet., internal mold of ventral valve, CEGH-UNC 22174 (×2.5); (F) “Spiriferellina” octoplicata (Sow-
erby) fragmentary ventral valve, CEGH-UNC 22176 (×4); (H) Phestia sp. aff. P. bellistriata (Stevens) outer view of left
valve, CEGH-UNC 22160 (×9); (I) Palaeolima retifera (Shumard) outer view of right valve, CEGH-UNC 22161 (×4).
(J–L) Nuculopsis sp. (J) internal mold of left valve, CEGH-UNC 22163 (×10); (K) internal mold of articulated valves,
CEGH-UNC 22162 (×14); (L) interior of left valve, CEGH-UNC 22164 (×11). Fossils with the prefix CEGH-UNC are
housed in the Centro de Investigaciones Paleobiológicas (Universidad Nacional de Córdoba) and those with the prefix IPI
are housed in the Fundación Miguel Lillo (Instituto de Paleontología).
142 Cisterna and Sterren

The “Postglacial Levipustula Fauna” tum more stable in which the suspension-feeding and epifaunal
organisms appear more diversified (epibyssate, libero-sessile, and
The first marine invertebrates of the “Postglacial Levipustula pedunculate habits). The relatively more benign climatic condi-
Fauna” have been identified ~20 m above of the uppermost diam- tions would have triggered an abundant food supply. Hence, the
ictic horizon of the Hoyada Verde Formation (Fig. 3). The fossil communities of this association have the highest faunal species
assemblages associated with the postglacial shales are composed richness and biovolume.
of bryozoans, brachiopods, and bivalves (Fig. 5), accompanied The fossiliferous interval that contains the upper brachio-
by less abundant gastropods and crinoids. This fauna is very pod association is characterized by abundant sandstones linked
abundant and highly diversified, and it exhibits compositional to a shallowing-upward trend. This association is dominated
variations throughout the fossiliferous interval studied. by Levipustula levis, a small quasi-infaunal productid, and
The fossil concentrations occur either as thin (1–5 cm) shell the ubiquitous Costuloplica leoncitensis accompanied by the
beds or nests. The bioclasts show poor sorting and a random dis- bivalve Phestia sp. aff. P. bellistriata (Fig. 6C). The values of
tribution. In cross section, the shells are mainly concordant to species richness and biovolume are relatively low and the pres-
slightly oblique to the bedding plane and display predominantly ence of the opportunistic organisms Levipustula levis and Phes-
concave-upward orientations. Low degrees of abrasion and frag- tia sp. aff. P. bellistriata could be related to fluctuating envi-
mentation have been observed in the shells. Similar proportions ronmental conditions (i.e., an unstable substratum and higher
of dorsal/ventral valves in brachiopods and left/right valves in sediment rate).
bivalves are recognized. Some articulated shells are present and The communities of the three associations described above
delicate details of fine sculpture, such as spines in Levipustula would have been developed in a stable marine environment, such
levis, are also preserved. The taphonomic features described sug- as an open shelf with moderate bottom currents (Cisterna, 1999).
gest biogenic fossil concentrations, produced by a gradual accu- Variations in these associations would have been controlled by
mulation of successive benthic colonizations (Sterren, 2002). substrate types and food supply fluctuations during the postgla-
Cisterna (1999) carried out a detailed paleoecologic analysis cial transgression.
that includes the dominant groups (bryozoans, brachiopods and
bivalves) of the postglacial fauna. From this paleoecologic study, DISCUSSION
Simanauskas et al. (2001) recognized three subfaunas (Fig. 6),
based primarily on the changes observed in the brachiopods in A complex glacial history with advances and retreats of gla-
the fossiliferous interval: the lower Costuloplica leoncitensis ciers might have been the main control on the distribution of the
subfauna, an intermediate Kitakamithyris sp. subfauna, and the Levipustula Fauna. Although this fauna is a conspicuous element
upper, Levipustula levis subfauna. in the Calingasta-Uspallata Basin, the glacial influence is more
From our most recent fieldwork we were able to confirm the evident in the faunas present in the Hoyada Verde Formation.
presence of the three brachiopod associations previously identified The identification of the “Intraglacial Levipustula Fauna” and the
by Simanauskas et al. (2001) in the postglacial interval (although “Postglacial Levipustula Fauna” described in this paper consti-
the bryozoans and the brachiopod Costuloplica leoncitensis tutes a new element for understanding the particular relationship
appear to be conspicuous elements throughout the interval). The between the faunal assemblages and the climatic variations due
lower association (Fig. 6A) is characterized by the low number of to the Gondwanan glaciation.
species and low total biovolume. The communities of this asso- The features discussed for the Intraglacial Fauna strongly
ciation are dominated by bivalves of the genus Streblochondria, suggest environmentally stressed conditions probably related to
accompanied, in decreasing order, by bryozoans and Costuloplica low (glacial) temperatures. Although low temperatures seem to
leoncitensis, organisms characterized as epifaunal and suspension be the dominant factor of stress for the faunas discussed herein,
feeding. This association would represent a gradual deepening of other variables such as oxygen and nutrient availability, salinity,
cold waters and a relatively low nutrient availability, which is evi- substrate type, and water depth would have been affected by the
denced by the low biovolume and the low species richness. glacial action.
The intermediate association (Fig. 6B) is dominated by bra- The very abundant and highly diversified “Postglacial Levi-
chiopods or bryozoans and characterized by the highest species pustula fauna” can be considered to represent the record of cli-
richness and the highest values of biovolume. The bivalves are matic amelioration in more stable environmental conditions.
less important in these communities but those of epibyssate habits This fauna exhibits compositional variations that reflect slight
(Streblochondria sanjuanensis Sterren, Streblochondria stappen- changes in the substrate stability and food supply, related to the
becki Reed, and Palaeolima retifera (Shumard)) show a relative moderate bottom currents. Some features of this fauna in the
increase. The brachiopod communities of this association are middle part of the postglacial interval, such as high diversity of
dominated by Costuloplica leoncitensis, accompanied by Kita- the filter-feeder and epifaunal organisms, suggest the presence
kamithyris sp., “Spiriferellina” octoplicata, Beecheria sp., and of maximum flooding conditions during the postglacial trans-
Levipustula levis as a subordinate element. This brachiopod asso- gression. The major sandstone component of the uppermost part
ciation could correspond to the maximum flooding with a substra- of the interval is related to a shallowing-upward trend. The low
A B C

D E F

G H I

J K L M

N O P Q

R S
Figure 5. Postglacial “Levipustula Fauna.” (A–C) Levipustula levis Maxwell. (A) dorsal valve partially deco-
rticated, IPI 3442 (×2.5); (B) internal mold of dorsal valve, IPI 4501 (×3); (C) ventral valve, IPI 3238 (×2).
(D–E) Costuloplica leoncitensis (Harrington); (D) ventral valve, IPI 3221 (×1); (E) dorsal valve partially de-
corticated, IPI 3223 (×2.5). (F–G) Kitakamithyris sp. (F) Internal mold of ventral valve incomplete, IPI 4502
(×1); (G) internal mold of dorsal valve, IPI 4503 (×1). (H–I) “Spiriferellina” octoplicata (Sowerby); (H) ven-
tral valve incomplete, IPI 3232 (×3.5); (I) internal mold of dorsal valve, IPI 3235 (×2.3). (J–K) Beecheria sp.;
(J) internal mold of ventral valve, IPI 4504 (×5); (K) internal mold of dorsal valve, IPI 4505 (×4). (L–M) Stre-
blochondria sanjuanensis Sterren; (L) paratype, composite mold of left valve, CEGH-UNC 19750 (×1.5);
(M) holotype, composite mold of right valve, CEGH-UNC 19748 (×1.5). (N–O) Streblochondria stappen-
becki (Reed); N, internal mold of left valve, CEGH-UNC 19757 (×2.5); (O) internal mold of right valve,
CEGH-UNC 19755 (×1.8). (P–Q) Palaeolima retifera (Shumard); (P) outer view of right valve, CEGH-UNC
22168 (×7); (Q) outer view of right valve, CEGH-UNC 19773 (×2.3). (R–S) Phestia sp. aff. P. bellistriata
(Stevens); (R) interior of left valve, CEGH-UNC 22165 (×4); (S) dorsal view of right valve showing denti-
tions, CEGH-UNC 19717 (×6).
144 Cisterna and Sterren

C a

b
b
a

i i
Figure 6. Hypothetical paleoecological
reconstructions of the three brachiopod
associations recognized in the post-
glacial interval. (A) Lower association
dominated by bivalves of the genus
Streblochondria, accompanied of bryo-
B j a
a
zoans and Costuloplica leoncitensis.
d (B) Middle association dominated by
brachiopod communities, character-
b ized by the highest species richness
f and the highest values of biovolume.
e (C) Upper association dominated by
g k Levipustula levis and Costuloplica
c leoncitensis, accompanied by Phestia
a h
sp. aff. P. bellistriata. a—Costuloplica
leoncitensis (Harrington), b—Levi-
pustula levis Maxwell, c—Beecheria
sp., d—“Spiriferellina” octoplicata
(Sowerby), e—Kitakamithyris sp., f—
Streblochondria sanjuanensis Sterren,
g—Streblochondria stappenbecki Reed,
h—Palaeolima retifera (Shumard), i—
A a a Phestia sp. aff. P. bellistriata, j—Fenes-
tella sp., k—Crinoidea indet.

f g g
a
j
f
k

diversity of the fauna in this part of the postglacial interval as tions drastically change in the postglacial phase and the increase
well as the presence of the opportunistic organisms may repre- of the nutrients accounts for the diversification of the mollusk
sent a relatively unstable substratum. assemblages. This model can be useful for understanding why
The distinctive pattern of the faunal distribution in the Hoyada the “Postglacial Levipustula Fauna” is more diversified than the
Verde Formation, as well as the persistence of some taxa after the “Intraglacial Levipustula Fauna”; it also allows us to understand
glaciation, have been recognized and studied in modern ecosys- the other variations of the postglacial fauna in the different asso-
tems close to glaciers. Faunal variations in the postglacial phases, ciations identified in this contribution.
mainly related to differences in substrate types, water depth, and Studies of the behavior of modern bivalve assemblages from
variations in clastic and organic content sediment rates, have been circumpolar regions suggest that during a glacial event the ice can
suggested by Gordillo and Aitken (2001) for the modern ecosys- negatively affect much of the benthic marine fauna in continental
tems of the Arctic region. These important changes are reflected shelves, but some species are able to survive using some form
in the development of different associations, from the onset of the of refugium. Various strategies for the subfossil mollusks have
deglaciation to the postglacial phase. When the glacier retreats, therefore been proposed, such as displacing into deeper waters,
large volumes of sediment discharged by the meltwater flows moving into subpolar regions, or surviving in nonglaciated pock-
can suppress the vertical circulation and the nutrient generation ets in continental shelves (Crame, 1996). This type of behavior in
in the surface water, limiting the marine primary production and glacial conditions could explain the recurrence of some bivalves
the food supply in ice-proximal benthic habitats. These condi- along the Hoyada Verde section, such as the presence of Phestia
“Levipustula Fauna” in central-western Argentina 145

sp. aff. P. bellistriata and Palaeolima retifera in the Intra- and Pleistocene Glacial Record: Cambridge, UK, Cambridge University
Postglacial Faunas. Press, p. 872–877.
Amos, A.J., and Rolleri, E.O., 1965, El Carbonífero medio en el Valle Calingasta-
The influence of the glaciation in the late Paleozoic marine Uspallata (San Juan-Mendoza): Boletín de Informes Petroleros, v. 368,
biota has been also documented in other basins of western Argen- p. 50–71.
tina such as the western Paganzo Basin. Pazos (2000) recognized Amos, A.J., Baldis, B., and Csaky, A., 1963, La fauna del Carbonífero medio
de la Formación La Capilla y sus relaciones geológicas: Ameghiniana,
a glacial opportunistic ichnofaunal assemblage characterized by v. 3, p. 123–132.
a low diversity and a high degree of burial, which could sug- Amos, A.J., Antelo, B., González, C.R., Mariñelarena, M.P., and Sabattini, N.,
gest stress conditions in a cold-water environment. Buatois et al. 1973, Síntesis sobre el conocimiento bioestratigráfico del Carbonífero-
Pérmico de Argentina: Actas, 5° Congreso Geológico Argentino, v. 3,
(2006, this volume) proposed that high sedimentation rates and p. 3–20.
fluctuations in the water salinity might have characterized the Archangelsky, S., Azcuy, C., González, C.R., and Sabattini, N., 1987, Cor-
depositional conditions during the Gondwanan glaciation. These relación general de las biozonas, in Archangelsky, S., ed., El Sistema
Carbonífero en la República Argentina: Córdoba, Argentina, Academia
authors suggested in addition that the postglacial assemblages Nacional de Ciencias de Córdoba, p. 281–292.
were developed during a transgressive event and the subse- Archangelsky, S., Azcuy, C., Césari, S., González, C.R., Hunickenm, M., Maz-
quent deltaic progradation, which set new conditions charac- zoni, A., and Sabattini, N., 1996, Correlación y edad de las biozonas,
in Archangelsky, S., ed., El Pérmico en la República Argentina y en la
terized by abundant supply of nutrients and oxygenation of the República Oriental del Uruguay: Córdoba, Argentina, Academia Nacional
water column. de Ciencias de Córdoba, p. 203–225.
The conclusions from the present study conducted in the Baldis, B., 1964, Estratigrafía y estructuras del Paleozoico al sur del Arroyo de
Las Cabeceras estancia El Leoncito San Juan: Boletín de Informaciones
Hoyada Verde Formation can be extended to other areas of the Petroleras, v. 368, p. 28–33.
Calingasta-Uspallata Basin. Although the Hoyada Verde section Buatois, L.A., and Limarino, C.O., 2003, El contacto entre las formaciones Hoy-
exhibits the best record of the glacial-postglacial transition and ada Verde y Tres Saltos, Carbonífero de la Cuenca Calingasta-Uspallata:
Su reinterpretación como una superficie de incisión de valle fluvial:
associated faunas, a fossil assemblage compositionally and tapho- Simposio Argentino del Paleozoico Superior, 3°, y Reunión del Proyecto
nomically equivalent to the “Intraglacial Levipustula Fauna” has 471, 2°, La Plata, Argentina, Resúmenes, p. 4.
been recently recognized in the Leoncito Formation (Cisterna Buatois, L.A., Netto, R., Mángano, M.G., and Balistieri, P.L., 2006, Extreme
freshwater release during the late Paleozoic Gondwana deglaciation and
and Sterren, 2008). Ostracods are unusual in the Carboniferous its impact on coastal ecosystems: Geology, v. 34, p. 1021–1024, doi: 10
sequences of the basins of western Argentina, and the presence .1130/G22994A.1.
of the genus Aurykirkbya Sohn in the “Intraglacial Levipustula Buatois, L.A., Netto, R.G., and Mángano, M.G., 2010, this volume, Ichnol-
ogy of late Paleozoic postglacial transgressive deposits in Gondwana:
Fauna” could have significant biostratigraphic implications. This Reconstructing salinity conditions in coastal ecosystems affected by
ostracod genus has been described from equivalent stratigraphic strong meltwater discharge, in López-Gamundí, O.R., and Buatois,
sequences in the Tepuel-Genoa Basin in Argentine Patagonia. L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions
in Gondwana: Geological Society of America Special Paper 468, doi:
Díaz Saravia and Jones (1999) considered that the ostracod fau- 10.1130/2010.2468(07).
nas described from this basin in the lower part of the Pampa de Campbell, K.W.S., 1961, Carboniferous fossils from Kuttung Rocks of New
Tepuel Formation would be early Namurian (lower part of the South Wales: Palaeontology, v. 4, p. 428–474.
Campbell, K.S.W., and McKellar, R.G., 1969, Eastern Australian Carbonifer-
Levipustula levis Zone). In this sense the new “Intraglacial Levi- ous invertebrates: Sequence and affinities, in Campbell, K.S.W., ed., Stra-
pustula Fauna” identified in the Hoyada Verde Formation would tigraphy and Palaeontology: Essays in Honour of Dorothy Hill: Canberra,
have biostratigraphic and paleogeographic implications for intra- Australian National University Press, p. 79–119.
Caputo, M.V., and Crowell, J.C., 1985, Migration of glacial centers across
and even interbasinal correlations. Gondwana during Paleozoic Era: Geological Society of America Bulletin,
v. 96, p. 1020–1036, doi: 10.1130/ 0016–7606(1985)96<1020:MOGCAG
ACKNOWLEDGMENTS >2.0.CO;2.
Chakraborty, A., and Bhattacharya, H.N., 2005, Ichnology of Late Paleo-
zoic (Permo-Carboniferous) glaciomarine environment, Talchir For-
The authors would like to thank CONICET, Consejo Nacio- mation, Saharjuri Basin, India: Ichnos, v. 12, p. 31–45, doi: 10.1080/
nal de Investigaciones Científicas y Técnicas (PIP 6144), and the 10420940590914480.
Cisterna, G.A., 1999, Paleoecología de niveles pelíticos superiores de la For-
Agencia Nacional de Promoción Científica y Tecnológica (PICT mación Hoyada Verde, Carbonífero superior, Precordillera de San Juan,
20752-PICT 32693) from Argentina. We acknowledge N. Emilio Argentina: Ameghiniana, v. 36, p. 259–267.
Vaccari for his help in the photographic work. The reviewers, Cisterna, G.A., and Sterren, A.F., 2003, Variaciones composicionales de la
“Fauna de Levipustula” en la Precordillera Argentina: Simposio Argen-
Lucia Angiolini and Marcello Simões, improved an earlier ver- tino del Paleozoico Superior, 3°, y Reunión del Proyecto 471, 2°, La Plata,
sion of the manuscript. Argentina, Resúmenes, p. 11.
Cisterna, G.A., and Sterren, A.F., 2004, Compositional variations of the “Levi-
pustula fauna” in the Argentine Precordillera and its relationships with
REFERENCES CITED the carboniferous glacial event in the southwestern Gondwanan margin:
International Geological Congress, 32nd, Part 2, p. 961.
Aigner, T., 1985, Storm depositional systems: Dynamic stratigraphy in modern Cisterna, G.A., and Sterren, A.F., 2008, Late Carboniferous Levipustula fauna
and ancient shallow marine sequences: Lecture Notes in Earth Sciences, in the Leoncito Formation, San Juan province, Argentine Precordillera:
v. 3, p. 1–174, doi: 10.1007/bfb0011412. Biostratigraphical and palaeoclimatological implications: Proceedings of
Amos, A.J., and López Gamundi, O.R., 1981, Late Paleozoic diamictites the Royal Society of Victoria, v. 118, p. 137–147.
of the Calingasta-Uspallata and Paganzo basins, San Juan and Men- Collinson, J.W., Isbell, J.L., Elliot, D.H., and Miller, J.M.G., 1994, Permian-
doza provinces, in Hambrey, M., and Harland, W., eds., Earth’s Pre- Triassic Transantarctic basin, in Veevers, J.J., and Powell, C. Mc.A., eds.,
146 Cisterna and Sterren

Permian-Triassic Pangean Basins and Foldbelts along the Panthalassan Lindsay, J.F., 1969, Stratigraphy and structure of the Palaeozoic sediments of
Margin of Gondwanaland: Geological Society of America Memoir 184, the lower Macleay region, north-eastern New South Wales: Journal and
p. 173–222. Proceedings of the Royal Society of New South Wales, v. 102, p. 41–55.
Crame, J.A., 1996, Evolution of high-latitude molluscan faunas, in Taylor, J., López Gamundí, O.R., 1983, Modelo de sedimentación glacimarina para la
ed., Origin and Evolutionary Radiation of the Mollusca: Oxford, UK, Formación Hoyada Verde, Paleozoico superior de la provincia de San
Oxford University Press, p. 119–131. Juan: Revista de la Asociación Geológica Argentina, v. 38, p. 60–72.
Crowell, J.C., 1983, Ice ages recorded on Gondwana continents: Geological López-Gamundí, O.R., 1989, Postglacial transgressions in late paleozoic basins
Society of South Africa Transactions, v. 86, p. 237–262. of western Argentina: A record of glacioeustatic sea level rise: Palaeogeog-
Díaz Saravia, P., and Jones, P.J., 1999, New Carboniferous (Namurian) gla- raphy, Palaeoclimatology, Palaeoecology, v. 71, p. 257–270, doi: 10.1016/
ciomarine ostracods from Patagonia, Argentina: Journal of Micropalae- 0031-0182(89)90054-0.
ontology, v. 18, p. 97–109. López-Gamundí, O.R., 1990, Mecanismos de formación, registro sedimentario
Freytes, E., 1971, Informe geológico preliminar sobre la Sierra de Tepuel y jerarquía estratigráfica de las transgresiones postglaciales en secuencias
(Departamentos de Languiñeo y Tehuelches, provincia de Chubut): neopaleozoicas de Argentina: Anales de la Academia de Ciencias Exac-
Yacimientos Petrolíferos Fiscales (Informe inédito). tas: Físicas y Naturales, v. 42, p. 165–182.
Fürsich, F.T., and Heinberg, K., 1983, Sedimentology, biostratinomy and palae- López-Gamundí, O.R., 1997, Glacial-postglacial transition in the late Paleozoic
oecology of an Upper Jurassic offshore sand bar complex: Bulletin of the basins of Southern South America, in Martini, I.P., ed., Late Glacial and
Geological Society of Denmark, v. 32, p. 67–95. Postglacial Environmental Changes—Quaternary, Carboniferous-Permian,
González, C.R., 1972, La Formación Las Salinas, Paleozoico Superior de Chubut and Proterozoic: Oxford, UK, Oxford University Press, p. 147–168.
(Argentina). Parte II. Bivalvia: Taxonomía y paleoecología: Revista de la López-Gamundí, O.R., 2010, this volume, Transgressions related to the demise
Asociación Geológica Argentina, v. 27, p. 188–213. of the late Paleozoic Ice Age: Their sequence stratigraphic context, in
González, C.R., 1980, Sobre la presencia de “glendonita” en el Paleozoico López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial
Superior de Patagonia: Revista de la Asociación Geológica Argentina, Events and Postglacial Transgressions in Gondwana: Geological Society
v. 35, p. 417–420. of America Special Paper 468, doi: 10.1130/2010.2468(01).
González, C.R., 1981, El Paleozoico Superior marino de la República Argen- López-Gamundí, O.R., and Espejo, I., 1993, Correlation of a paleoclimatic
tina, Bioestratigrafía y Paleoclimatología: Ameghiniana, v. 18, p. 51–55. mega-event: The Carboniferous Glaciation in Argentina: Compte Rendu,
González, C.R., 1990, Development of the Late Paleozoic glaciations of the 12e Congrès International Stratigraphie et Géologie du Carbonifère et
South American Gondwana in Western Argentina: Palaeogeography, Permien, v. 1, p. 313–324.
Palaeoclimatology, Palaeoecology, v. 79, p. 275–287, doi: 10.1016/0031 López-Gamundí, O.R., and Martínez, M., 2000, Evidence of glacial abrasion in
-0182(90)90022-Y. the Calingasta-Uspallata and western Paganzo basins, mid-Carboniferous
González, C.R., 1993, Late Paleozoic faunal succession in Argentina: Compte of western Argentina: Palaeogeography, Palaeoclimatology, Palaeoecol-
Rendu, 12e Congrès International Stratigraphie et Géologie du Carboni- ogy, v. 159, p. 145–165, doi: 10.1016/S0031-0182(00)00044-4.
fère et Permien, v. 1, p. 537–550. López-Gamundí, O.R., and Rosello, E., 1995, Pavimento glacial en la For-
González, C.R., 2002, Bivalves from Carboniferous glacial deposits of western mación Leoncito (Carbonífero), Precordillera occidental, San Juan:
Argentina: Paläontologische Zeitschirft, v. 76, p. 127–148. Revista de la Asociación Geológica Argentina, v. 50, p. 1–4.
González, C.R., and Taboada, A.C., 1987, Nueva localidad fosilífera del Car- Mángano, M.G., Buatois, L.A., Limarino, C.O., Tripaldi, A., and Caselli, A.,
bónico marino de la provincia de San Juan: Actas, 10° Congreso Geológico 2003, El icnogénero Psammichnites Torell, 1870 en la Formación Hoyada
Argentino, v. 3, p. 103–105. Verde, Carbonífero Superior de la cuenca Calingasta-Uspallata: Ameghin-
Gordillo, S., and Aitken, A.E., 2001, Postglacial succession and palaeo- iana, v. 40, p. 601–608.
ecology of Late Quaternary macrofaunal assemblages from the central Maxwell, W.G., 1951, Upper Devonian and Middle Carboniferous brachiopods
Canadian Arctic Archipelago: Boreas, v. 30, p. 61–72, doi: 10.1080/ of Queensland: University of Queensland Papers, Department of Geol-
030094801300062329. ogy, v. 3, p. 1–27.
Hambrey, M.J., and Harland, W.B., 1981, Earth’s Pre-Pleistocene Glacial Maxwell, W.G., 1964, The geology of the Yarrol region, Part 1, Biostratigraphy:
Record: Cambridge, UK, Cambridge University Press, 1004 p. University of Queensland Papers, Department of Geology, v. 5, p. 1–79.
Jones, P.J., Campbell, K.S.W., and Roberts, J., 1973, Correlation chart for McLachlan, I.R., Tsikos, H., and Cairncross, C., 2001, Glendonites (pseudo-
the Carboniferous System of Australia: Bulletin, Bureau of Mineral morphs after ikaite) in late Carboniferous Marine Dwyka beds in South-
Resources, Geology and Geophysics, Australia, v. 156A, p. 1–40. ern Africa: South African Journal of Geology, v. 104, p. 265–272, doi:
Kammer, T.W., Brett, C.E., Boardman, D.R., and Mapes, R.H., 1986, Ecologic 10.2113/1040265.
stability of the dysaerobic biofacies during the Late Paleozoic: Lethaia, McKellar, R.G., 1965, An Upper Carboniferous brachiopod fauna from the
v. 19, p. 109–121, doi: 10.1111/j.1502-3931.1986.tb00720.x. Monto district, Queensland: Publications of the Geological Survey of
Keidel, J., and Harrington, H.J., 1938, On the discovery of Lower Carbonifer- Queensland, v. 328, p. 1–15.
ous tillites in the Precordillera of San Juan: Geological Magazine, v. 75, Mésigos, M., 1953, El Paleozoico Superior de Barreal y su continuación austral,
p. 103–129, doi: 10.1017/S0016756800091433. Sierra de Barreal (Prov. de San Juan): Revista de la Asociación Geológica
Kidwell, S.M., and Baumiller, T.M., 1990, Experimental disintegration of regu- Argentina, v. 8, p. 65–109.
lar echinoids: Roles of temperature, oxygen and decay thresholds: Paleo- Pazos, P.J., 2000, Trace fossils and facies in glacial to postglacial deposits from
biology, v. 16, p. 247–271. the Paganzo basin (Late Carboniferous), central Precordillera, Argentina:
Kidwell, S.M., and Bosence, D.W., 1991, Taphonomy and time-averaging of Ameghiniana, v. 37, p. 23–38.
marine shelly faunas, in Allison, P.A., and Briggs, B., eds., Data Locked Peterson, C.H., 1985, Patterns of lagoonal bivalve mortality and their paleonto-
in the Fossil Record: New York, Plenum Press, Topics in Geobiology, v. 9, logical significance: Paleobiology, v. 11, p. 139–153.
p. 115–209. Ramos, V., and Palma, R., 1996, Tectónica, in El Sistema Pérmico en la
Klets, A.G., 1983, A new carboniferous productid genus: Paleontological Jour- República Argentina y en la República Oriental del Uruguay: Córdoba,
nal, v. 17, p. 70–75. Argentina, Academia Nacional de Ciencias de Córdoba, p. 239–254.
Lebold, J.G., and Kammer, T.W., 2006, Gradient analysis of faunal distribu- Ramos, V., Jordan, T., Allmendinger, R., Mpodozis, C., Kay, S., Cortes, J., and
tions associated with rapid transgression and low accommodation space Palma, M., 1986, Paleozoic terranes of the central Argentine-Chilean
in a Late Pennsylvanian marine embayment: Biofacies of the Ames Mem- Andes: Tectonics, v. 5, p. 855–880, doi: 10.1029/TC005i006p00855.
ber (Glenshaw Formation, Conemaugh Group) in the northern Appala- Reed, F.R.C., 1927, Upper Carboniferous fossils from the Argentina, in Du
chian Basin, USA): Palaeogeography, Palaeoclimatology, Palaeoecology, Toit, A.L., ed., A Geological Comparison of South America with South
v. 231, p. 291–314, doi: 10.1016/j.palaeo.2005.08.005. Africa: Publications of the Carnegie Institution of Washington, v. 381,
Lech, R.R., 1989, Algunos braquiópodos de la Formación Leoncito, Carboníf- p. 129–149.
ero inferior de la provincia de San Juan, Argentina: Actas, 4° Congreso Roberts, J., 1976, Carboniferous chonetacean and productacean brachiopods
Argentino de Paleontología-Bioestratigrafía, v. 4, p. 5–10. from eastern Australia: Palaeontology, v. 19, p. 17–77.
“Levipustula Fauna” in central-western Argentina 147

Roberts, J., Hunt, J.W., and Thompson, D.M., 1976, Late Carboniferous marine Sterren, A.F., 2000, Moluscos bivalvos en la Formación Río del Peñón, Car-
invertebrate zones of eastern Australia: Alcheringa, v. 1, p. 197–225, doi: bonífero tardío-Pérmico temprano, provincia de La Rioja: Ameghiniana,
10.1080/03115517608619071. v. 37, p. 421–438.
Roberts, J., Jones, P.J., and Jenkins, T.B.H., 1993, Revised correlations for Sterren, A.F., 2002, Paleoecología, tafonomía y taxonomía de los moluscos
Carboniferous marine invertebrate zones of eastern Australia: Alcheringa, bivalvos del Carbonífero-Pérmico en las cuencas de Río Blanco y Calin-
v. 17, p. 353–376, doi: 10.1080/03115519308619598. gasta-Uspallata. [Ph.D. thesis]: Córdoba, Argentina, Universidad Nacio-
Roberts, J., Claoué-Long, J.C., Jones, P.J., and Foster, C.B., 1995, SHRIMP nal de Córdoba, p. 1–203.
zircon age control of Gondwanan sequences in Late Carboniferous Sterren, A.F., 2003, Bivalvos carboníferos de la sierra de Barreal, cuenca
and early Permian Australia, in Dunay, R.E., and Hailwood, E.A., eds., de Calingasta-Uspallata, provincia de San Juan: Ameghiniana, v. 40,
Non-Biostratigraphical Methods of Dating and Correlation: Geological p. 469–481.
Society [London] Special Publication 89, p. 145–174, doi: 10.1144/GSL Sterren, A.F., 2005, Bivalvos carboníferos de la Formación La Capilla en el
.SP.1995.089.01.08. área de Las Cambachas, provincia de San Juan: Ameghiniana, v. 42,
Rocha-Campos, A.C., Carvalho, R.G., and Amos, A.J., 1977, A Carboniferous p. 209–219.
(Gondwana) fauna from Subandean Bolivia: Revista Brasileira de Geo- Sterren, A.F., and Cisterna, G.A., 2006, La fauna de Levipustula en la For-
ciências, v. 7, p. 287–303. mación Hoyada Verde: Control paleoecológico versus resolución
Sabattini, N., 1972, Los Fenestellidae Acanthocladiidae y Rhabdomesidae bioestratigráfica: Congreso Argentino de Paleontología y Bioestratigrafía,
(Bryozoa, Cryptostomata) del Paleozoico superior de San Juan y Chubut, 9°, Córdoba, Argentina, Resúmenes, p. 192.
Argentina: Revista del Museo de la Plata, n.s. 6, Paleontología, v. 42, Swainson, I.P., and Hammond, R.P., 2001, Ikaite, CaCO36H2O: Cold comfort
p. 255–377. for glendonites as paleothermometers: The American Mineralogist, v. 86,
Sabattini, N., 1980, Gastrópodos marinos carbónicos y pérmicos de la Sierra de p. 1530–1533.
Barreal (Provincia de San Juan): Ameghiniana, v. 17, p. 109–119. Taboada, A.C., 1997, Bioestratigrafía del Carbonífero marino del valle de
Scotese, C.R., and Barret, S.F., 1990, Gondwana’s movement over the South Calingasta-Uspallata, provincias de San Juan y Mendoza: Ameghiniana,
Pole during the Palaeozoic: Evidence from lithological indicators of cli- v. 34, p. 215–246.
mate, in McKerrow, W.S., and Scotese, C.R., eds., Palaeozoic Palaeo- Taboada, A.C., 2006, Levipustula Maxwell, 1951 y Lanipustula Klets, 1983
geography and Biogeography: Geological Society [London] Memoir 12, (Brachiopoda, Levipustulini) en Argentina: Revisión preliminar: Con-
p. 75–85. greso Argentino de Paleontología y Bioestratigrafía, 9°, Córdoba, Argen-
Scotese, C.R., and McKerrow, W.S., 1990, Revised World maps and introduc- tina, Resúmenes, p. 193.
tion, in McKerrow, W.S., and Scotese, C.R., eds., Palaeozoic Palaeo- Taboada, A.C., 2008, First record of the Late Paleozoic brachiopod Verchojania
geography and Biogeography: Geological Society [London] Memoir 12, in Patagonia, Argentina: Proceedings of the Royal Society of Victoria,
p. 1–21. v. 120, no. 1, p. 305–319.
Sessarego, H., and Amos, A.J., 1987, Diamictitas en la Formación La Capilla Taboada, A.C., and Carrizo, H.A., 1992, La Formación Yalguaraz, Paleozoico
(Carbonífero), zona de Las Juntas de los ríos Castaño y Los Patos, provin- superior de la Cordillera Frontal Argentina. Bioestratigrafía, paleoambi-
cia de San Juan, Argentina: Annual Meeting IUGS-UNESCO Project 211, entes y paleogeografía: Acta Geológica Lilloana, v. 17, p. 115–128.
Late Paleozoic of South America, Abstracts, p. 85. Taboada, A.C., and Cisterna, G., 1996, Elythinae (Brachiopoda) del Paleozoico
Simanauskas, T., 1996, Una nueva especie de Lanipustula (Productoidea, Bra- Superior de Argentina: Ameghiniana, v. 33, p. 83–94.
chiopoda) del Paleozoico Superior de Argentina: Ameghiniana, v. 33, Taboada, A.C., and Sabattini, N., 1987, Nuevos Eotomariidae (Gastropoda) del
p. 301–305. Paleozoico Superior de Argentina: Ameghiniana, v. 24, p. 175–180.
Simanauskas, T., and Cisterna, G.A., 2000, A palaeo-opportunistic brachiopod Thomas, S.G., Fielding, C.R., and Frank, T.D., 2007, Lithostratigraphy of the
from the Early Permian of Argentina: Alcheringa, v. 24, p. 45–53, doi: late Early Permian (Kungurian) Wandrawandian Siltstone, New South
10.1080/03115510008619522. Wales: Record of glaciation? Australian Journal of Earth Sciences, v. 54,
Simanauskas, T., and Cisterna, G.A., 2001, Los braquiópodos articulados de la p. 1057–1071.
Formación El Paso, Paleozoico Tardío, Precordillera Argentina: Revista Trujillo Ikeda, H., 1989, Nuevo hallazgo de fósiles de la Formación Taiguari en
Española de Paleontología, v. 16, p. 209–222. la Serranía Caipipendi, Santa Cruz, Bolivia: Revista Técnica de Yacimien-
Simanauskas, T., and Sabattini, N., 1997, Bioestratigrafía del Paleozoico Supe- tos Petrolíferos Fiscales Bolivianos, v. 10, p. 7–11.
rior marino de la Cuenca Tepuel-Genoa, Chubut, Argentina: Ameghini- Vallecillo, G., and Bercowski, F., 1998, Litofacies y paleoambientes de la For-
ana, v. 34, p. 49–60. mación La Capilla (Carbonífero), Calingasta, provincia San Juan, Argen-
Simanauskas, T., Cisterna, G.A., and Sterren, A.F., 2001, Evolución de las fau- tina: Actas, 10° Congreso Latinoamericano de Geología y 6° Congreso
nas bentónicas marinas de la Formación Hoyada Verde, Carbonífero tar- Nacional de Geología Económica, v. 1, p. 243–248.
dío de la sierra de Barreal, San Juan: Simposio Argentino del Paleozoico
Superior, 2°, Trelew, Argentina, Resúmenes, p. 26. MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

Ichnology of late Paleozoic postglacial transgressive deposits


in Gondwana: Reconstructing salinity conditions in coastal
ecosystems affected by strong meltwater discharge

Luis A. Buatois
Department of Geological Sciences, University of Saskatchewan, 114 Science Place, Saskatoon, SK S7N 5E2, Canada

Renata G. Netto
Unisinos, Programa de Pós-graduação em Geologia, Av. Unisinos 950, 93022-000 São Leopoldo RS, Brazil

M. Gabriela Mángano
Department of Geological Sciences, University of Saskatchewan, 114 Science Place, Saskatoon, SK S7N 5E2, Canada

ABSTRACT

Late Paleozoic ichnofaunas from eight different Gondwanic basins (Paganzo,


San Rafael, Tarija, Paraná, Karoo, Falkland, Transantarctic, and Sydney) provide
valuable evidence for reconstructing the environmental conditions of postglacial
transgressions. The depositional environment of most of these transgressive fine-
grained deposits historically has been controversial, with interpretations ranging
from freshwater lacustrine to brackish-water estuarine, and even normal-salinity,
open-marine platforms. Although the various units differ in the degree of marine con-
nection, the common theme in all is the presence of freshwater ichnofaunas in direct
association with glacially influenced coasts affected by strong discharges of meltwa-
ter. Ichnofaunas are typically dominated by nonspecialized grazing trails (Mermia,
Helminthopsis, Helminthoidichnites), simple feeding traces (Treptichnus), arthropod
trackways (Diplichnites, Umfolozia), and fish trails (Undichna), representing exam-
ples of the Mermia and, to a lesser extent, the Scoyenia ichnofacies. A complex paleo-
geography of fjords and deep, large coastal lakes is suggested. Freshwater conditions
were prevalent during most of the time because these areas were affected by a strong
discharge of fresh water due to melting of the ice masses during deglaciation. The sim-
ple dichotomy between marine and nonmarine settings is misleading because these
peculiar assemblages should first be understood in terms of their paleoecologic signifi-
cance, and subsequently placed within a larger paleoenvironmental context. Laterally
persistent, albeit diachronous, peri-Gondwanan ichnofaunas characterize melting of
the late Paleozoic ice caps. Temporal recurrence of these ichnofaunas through the
Late Carboniferous–Middle Permian indicates a common response of benthic faunas
under similar ecological conditions during deglaciation events.

Buatois, L.A., Netto, R.G., and Mángano, M.G., 2010, Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana: Reconstructing salinity condi-
tions in coastal ecosystems affected by strong meltwater discharge, in López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial Events and Post-
glacial Transgressions in Gondwana: Geological Society of America Special Paper 468, p. 149–173, doi: 10.1130/2010.2468(07). For permission to copy, contact
[email protected]. ©2010 The Geological Society of America. All rights reserved.

149
150 Buatois et al.

INTRODUCTION et al., 2003; Isbell et al., 2003a, 2003b; Trosdtorf et al., 2005).
Deglaciation sequences are diachronous, reflecting the different
Upper Paleozoic rocks in the Southern Hemisphere con- ages of the Gondwana glaciation.
tain an extensive record of the Gondwana glaciations and the Although Gondwana glacial events and associated deglacia-
associated postglacial transgressions (see López-Gamundí, this tion episodes may have strongly affected benthic communities
volume). The third Gondwanan Ice House Age spanned at least inhabiting terrestrial, freshwater, and coastal ecosystems, sur-
90 m.y. (Caputo and Crowell, 1985), starting in western South prisingly little is known about late Paleozoic glacial ecosystems.
America during the Tournaisian (Mississippian) and ending at the The global impact of the Gondwana glaciation in marine eco-
base of the Guadalupian (Middle Permian) in Australia. It charac- systems was recently evaluated by Stanley and Powell (2003)
terizes the longest continuous glaciation during the Phanerozoic and Powell (2005), based on J.J. Sepkoski’s (2002) database and
(Eyles, 1993). Upper Paleozoic glacial diamictites and postgla- an expanded database of brachiopod occurrences, respectively.
cial fine-grained deposits occur in Australia, Southeast Asia, These authors noted that rates of origination and extinction for
India, Pakistan, Arabian Peninsula, North Africa, Namibia, South marine organisms dropped to low levels during the glaciation.
Africa, southern South America, and Antarctica (Fig. 1) (e.g., However, the direct impact of glacial and deglaciation episodes
Veevers and Powell, 1987; López Gamundí, 1989, 1997; Eyles on late Paleozoic Gondwana communities has remained essen-
and Young, 1994; Visser, 1997; Limarino et al., 2002; Angiolini tially underexplored. In part, lack of studies may be due to the

Arabian Peninsula
Africa

India

h
out ica
S er

A
us
Am 4 Antarctica

tr
al
ia
1
5
6
7
2
3 Falkland
Pan Islands
8 rg in
tha
lass n ma
an s sa
ma
rgin New Zealand th ala
Polar Circle n
(290 Ma)
Pa
1,000 km
Discussed Gondwanan basins with Upper Paleozoic glaciogenic successions (and postglacial freshwater ichnofaunas):

1 - Tarija Basin 5 - Karoo Basin (including Kalahari Basin)


2 - Paganzo Basin 6 - Falkland Basin
3 - San Rafael Basin 7 - Transantarctic Basin
4 - Paraná Basin (including Chaco-Paraná 8 - Sydney Basin
and North Uruguayan Basins)

Other Gondwanan glacial basins with Upper Paleozoic glaciogenic successions

Figure 1. Location map of late Paleozoic Gondwanan basins (modified from Isbell et al., 2003a).
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 151

low preservation potential of body fossils in areas adjacent to gla- ian (López Gamundí, 1989; Limarino et al., 2002). Outcrops of
ciated margins, such as modern fjords (e.g., Aitken, 1990). Most the Devonian–Early Carboniferous glaciation are well exposed
paleontologic studies dealing with late Paleozoic faunas focus in the Titicaca Basin of Bolivia (e.g., Díaz Martínez et al.,
on the extremely rich record of the Northern Hemisphere and, in 1993), and the Amazonas and Parnaíba Basins of Brazil
addition, those on Gondwanic faunas have for the most part dealt (Rocha-Campos, 1981a, 1981b). The early Late Carboniferous
with biostratigraphic rather than paleoecologic aspects. glacial event is well represented in the Andean basins of Argen-
Interestingly, upper Paleozoic rocks of Gondwana com- tina, including the Tarija, Calingasta-Uspallata, Paganzo, and
monly contain distinctive trace-fossil assemblages that are com- San Rafael Basins (e.g., López Gamundí, 1997; del Papa and
monly preserved in postglacial deposits and, more rarely, in Martínez, 2001; Limarino et al., 2002; Marenssi et al., 2005),
fine-grained intervals interbedded with or underlying glacial as well as the Sydney Basin of eastern Australia (e.g., Isbell
diamictites. Therefore, ichnologic evidence provides a proxy for et al., 2003a; Birgenheier et al., 2009). The Late Carboniferous–
use in reconstructing ecosystems associated to the Gondwana gla- Early Permian glaciation is the most widespread and has been
ciation in rocks that are commonly devoid of body fossils. These extensively documented in the Paraná Basin of Brazil, Paraguay,
peri-Gondwanic ichnofaunas are characterized by the dominance and Uruguay (e.g., Eyles et al., 1993; França, 1994; Vesely and
of simple grazing trails and arthropod trackways (Buatois et al., Assine, 2006), the Karoo Basin of South Africa (e.g., Visser,
2006). Their paleoecological and paleoenvironmental significance 1990, 1997), the Kalahari Basin of Namibia, Botswana, and
has largely been debated, with interpretations ranging from fresh- South Africa (e.g., Visser, 1997; Key et al., 1998), the Falkland
water lacustrine to brackish-water deltaic or estuarine, and even Basin (Trewin et al., 2002), the Transantarctic Basin of Ant-
fully marine (e.g., Savage, 1970, 1971; Anderson, 1970, 1975a, arctica (e.g., Isbell et al., 1997), the Gondwana Master Basins
1975b, 1976, 1981; Guerra-Sommer et al., 1984; Fernandes et al., of India (e.g., Wopfner and Casshyap, 1997), and the Sydney,
1987; Dias-Fabrício and Guerra-Sommer, 1989; Bercowski et al., Officer, Carnavon and Canning Basins of Australia (e.g., Veevers
1990; Buatois and Mángano, 1992, 1993, 1995a, 1995b, 2003; and Powell, 1987; O’Brien et al., 1998; Eyles and Eyles, 2000;
Miller and Collinson, 1994; Peralta and Milana, 1999; Pazos, Eyles et al., 2003). Our study is focused on the latter two of these
2000, 2002a, 2002b; Trewin, 2000; Isbell et al., 2001; Nogueira three glacial events because no detailed ichnologic information
and Netto, 2001a, 2001b; Balistieri et al., 2002, 2003a; Trewin is available for the Devonian–Early Carboniferous glacial event.
et al., 2002; Buatois and del Papa, 2003; Gandini et al., 2007; Discussion of the ichnofaunas of the San Rafael, Karoo, Falk-
Pazos et al., 2007). In an attempt to reconcile sedimentologic, land, Transantarctic, and Sydney Basins is based on literature
ichnologic, and paleontologic data, it has recently been suggested review and/or study of trace-fossil specimens.
that freshwater conditions—resulting from a strong discharge of
freshwater due to melting of the ice masses—were prevalent in The Paganzo Basin (Western Argentina)
coastal areas during deglaciation (Buatois et al., 2006). Within
this environmental scenario, glacial melting due to climatic ame- The Upper Carboniferous–Upper Permian Paganzo Basin
lioration led to the formation of freshwater water bodies that were covers ~150,000 km2 of western Argentina (Fig. 1), and hosts
physically connected with the open sea. In this setting, the simple sedimentary successions up to 3000 m thick. It is currently con-
dichotomy between marine and nonmarine settings is mislead- sidered a foreland basin related to subduction of the Pacific plate
ing. To complicate things further, some of the discussions sur- beneath the western continental margin of Gondwana, which
rounding the paleosalinity significance of these ichnofaunas are probably evolved to a rift system during the Permian (Ramos,
poorly constrained stratigraphically because trace-fossil assem- 1988). The Protoprecordillera was a north-south-trending topo-
blages coming from different stratigraphic levels are analyzed as graphic high, albeit discontinuous, that separates mostly marine
a whole and paleoenvironmental interpretations are commonly deposits of the Calingasta-Uspallata and Río Blanco Basins on
uncritically extrapolated from one interval or locality to another. the west from the mostly continental deposits of the Paganzo
The aim of this paper is to further support the strong meltwa- Basin on the east. The magmatic arc was located farther to the
ter discharge hypothesis by evaluating the paleoecologic and west, in present-day Chile. The Paganzo Basin was limited to
paleoenvironmental significance of late Paleozoic ichnofaunas the east and south by the Pampean and Pie de Palo topographic
from eight different sedimentary basins along the Panthalassan highs, respectively, whereas the Puna high represents its northern
margin of Gondwana. boundary. Sedimentation in the Paganzo Basin occurred in sub-
basins separated by internal basement highs. Overall, two major
GLACIAL EVENTS AND POSTGLACIAL areas can be distinguished within the Paganzo Basin: an eastern
TRANSGRESSIONS: SEDIMENTOLOGIC zone dominated by continental environments, and a western one
AND ICHNOLOGIC SIGNATURES with increased marine influence (Limarino et al., 2002).
Strata of the Paganzo Basin constitute the upper Paleozoic
Three main glacial events are recognized in Gondwana, Paganzo Group (Fig. 2), although different nomenclatures have
depending on paleolatitude: Devonian–Early Carboniferous, been established in various areas of the basin (Azcuy and Morelli,
early Late Carboniferous, and Late Carboniferous–Early Perm- 1970). Glacial diamictites and postglacial fine-grained deposits
152 Buatois et al.

Depauperate Cruziana -Skolithos


Ichnofacies (Brackish Water)
SUBAQUEOUS
Scoyenia Ichnofacies
OUTWASH
(Terrestrial to Freshwater)
TS
DISTRIBUTARY PLAIN
FJORD EMBAYMENT

Rio do Sul Fm
Depauperate Cruziana -Skolithos
DELTA FRONT Ichnofacies (Brackish Water)
TS
DELTA FRONT

PRODELTA
Mermia Ichnofacies
(Freshwater ) FJORD EMBAYMENT
Depauperate Cruziana -Skolithos
Tarija Fm

LAKE OR FJORD Ichnofacies (Brackish Water)


Guandacol Fm

MFS
EMBAYMENT
TS Glossifungites Ichnofacies
Mermia Ichnofacies TS (Marine Firmground)
(Freshwater)
Depauperate Cruziana -Skolithos
PRODELTA Ichnofacies (Brackish Water)
SUBAQUEOUS
OUTWASH

Mafra Fm
DELTA FRONT
MFS

FJORD EMBAYMENT
Depauperate Cruziana -Skolithos
Scoyenia Ichnofacies Ichnofacies (Brackish Water)
Depauperate Cruziana (Terrestrial to Freshwater)
50 m

Ichnofacies (Brackish Water)


LAKE OR FJORD PRODELTA
TS EMBAYMENT Mermia-Scoyenia Ichnofacies
Itacuami Fm

TS (Freshwater + Terrestrial to Freshwater)

SUBAQUEOUS DELTA FRONT FJORD EMBAYMENT


50 m
50 m

OUTWASH TS
Scoyenia Ichnofacies
(Terrestrial to Freshwater)
PRODELTA
CS Cg CS Cg CS Cg
Vf FMC Vc Vf FMC Vc Vf FMC Vc
Sand Sand Sand

Paganzo Basin Tarija Basin Paraná Basin


Key
Trough cross-stratification Wave ripple cross-lamination Climbing ripple cross-lamination

Dropstone Current ripple cross-lamination Hummocky cross-stratification TS Transgressive surface

Diamictite Planar cross-stratification Marine Fauna MFS Maximum Flooding surface

Figure 2. Simplified stratigraphic logs of the Paganzo, Tarija, and Paraná Basins showing distribution of trace-fossil associations (after
Buatois et al., 2006, and references therein).

occur in the lower Upper Carboniferous (Namurian–Westphalian) 2005). Deglaciation led to the establishment of large freshwa-
Guandacol Formation of the western area, and in the Agua Colo- ter bodies to the east, and was accompanied by a marine incur-
rada Formation of the eastern area (Limarino et al., 2002). Far- sion from the west that flooded valleys giving place to a series
ther to the east, the coeval Malanzán Formation records the same of fjords along the coastal area (Limarino et al., 2002; Kneller
depositional cycle (Andreis et al., 1986; Buatois and Mángano, et al., 2004). These water bodies were filled during the subse-
1995a). Sedimentation began with coarse-grained alluvial fan quent regression mostly by deltaic progradation that eventually
deposits, braided fluvial deposits, and tillite that gave way rapidly led to the establishment of fluvial systems across the area, rep-
to transgressive mudstone and sandstone (Limarino and Césari, resented in the overlying Upper Carboniferous–Lower Permian
1988; Buatois and Mángano, 1995b; López Gamundí and Mar- (Stephanian–Asselian) Tupe Formation and coeval units (Ottone
tínez, 2000; Limarino et al., 2002; Pazos, 2002a; Marenssi et al., and Azcuy, 1986; Limarino, 1988; Desjardins et al., 2009).
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 153

The ichnofauna of the Guandacol Formation has been ana- restrially derived palynomorphs (Buatois and Mángano, 2003).
lyzed in a number of studies, mostly in Cuesta de Huaco (Pazos, Body fossils are lacking in these deposits. The most likely envi-
2000) and Huerta de Huachi (Buatois and Mángano, 2003), ronmental scenario for the Guandacol Formation is that of fjords
San Juan Province. In Cuesta de Huaco, it has been subdivided filled by prograding deltaic deposits (Limarino et al., 2002).
into two main trace-fossil assemblages that yield insights into the In the eastern region, the Mermia assemblage has been
ecological conditions during glacial retreat: the Didymaulichnus recorded in the coeval Agua Colorada Formation at Sierra de
and the Mermia trace-fossil assemblages. Deposits bearing these Narváez, Catamarca Province (Buatois and Mángano, 1993)
two assemblages are separated by nonbioturbated black shale rep- (Fig. 4). High-diversity suites occur in parallel-laminated mud-
resenting maximum flooding. The Didymaulichnus assemblage stone formed by suspension fallout and low-density turbidity cur-
consists of monospecific suites of bilobate trails that historically rents, and in delta-fed, very fine grained sandstone and siltstone
have been assigned to Didymaulichnus lyelli, although the taxo- deposited from underflow currents. This ichnofauna includes
nomic affinities of these bilobate trace fossils are currently under Circulichnis montanus, Cochlichnus anguineus, Gordia marina,
revision. This assemblage occurs in a wide variety of lithologies, Gordia indianaensis, Helminthoidichnites tenuis, Helminthopsis
including mudstone, very fine- to fine-grained sandstone with tenuis, Mermia carickensis, Orchesteropus atavus, Rusophycus
current and combined-flow ripples, or granule conglomerate and isp., Treptichnus pollardi, Undichna britannica, and U. inso-
very coarse-grained sandstone deposited from debris flows (Bua- lentia, among other forms. A low-diversity suite, consisting of
tois et al., 2006). Dropstones are abundant in this interval. This Gordia marina and Mermia carickensis, occurs at the top of tem-
assemblage is present in the lowermost strata of the transgres- pestite and turbidite fine- to very fine grained sandstone, record-
sive interval, and has been found only in the western area. Low- ing post-event colonization after major breaks in environmental
diversity suites of linguliformean brachiopods (cf. Oehlertella conditions (Buatois and Mángano, 1995b). The palynologic
sp.) occur in one bed (Martínez, 1993). The ichnogenus Didy- assemblage consists only of terrestrially derived forms, and no
maulichnus occurs in both continental (e.g., Miller, 1986), and acritarchs have been reported to date (Vergel et al., 1993). Inte-
marine (e.g., Buatois and López Angriman, 1992) deposits. In grated sedimentologic and ichnologic data together with the lack
the present context, the Didymaulichnus assemblage most likely of marine indicators suggest deposition in large and deep lakes
represents brackish-water conditions in a stressful environment. (Buatois and Mángano, 1994, 1995b), although deposition in the
Late Paleozoic brackish-water ichnofaunas are characterized by inner zone of fjords affected by strong meltwater discharge can-
low-diversity (commonly monospecific) suites of very simple not be disregarded.
forms (Mángano and Buatois, 2004). Subsequently, a similar assemblage, although poorly pre-
The Mermia trace-fossil assemblage occurs in both Cuesta served and less diverse, has been found in the Agua Colorada
de Huaco and Huerta de Huachi (Figs. 3A–3F). It is relatively Formation at Bajo El Manzano (La Rioja Province). Trace fossils
diverse and is dominated by nonspecialized grazing trails in this locality occur in parallel-laminated siltstone immediately
(Mermia carickensis, Gordia marina, Helminthopsis tenuis, above a dropstone-bearing siltstone interval. These fine-grained
Helminthoidichnites tenuis), simple feeding traces (Treptichnus deposits are sandwiched between pebble conglomerate and very
pollardi, Circulichnis montanus), arthropod trackways (Diplich- coarse to medium-grained sandstone that formed in low-sinuosity
nites gouldi, Umfolozia isp., Maculichna carboniferus, Orches- fluvial systems, and sandstone, siltstone, and conglomerate of
teropus atavus), and fish trails (Undichna insolentia, Undichna fluvio-deltaic origin. Farther to the east, the coeval Malanzán
britannica). Because these structures are preserved along bed- Formation contains a low-diversity suite of grazing trails, record-
ding planes, recording emplacement in very shallow tiers, degree ing colonization of turbidite sandstone (Buatois and Mángano,
of bioturbation (as seen in cross section) is typically zero. The 1995a). Acritarchs have been recorded in the Malanzán succes-
Mermia assemblage is present not only in the western area of the sion, although not at the same levels as the trace fossils (Gutiérrez
Paganzo Basin but also in the eastern region. and Limarino, 2001).
In the Guandacol Formation, the Mermia assemblage occurs
in parallel-laminated siltstone and current ripple cross-laminated The San Rafael Basin (Western Argentina)
and parallel-laminated very fine grained sandstone mostly depos-
ited by delta-fed underflow currents and suspension fallout. This The San Rafael Basin is a northwest-southeast trending late
assemblage is present through the middle interval of the postgla- Paleozoic back-arc basin located in western Argentina (Espejo
cial succession, representing early highstand systems tract depo- et al., 1996). It is limited to the south and southwest by the
sition. This assemblage is typical of freshwater environments and Neuquén Basin, and to the north and northeast by the Pampean
represents an example of the Mermia ichnofacies. However, the Ranges. The Upper Carboniferous column is subdivided into the
presence of acritarchs in some beds, mostly in the western region, El Imperial and Agua Escondida Formations (Azcuy et al., 1999;
indicates a marine connection and the existence of intermittent Césari and Gutiérrez, 2000). The El Imperial Formation consists
periods of brackish-water conditions. Integration of palynologic mostly of siltstone and sandstone, with diamictite toward the base
and ichnologic data shows that acritarchs do not occur in beds of the succession. This unit has been subdivided into three main
containing trace fossils, which are commonly associated with ter- intervals (Arias and Azcuy, 1986). The lower interval consists of
A E

Figure 3. Trace fossils from the Guandacol Formation (Huerta de Huachi, Paganzo Basin, western Argentina). (A) Helminthoid-
ichnites tenuis. (B) Helminthopsis tenuis. (C) Diplichnites gouldi showing poor preservation of individual appendage imprints.
(D) Maculichna carboniferous. Preservational variety showing only one of the elements of the paired tracks. (E) Orchesteropus
atavus. (F) Undichna britannica. Specimens housed at the Invertebrate Paleontology collection of the Instituto Miguel Lillo, Uni-
versidad Nacional de Tucumán, Argentina. All bars are 1 cm (after Buatois and Mángano, 2003).
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 155

A B

D E

Figure 4. Trace fossils from the Agua Colorada Formation (Sierra de Narváez, Paganzo Basin, western Argentina). (A) Cochlichnus anguineus.
(B) Helminthoidichnites tenuis. (C) Treptichnus pollardi. (D) Mermia carickensis. (E) Undichna insolentia. Specimens housed at the Inverte-
brate Paleontology collection of the Instituto Miguel Lillo, Universidad Nacional de Tucumán, Argentina. All bars are 1 cm (after Buatois and
Mángano, 1993).

very coarse to fine-grained sandstone, siltstone, and pebble to event, the postglacial transgression, and the subsequent highstand
cobble conglomerate deposited by turbidity currents (Arias and progradation (Pazos et al., 2007).
Azcuy, 1986). Pazos et al. (2007) mentioned evidence of wave The trace-fossil content of the glacial to postglacial event in
action in some of the sandstone, indicating that at least part of the the San Rafael Basin was recently analyzed by Pazos et al. (2007).
turbidite succession was emplaced in shallow water. The middle These authors recognized three main trace-fossil assemblages
interval consists of diamictite and interbedded medium- to very in the El Imperial Formation. Their assemblage A occurs in the
fine grained sandstone and siltstone deposited in delta-front to lower interval and is dominated by Diplopodichnus biformis and
delta-plain settings (Arias and Azcuy, 1986; Pazos et al., 2007). Diplichnites gouldi, with subordinate occurrences of Archaeo-
Fine-grained distal deposits commonly contain dropstones, nassa fossulata. Assemblage B is present in the lower part of the
whereas slumps occur in delta front deposits. The maximum middle interval and contains simple feeding structures (Treptich-
flooding surface is contained within a black shale package. The nus pollardi), grazing trails (Helminthoidichnites tenuis, Mermia
upper interval consists of coarse- to fine-grained sandstone and carickensis, Gordia marina, Cochlichnus anguineus), arthropod
siltstone recording deposition in anastomosed fluvial systems trails and trackways (Diplopodichnus biformis, Maculichna
(Arias and Azcuy, 1986). The middle interval records the glacial carboniferus), and fish trails (Undichna isp.). This assemblage
156 Buatois et al.

favors comparison with the Mermia assemblage of the Guanda- layers of very fine grained silty sandstone and siltstone, locally
col and Agua Colorada Formations. Assemblage C occurs in the with rippled tops and convolute lamination, interpreted as the
black shale package of the middle interval and consists only of product of underflow currents, and thin mudstone partings cover-
Didymaulichnus lyelli and Diplopodichnus biformis. This assem- ing medium- to fine-grained sandstone with wave ripples, wave
blage is similar to the Didymaulichnus assemblage of the Guan- ripple-cross lamination, and microhummocky cross-stratification,
dacol Formation. deposited by oscillatory flows. These deposits overlie diamictite
Terrestrially derived palynomorphs occur throughout the and sandy turbidites, and underlie a thick diamictite package. The
whole succession (García, 1995). Rare acritarchs (Gorgonispha- ichnofauna from the Tarija Formation is assigned to the Mermia
eridium spp.) have been recorded in only one sample from the ichnofacies. Taxonomic composition, nature of preservation, and
middle interval of El Imperial Formation (Pazos et al., 2007). overall features of the ichnofauna (e.g., presence of very simple
Intense fluvial runoff into fjord environments and sporadic grazing patterns of epifaunal animals and absence of trace fossils
marine connections were indicated for the El Imperial Formation of infaunal organisms and typical ichnotaxa of marine environ-
(Pazos et al., 2007). ments) suggest freshwater conditions during accumulation of the
ichnofossil-bearing deposits. This is consistent with the absence
The Tarija Basin (Northwestern Argentina) of marine body fossils and acritarchs in the unit. These condi-
tions may have occurred in lakes or fjords affected by a strong
The Tarija Basin extends ~1000 km from southeastern Peru discharge of fresh water due to melting of the ice masses.
in the north, through the Subandean region of Bolivia, and into
northwestern Argentina toward the south (Fig. 1). It is considered The Paraná Basin (Southern Brazil)
an asymmetric intracratonic basin bounded by the Guapore Cra-
ton to the north and east, the Puna arch to the south and west, and The Paraná Basin is a widespread intracratonic depression
the Michicola arch to the south and east (Eyles et al., 1995; Azcuy that covered the entire southern portion of Brazil, as well as south-
and di Pasquo, 1999; Starck and del Papa, 2006). In northwest eastern Paraguay, northeastern Argentina, and northern Uruguay
Argentina, the Tarija Basin stratigraphic succession is subdivided (Fig. 1). It was formed during the Ordovician, being completely
into three groups: the Macharetí, Mandiyutí, and Cuevo. The filled by the Late Cretaceous. During the Late Carboniferous–
lower Upper Carboniferous (Namurian–Westphalian) Macharetí Early Permian, the landscape of the Paraná Basin area contained
Group contains rocks of the Gondwana glaciations, and is in turn multi-lobed glacial fronts, forming fjord-like incised valleys
subdivided into the Tupambi, Itacuamí, and Tarija Formations opening toward a shallow epicontinental sea (Santos, 1987;
(Azcuy and di Pasquo, 1999). The Tupambi Formation is domi- Santos et al., 1996). The dominantly glaciomarine deposits of
nated by sandstone and represents sedimentation in periglacial the Rio do Sul Formation represent the last glacial episode in
environments characterized by estuarine and deltaic deposits fill- the Paraná Basin, and the maximum flooding event related to the
ing incised valleys (López Gamundí, 1986; Starck et al., 1993; Gondwana deglaciation. Its eastern counterpart is represented by
Starck and del Papa, 2006). The Itacuamí Formation consists of the Dwyka Series deposits, revealing a southeastern connection
parallel-laminated mudstone with dropstones recording deposi- between the Paraná and the Karoo Basins (South Africa) during
tion in periglacial settings (Starck et al., 1993; del Papa and Mar- this time (Santos, 1987). Its southern counterpart is represented
tínez, 2001; Starck and del Papa, 2006). The Tarija Formation by the San Gregorio Formation from the North Uruguayan Basin
consists of glacial diamictite, medium- to fine-grained sandstone (Goso, 1995).
with current and combined-flow ripples, and parallel-laminated Six second-order sequences were recognized in the Paraná
mudstone with dropstones (del Papa and Martínez, 2001; Starck Basin (Milani, 1997). The Gondwana I sequence contains the
and del Papa, 2006). glaciogenic deposits of the Upper Carboniferous–Lower Perm-
Buatois and del Papa (2003) documented trace fossils in ian (Stephanian–Sakmarian) Itararé Group. In outcrop, the Ita-
the uppermost part of the Itacuamí Formation, and in the lower raré Group is subdivided into four formations: the Aquidauana,
interval of the Tarija Formation (Fig. 2). The Itacuamí ichno- Campo do Tenente, Mafra, and Rio do Sul Formations (Sch-
fauna consists of the locomotion trail Diplopodichnus biformis, neider et al., 1974). Fluvial fine- to coarse-grained sandstone and
locally intergrading with the trackway Diplichnites gouldi glaciomarine diamictite are the main facies associations in the
(Diplopodichnus-Diplichnites assemblage) (Fig. 5B). It occurs Aquidauana Formation, whereas glaciolacustrine massive silt-
in suspension-fallout, parallel-laminated mudstone underlying stone and shale, and glaciofluvial diamictite, sandstone, and con-
diamictite. This assemblage is assigned to the Scoyenia ichnofa- glomerate dominate in the Campo do Tenente Formation (Castro
cies. Overall, the ichnofauna from the Tarija Formation (Mermia et al., 1994). Striated pavements and striated and faceted clasts
assemblage) is dominated by tiny nonspecialized grazing trails are common.
(Mermia carickensis, Gordia marina, Helminthopsis tenuis, The basin depocenter at this time was located in the mid-
Helminthoidichnites tenuis), with subordinate locomotion traces southern portion of Brazil (central and northern Santa Catarina
(Cochlichnus isp., Diplopodichnus biformis, Diplichnites gouldi) State to south of Paraná State), where the thickest successions
(Fig. 5A, 5C, and 5D). The Tarija ichnofauna is present in thin of the Itararé Group are well exposed. The Mafra and the Rio do
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 157

A B

C
H

G
D D
Figure 5. Trace fossils from the Itacuamí, and Tarija Formations (Tarija Basin, northwest Argentina). (A) Diplopodichnus biformis. Nega-
tive epirelief in mudstone partings blanketing sandstone with wave ripples. Aguas Blancas. Tarija Formation. (B) Diplichnites gouldi (white
arrow), and Diplopodichnus biformis (gray arrow). Negative epirelief in laminated mudstone. Arroyo Iquira. Itacuamí Formation. PIL 14843.
(C) Mermia carickensis preserved in mudstone parting blanketing sandstone with wave ripples. Aguas Blancas. Tarija Formation. PIL 14854.
(D) Helminthopsis tenuis (H), Diplopodichnus biformis transitional with Diplichnites gouldi (D), and Gordia marina (G). Positive hyporelief
in sandstone with wave ripples. Aguas Blancas. Tarija Formation. Specimens housed at the Invertebrate Paleontology collection of the Instituto
Miguel Lillo, Universidad Nacional de Tucumán, Argentina. All bars are 1 cm (after Buatois and del Papa, 2003).

Sul Formations are composed chiefly of deposits that originated and possibly hummocky cross-stratification. Matrix-supported
during deglaciation (Fig. 2). The lower interval of the Mafra For- diamictite with faceted clasts occurs throughout the whole suc-
mation consists of a glacio-deltaic succession that is replaced cession. Widespread marine fossiliferous shale recording maxi-
upward by a tide-influenced shallow-marine succession, consist- mum flooding characterizes the base of the overlying Rio do Sul
ing of medium- to large-scale, trough cross-stratified sandstone Formation, which consists mostly of fine-grained heterolithic
representing tidal bars (Schneider et al., 1974). The middle inter- deposits; diamictites are less abundant than in the Mafra Forma-
val consists of very thinly bedded silty-muddy rhythmites filling tion. Dropstones of different sizes are present.
an incised valley. Dropstones are abundant, ranging from gran- The ichnology of the Itararé Group has been analyzed in
ules (more common) to boulders. The upper interval of the Mafra numerous papers (Guerra-Sommer et al., 1984; Fernandes et al.,
Formation consists of thinly interbedded fine- to very fine grained 1987; Dias-Fabrício and Guerra-Sommer, 1989; Nogueira and
sandstone and siltstone, and fine- to medium-grained sandstone Netto, 2001a, 2001b; Balistieri and Netto, 2002; Balistieri et al.,
with trough cross-stratification, herringbone cross-stratification, 2002, 2003a; Lermen, 2006; Gandini et al., 2007; Netto et al.,
158 Buatois et al.

2009). The rhythmites of the Mafra Formation are well exposed behaviors, revealing subaqueous colonization of soft substrates.
in northern Santa Catarina State, where they make up part of the The behavioral characteristics of the assemblage, the absence of
middle interval of the unit and contain two distinctive trace-fossil truly marine ichnotaxa, the presence of tiny trails, and the domi-
assemblages (Balistieri et al., 2002, 2003a). The first assemblage nance of simple feeding strategies are consistent with character-
(Diplopodichnus-Diplichnites assemblage) occurs all through- istics of the Mermia ichnofacies, suggesting colonization by a
out the rhythmite package, and consists exclusively of arthropod freshwater fauna (Balistieri, 2003; Balistieri et al., 2002, 2003a;
trackways (Diplichnites gouldi) (Fig. 6A) and trails (Diplopod- Netto et al., 2009). Notably, in the thicker-bedded rhythmites,
ichnus biformis) commonly associated with wrinkle marks. Both both trace-fossil assemblages coexist, representing distinct suites,
ichnotaxa are commonly intergradational, reflecting production with elements of the Mermia ichnofacies crosscut by the terres-
by the same tracemaker, most likely a myriapod, but in a sub- trial suite, suggesting that the ponds periodically dried up (see
strate ranging from soft (Diplopodichnus biformis) to slightly Buatois and Mángano, 2004, 2007).
firm (Diplichnites gouldi) (Johnson et al., 1994; Keighley and The upper interval of the Mafra Formation contains diam-
Pickerill, 1996; Buatois et al., 1998; Paz et al., 2002, 2003, 2004; ictite and fine-grained heterolithic deposits that occur above the
Balistieri, 2003; Balistieri et al., 2002, 2003a; Netto et al., 2009). marine shale interval. The heterolithic deposits consist of inter-
Balistieri et al. (2002, 2003a) interpreted this trace-fossil assem- bedded fine- to very fine grained sandstone, and of siltstone with
blage as recording arthropod locomotion during periods of sub- flaser and wavy bedding. The heterolithic facies contains ?Areni-
aerial exposure of the substrate, most likely in mud flats flanking colites isp., Chondrites isp., Diplocraterion isp., Palaeophycus
the fjord valley. Dryness is important in glacial climates, and isp., Planolites isp., Rhizocorallium isp., and Thalassinoides isp.
promotes constant water evaporation and sublimation, especially (Balistieri, 2003). This trace-fossil assemblage occurs at several
during periods of colder temperatures when shallow-water bod- stratigraphic intervals within the Mafra Formation and extends
ies usually freeze. In contrast, during periods of higher tempera- into the overlying Rio do Sul Formation, invariably associated with
tures, these shallow-water bodies defrost, supplying sediment to heterolithic tide-influenced facies. Most ichnotaxa in this assem-
outwash plains. blage are facies-crossing trophic generalists, a strong evidence
The second assemblage (Mermia assemblage) occurs toward of opportunistic behavior (Ekdale, 1985; Bromley, 1996). These
the upper part of the rhythmite package and is dominated by non- characteristics suggest substrate colonization by a brackish-water
specialized grazing trails (e.g., Cochlichnus anguineus, Gordia fauna, producing an impoverished mixed Cruziana-Skolithos
arcuata, Gordia marina, Helminthoidichnites tenuis) and feed- ichnofacies (Pemberton and Wightman, 1992; MacEachern and
ing traces (Hormosiroidea meandrica, Treptichnus pollardi) Pemberton, 1994; Pemberton et al., 2001; Netto and Rossetti,
(Fig. 6B). Accessory components include arthropod locomotion 2003; Buatois et al., 2005; MacEachern and Gingras, 2007).
(Cruziana cf. problematica) and resting traces (Rusophycus cf. At the base of Rio do Sul Formation, heterolithic deposits
carbonarius, Gluckstadtella isp.) and fish trails (Undichna con- are punctuated by massive siltstone that contains bivalves, articu-
sulca). Shallow grazing and feeding are the main ethological lated brachiopods (Rocha-Campos, 1970), and a Glossifungites

A B

Figure 6. Trace fossils from rhythmites of the Mafra Formation (northern Santa Catarina State, Paraná Basin, southern Brazil). (A) Diplichnites
gouldi. (B) Hormosiroidea meandrica. Specimens housed at the Paleontology collection of Centro Paleontológico de Mafra, Brazil. All bars are
1 cm (after Balistieri et al., 2002, 2003a).
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 159

suite (Balistieri and Netto, 2002) (Fig. 7). This suite is domi- arthropod resting traces (Gluckstadella cooperi) (Gandini et al.,
nated by Thalassinoides isp., and also contains Diplocraterion 2007) (Fig. 8). Balistieri et al. (2002) also recorded chevronate
isp., Palaeophycus isp., P. striatus, ?Rhizocorallium isp., and trails in rhythmites of the base of the Rio do Sul Formation,
Gyrolithes-like burrows (see Netto et al., 2007), revealing bur- assigned to the ichnogenus Protovirgularia, and noted that struc-
rowing in compacted, firm mud (Balistieri and Netto, 2002). tures previously assigned to Gyrochorte isp. by Guerra-Sommer
The substrate-controlled Glossifungites suite delineates a firm- et al. (1984) and Marques-Toigo et al. (1989) most likely repre-
ground surface formed by erosional exhumation during trans- sent Protovirgularia. Regardless of the ichnotaxonomic assign-
gression (MacEachern et al., 1992; Pemberton et al., 2001). ment, these structures are most likely produced by arthropods
The transgressive trend is further revealed by the presence of rather than bivalves. The two arms of the V-shaped structures are
fossiliferous black shale (Lontras Shale) overlying this interval. not perfectly articulated at the axial area, indicating that they are
This shale contains linguliformean brachiopods (Lingula sp. and not the result of the impression of a split muscular foot, but rather
Orbiculoidea guaraunensis), articulated brachiopods (Anthra- reflect the active and passive phases of an appendage stroke.
coneilo sp., Barroisella imbituvensis, Chonetes riograndensis, The upper interval of the Rio do Sul Formation, which crops
Chonetes rionegrense, Crurithyris aff. planoconvexa, Crurithy- out in central Santa Catarina State, contains arthropod track-
ris roxoi, and Langella imbituvensis) bivalves (Leda woodworthi ways (Diplichnites isp., Umfolozia isp.), trails (Cruziana isp.,
and Nuculana woodworthi), paleoniscid and celacanthiid scales, Diplopodichnus isp.), and resting trace fossils (Gluckstadtella
shark teeth, arenaceous forams, and sponge spicules (Ruedmann, isp., Rusophycus isp.), together with grazing trails (Helmint-
1929; Oliveira, 1930; Mendes, 1952; Mezzalira, 1956; Rocha- hoidichnites isp.) (Nogueira and Netto, 2001a, 2001b). The Rio
Campos, 1970). Terrestrial elements in this shale interval include do Sul assemblage is similar to the assemblages recorded in the
blattoid wings and terminal stems of gymnosperms (Pinto and rhythmites of the Mafra Formation, albeit with a lower participa-
Sedor, 2000). The shale interval indicates flooding of marginal- tion of grazing trails. The sporadic presence of marine plankton
marine areas during the transgressive peak. indicates periodic connection with the sea.
The lower interval of the Rio do Sul Formation is also In the southernmost portion of the Paraná Basin (Rio Grande
exposed in northern Santa Catarina State. These deposits contain do Sul State), the glaciogenic deposits of the Itararé Group have not
abundant Protovirgularia isp., arthropod trackways (Diplichnites been divided into formations (Schneider et al., 1974). These con-
gouldi, Diplopodichnus biformis, Maculichna varia, Umfolo- sist of very thin-bedded silty-muddy rhythmites overlying thick-
zia sinuosa), common shallow burrows and trails (Cochlichnus bedded diamictite and interbedded with fine-grained sandstone,
anguineus, Hormosiroidea meandrica, Treptichnus isp.), and rare interbedded fine-grained sandstone and siltstone, and black shale

A B
Figure 7. Trace fossils from a ravinement surface at the top of the Mafra Formation (northern Santa Catarina State, Paraná Basin, southern Bra-
zil). (A) Diplocraterion isp. (B) Thalassinoides isp. Specimens not collected. All bars are 1 cm (after Balistieri and Netto, 2002).
160 Buatois et al.

A B

Gc

Db Hm
C D
Hm

Dg
Mv
Us
Dg
Mv
Db

Figure 8. Trace fossils from the middle interval of Mafra Formation, and lower interval of the Rio do Sul Formation (northern Santa Catarina
State, Paraná Basin, southern Brazil). (A) Gluckstadella cooperi (Gc) overprinted to Protovirgularia isp., forming a palimpsest surface (P).
(B) Treptichnus isp. (C) Diplopodichnus biformis (Db), Maculichna varia (Mv), Hormosiroidea meandrica (Hm), and Diplichnites gouldi (Dg).
(D) Palimpsest surface with Umfolozia sinuosa (Us), intergradational Diplopodichnus biformis (Db) and Diplichnites gouldi (Dg), Maculichna
varia (Mv), and Hormosiroidea meandrica (Hm). Specimens housed at the Paleontology collection of the Centro Paleontológico de Mafra, and
at the Invertebrate Paleontology collection of Universidade do Vale do Rio dos Sinos, Brazil. All bars are 1 cm (after Gandini et al., 2007).

(Correa da Silva, 1978). Dropstones are common in fine-grained Pinto, 1947, 1949, 1955, 2000; Martins, 1948, 1951; Martins and
deposits, and diamictites contain striated and faceted clasts. Stri- Sena Sobrinho, 1950; Pinto and Purper, 2000; Richter, 2000).
ated pavements suggest long periods of nondeposition, possibly This fossil assemblage is, in part, similar to that preserved in
related to ice cap migration over highlands formed by basement the Lontras Shale (Santa Catarina State) and the Passinho Shale
rocks (Tomazelli and Soliani, 1982). The black shale (locally (Paraná State).
known as Budó Shale) contains a mix of marine and nonmarine Trace fossils from the Itararé Group rhythmites of Rio Grande
fossils, composed of bivalves (Aviculopecten cambahyensis), do Sul State were originally reported by Guerra-Sommer et al.
brachiopods (Langella imbituvensis, Orbiculoidea maricaen- (1984) and Dias-Fabrício and Guerra-Sommer (1989). A recent
sis), marine scolecodonts (Nereidavus moreirai, Nereidavus ichnotaxonomic review of this material (Lermen, 2006) reveals
beetleae, Ildraites langei, Arabellites almeidai, Arabellites san- the presence of two distinct paleoichnocoenoses, one dominated
tosi), ostracods, sponge spicules, fish remains (Elasmobranchii, by arthropod trackways (Fig. 9) and preserved in thin silty-muddy
paleonisciform and cladontiform teeth, coelacanthid and pale- rhythmites, and the other one recorded by a Chondrites-Plano-
oniscid scales), paraplecopteran insects (Narkemina rochacam- lites-Palaeophycus composite ichnofabric, occurring in hetero-
posi), a Glossopteris flora, and algal remains (Dolianiti, 1945; lithic fine-grained sandstone and siltstone. Shallow burrows, as
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 161

C D
Figure 9. Trace fossils from the Itararé Group rhythmites (Rio Grande do Sul State, Paraná Basin, southern Brazil). (A) Cruziana problemati-
ca. (B) Maculichna varia. (C) Kouphichnium isp. (D) Protichnites isp. Specimens housed at the Paleobotany Paleontology collection of Univer-
sidade Federal do Rio Grande do Sul, Brazil. All bars are 1 cm (after Lermen, 2006).

well as arthropod resting traces and intrastratal trails, are subordi- sive and reflects deposition controlled by basement highs (Goso,
nate elements in the rhythmite paleoichnocoenosis. 1995). This unit contains a relatively diverse marine fauna,
Trace-fossil information suggests a marginal-marine, possi- including orthocone cephalopods (Dolorthoceras chubutensis),
bly fjord-like depositional setting, similar to that represented by goniatites (Eoasianites (Glaphyrites) rionegrensis), radiolarians,
the Mafra Formation deposits. The rhythmite ichnocoenosis has a sponges (Itararella gracilis, Microhemidis ortmanni), fish remains
strong freshwater signature, whereas the Chondrites-Planolites- (Coccocephalichthys tesselatus, Carbonilepsis uruguayensis,
Palaeophycus composite ichnofabric most likely formed under Daphnaechelus formosus, Elonichthys macropercularis, Gond-
brackish-water conditions. The dominance of arthropod track- wanaichthys maximus, Itaratichthys microphtalmus, Mesonich-
ways in the rhythmite ichnocoenosis suggests affinities with thys antipodeus, Rhadinichthys rioniger), and arthropods (Mar-
the Scoyenia ichnofacies, whereas the heterolithic-facies ichno- tínez Machiavello, 1963; Closs, 1967, 1969; Kling and Reif,
coenosis represents an impoverished Cruziana ichnofacies. The 1969; Ybert and Marques-Toigo, 1970; Marques-Toigo, 1970,
trace-fossil content of the Itararé Group in Rio Grande do Sul 1973a, 1973b, 1974; Beltan, 1978, 1981; Mones and Figueiras,
State is similar to that preserved in rhythmites of the Rio do Sul 1980; da Silva, 1985; Beri, 1987, 1991; Beri and Daners, 1994).
Formation in northern and central Santa Catarina State, further The precise age of the San Gregorio Formation has been debated.
supporting the hypothesis that these deposits represent the last The cephalopod and fish remains suggest a Pennsylvanian age
deglaciation events in the Paraná Basin. whereas the palynomorphs indicate Artinskian-Sakmarian.
The Paraná Basin extends toward the south into Uruguay The San Gregorio Formation contains an ichnofauna that
(the so-called North Uruguayan Basin), occupying an area of includes typical marine ichnogenera, such as Phycosiphon,
~90,000 km2. This basin was filled mostly by siliciclastic Paleo- Chondrites, Rhizocorallium, Thalassinoides, and Paleobullia,
zoic and Mesozoic deposits (Goso, 1995). Four major sedimen- accompanied by some facies-crossing forms (e.g., Palaeophy-
tary sequences have been recognized in this region, ranging from cus, Diplocraterion, Taenidium, Cruziana, Gordia, Planolites,
Middle Devonian to Paleogene. The glacial-influenced deposits Protichnites) (Netto and Goso, 1998; Balistieri et al., 2003b).
of the San Gregorio Formation correspond to the initial deposi- This ichnofauna is similar to the mixed Skolithos-Cruziana
tion of the Permian–Triassic sequence represented by the Cerro association recorded in the heterolithic deposits of the lower
Largo Group (Preciozzi et al., 1988). interval of the Rio do Sul Formation in northern Santa Catarina
The San Gregorio Formation (equivalent of the Itararé Group State (southern Brazil). The low to moderate intensity of bio-
and the middle interval of the Ecca Group) is regionally exten- turbation and ichnodiversity, and the dominance of trace fossils
162 Buatois et al.

made by trophic generalists, suggests brackish-water conditions Descriptions of the Dwyka and Ecca ichnofaunas have been
in marginal-marine environments. This interpretation is consis- provided by Savage (1970, 1971) and Anderson (1970, 1975a,
tent with those proposed by da Silva (1985), who regarded these 1975b, 1976, 1981). The Dwyka ichnofauna consists of arthro-
glaciogenic deposits as estuarine on the basis of palynomorphs, pod trackways (Umfolozia sinuosa, U. longula, Maculichna
and by Goso (1995), who suggested that the absence of marine varia, Diplichnites isp.) and resting traces (Gluckstadtella coo-
plankton resulted from meltwater discharge. However, the abun- peri, Kingella natalensis), and fish trails (Undichna insolentia,
dance of ichnofabrics dominated by Chondrites and Phycosi- U. simplicitas, U. bina, Undichna isp.). At least part of the mate-
phon points to periods of more marine influence, more typical of rial originally referred to Gyrochorte isp. by Savage (1971) may
prodeltaic settings. be included in Protovirgularia, although a bivalve affinity is ques-
tionable, and the structures were most likely produced by arthro-
The Karoo Basin (South Africa) pods, as in the case of the specimens from the Paraná Basin. The
nature of the tracemaker is significant because bivalve-generated
The Karoo Basin is a retro-arc foreland basin located chevronate structures are produced by cleft-foot bivalves, which
north of the Cape Fold Belt that contains up to 8 km of strata are exclusively marine. Most of the trace fossils were recorded in
(Rubidge et al., 2000). It represents the remnant of a more exten- a single quarry near Swart Umfolozi in northern Natal (Fig. 10),
sive roughly north-south-trending basin along part of the Gond- but other localities in western South Africa and southern Namibia
wanic paleo-Pacific margin (Visser, 1987). Reconstructions sug- have provided specimens, as well as a core in Orange Free
gest that the most important ice-spreading centers were located State. The freshwater affinities of the Dwyka ichnofauna have
over south-central Africa (Visser, 1997; Catuneanu, 2004). The been noted by almost all workers in the area (Anderson, 1970,
Upper Paleozoic column of the Karoo Basin is included in the 1975a, 1976, 1981; Savage, 1970, 1971). This is consistent with
Karoo Supergroup. The Upper Carboniferous–Upper Permian the restricted distribution of marine body fossils during Dwyka
succession has been subdivided into three groups, Dwyka, Ecca, times (McLachlan and Anderson, 1973). Cephalopods and nucu-
and Beaufort (Visser, 1983, 1995, 1996, 1997; Catuneanu et al., lanid bivalves have been recorded in the western part of the basin,
1998). The Upper Carboniferous–Lower Permian (Stephanian– whereas acritarchs are known from boreholes in the south. The
Asselian) Dwyka Group consists of glacial diamictite separated Dwyka ichnofauna is virtually identical to the one described from
by varve shale and siltstone, representing glacial and interstadial the Rio do Sul Formation rhythmites in Brazil.
phases, respectively (Visser, 1983, 1997). The Ecca ichnofauna is more variable. Arthropod trackways
The stratigraphy of the Lower–Middle Permian (Asselian– (Umfolozia sinuosa, U. longula) and fish trails (Undichna inso-
Capitanian) Ecca Group varies from section to section (Catu- lentia, U. simplicitas, U. bina) have been described from a num-
neanu et al., 2005). The lower Ecca includes three formations, ber of localities both in the southern “turbidite” facies and in the
Prince Albert, Whitehill, and Collingham. The Prince Albert For- basin center facies (Anderson, 1970, 1976, 1981). Also, Kouph-
mation consists of mudstone with chert, carbonate, and phosphate ichnium isp. was recorded at the top of the Whitehill Formation,
nodules, recording syn- to postglacial suspension fallout and epi- and in the upper Ecca close to the boundary with the Beaufort
sodic deposition from turbidity currents and mud flows (Visser Group (Anderson, 1975b). Marine ichnotaxa (e.g., Chondrites,
and Young, 1990). The Whitehill Formation consists of black Lorenzinia, Lophoctenium) have been mentioned (but not illus-
organic-rich shale, representing deposition under anoxic condi- trated) by Johnson et al. (2001) in turbidites from the Skoorsteen-
tions (Visser, 1992; Johnson et al., 1997). The Collingham For- berg Formation in the southwest portion of the basin. Marine
mation consists of chert, mudstone, ash, and sandstone, recording trace-fossil assemblages containing Siphonichnus eccaensis,
suspension fallout alternating with low-density turbidity currents Spirodesmos archimedeus, Rhizocorallium isp., Skolithos isp.,
in a basin floor (Viljoen, 1994). Toward the foreland, the upper and Diplocraterion isp. have been documented in deltaic deposits
Ecca is subdivided into the Tierberg, Skoorsteenberg, Kookfon- of the Middle Ecca in Natal (Hobday and Tavener-Smith, 1975;
tein, and Waterford Formations (Wickens, 1996; Catuneanu et al., Stanistreet et al., 1980; Turner et al., 1981; Mason et al., 1983;
2005). The Tierberg Formation consists of dark basinal shale, Tavener-Smith and Mason, 1985). Other ichnotaxa described
whereas the Skoorsteenberg Formation records five major sand- from the Ecca need to be reevaluated. Marine body fossils are
rich turbidite systems separated by fine-grained intervals (Scott rare in the Ecca. The most abundant marine fauna occurs in the
et al., 2000; Johnson et al., 2001). The Kookfontein Formation Whitehill Formation and consists of the reptile Mesosaurus, the
consists of thick sandstone units separated by thin mudstone fish Palaeoniscus, and the crustacean Notocaris tapscotti (Smith
intervals, mostly recording deltaic progradation (Wickens, 1996). et al., 1993; Johnson et al., 1997). Cephalopods have also been
The Waterford Formation consists of sandstone and siltstone, and described in middle Ecca beds in Natal (Rilett, 1963; Teichert
records shallow-marine to paralic deposition (Rubidge et al., 2000; and Rilett, 1974), and bivalves are known from the Waterford
Catuneanu et al., 2005). The Upper Permian (Wuchiapingian– Formation in southern Cape Province (Cooper and Kensley,
Changhsingian) to Middle Triassic Beaufort Group consists of 1984; Rubidge et al., 2000). Teichert and Rilett (1974) noted the
interbedded mudstone and sandstone, recording deltaic, fluvial, presence of a bed containing a marine fauna 1.5 m above a bed
and lacustrine environments (van Dijk et al., 1978). with freshwater bivalves. Integration of ichnologic information
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 163

Us
Us
Figure 10. Trace fossils from the Dwyka
Group (Swart Umfolozi River, Karoo
Basin, South Africa). Umfolozia sinu-
osa (Us) and Undichna isp. (U). Slab is
U housed at the British Museum of Natu-
ral History, London. Scale bar is 1 cm.

within a sedimentologic, paleontologic, and stratigraphic frame- (Frakes and Crowell, 1967; Trewin et al., 2002). Trewin et al.
work is essential for unraveling paleosalinity conditions and the (2002) recorded the presence of the arthropod trackway Umfolo-
complex depositional history of the Ecca Group. zia in a mudstone unit near the base of this formation. The Lower–
Late Paleozoic deposits extend toward Namibia in the intra- Middle Permian (Artinskian-Capitanian) Port Sussex Formation
cratonic Kalahari Basin. Paleontologic evidence from the Dwyka consists of sandstone and mudstone with dropstones, represent-
and Ecca Groups in southern Namibia indicates marine conditions ing the postglacial transgression, commonly under anoxic condi-
(Martin and Wilczewski, 1970). Mollusc shells (e.g., Eurydesma tions (Trewin et al., 2002). Thin-bedded turbidites occur toward
mytiloides), bryozoan colonies, cephalopods, echinoderms, fishes the top of the formation (Shepherds Brook Member).
(Namaichthys schroederi), algal mats, and algal-serpulid build- The Middle Permian (Capitanian) Brenton Loch Formation
ups occur in the Dwyka Group in Namibia (Dickins, 1961; Gar- consists of massive turbidites with abundant dewatering struc-
diner, 1962; Martin and Wilczewski, 1970; Grill, 1997; Visser, tures and rhythmically laminated sandstone and mudstone cou-
1983). The trace fossils recorded in this unit are Rhizocorallium plets, most likely emplaced from density underflows. This unit
irregulare, Rosselia isp., and Planolites isp., indicating marine represents deposition in turbidite systems and prograding deltas
conditions (Grill, 1997). Also, a marine ichnofauna (Siphonich- (Trewin et al., 2002). Trewin et al. (2002) documented the ichnol-
nus isp., Arenicolites isp.) has been documented by this author in ogy of this unit, recording a relatively high diversity of trace fos-
shoreface deposits of the Auob Sandstone Member of the Prince sils in the middle member of the formation (Cantera Member).
Alpert Formation (Ecca Group). The Cantera Member ichnofauna consists of arthropod track-
ways (Umfolozia longula, Kouphichnium isp.), arthropod resting
The Falkland Basin (Falkland Islands) trace fossils, fish trails (Undichna bina, U. cf. insolentia), simple
grazing trails (Helminthoidichnites, Cochlichnus, Spirodesmos),
The late Paleozoic Falkland Basin (Fig. 1) is thought to rep- simple feeding trace fossils (Planolites, Treptichnus), and ver-
resent a “missing” southeast corner of the Karoo Basin after a tical burrows (Diplocraterion). A similar, although less diverse,
180° rotation of the Falkland block (Adie, 1952; Mitchell et al., ichnofauna occurs in the upper Saladero Member.
1986; Taylor and Shaw, 1989; Marshall, 1994; Trewin et al., Trewin et al. (2002) noted the remarkable similarities
2002). Upper Paleozoic rocks are subdivided into the Bluff Cove, between the Falkland ichnofaunas and those from the Paganzo
Fitzroy Tillite, Port Sussex, Brenton Loch, and Bay of Harbours Basin of Argentina and the Karoo Basin of South Africa, stress-
Formations (Trewin et al., 2002). ing the absence of any marine indicator. Therefore, they inter-
The Upper Carboniferous–Lower Permian (Stephanian- preted the late Paleozoic succession as representing deposition in
Sakmarian) Fitzroy Tillite Formation is dominated by diamictite a freshwater lake. These authors also noticed that freshwater con-
with subordinate presence of sandstone and mudstone, represent- ditions in the Falkland Basin were consistent with the absence of
ing glacially related subaqueous debris flow and esker deposits marine fossils in the eastern portion of the Karoo Basin.
164 Buatois et al.

The Transantarctic Basin (Antarctica) tuff and thinly interbedded sandstone, and siltstone, and has been
interpreted as deposited in fluvial and lacustrine environments
The Transantarctic Basin of the central Transantarctic Moun- adjacent to a volcanic arc (Campbell, 1969; Benson, 1981; Isbell
tains (Fig. 1) contains a thick succession of Permian–Jurassic et al., 2003a; Birgenheier et al., 2009). Detailed integrated sedi-
continental rocks (Barrett et al., 1986; Collinson et al., 1994). The mentologic and ichnologic studies are lacking, but an ichnofauna
Permian column is subdivided in the Pagoda, Mackellar, Fair- of arthropod trackways (Umfolozia sinuosa) and trails (Cruziana
child, and Buckley Formations. The Upper Carboniferous–Lower problematica) is present in a siltstone interval overlying glacial
Permian Pagoda Formation consists of diamictite and sandstone diamictite (Fig. 11).
recording periglacial and glacial deposition (Barrett et al., 1986;
Miller, 1989; Isbell et al., 2001). The Lower Permian Mackel- DISCUSSION
lar Formation consists of fine-grained deposits, essentially shale,
siltstone, and thin-bedded sandstone, and records the postglacial Recent research in late Quaternary marine deposits associated
transgression in a large and deep lake affected by underflow and with glacial meltwater has emphasized the role of hyperpycnal
turbidity currents (Miller and Collinson, 1994; Collinson et al., flows in connection with high freshwater discharges (e.g., Piper
2004; Miller and Isbell, this volume). The Lower–Middle Perm- et al., 2007; Tripsanas et al., 2007). High freshwater discharges
ian Fairchild Formation consists of sandstone formed in low- due to glacier melting and associated catastrophic outburst floods
sinuosity fluvial systems that prograded into the lake, forming a have been documented for the Baltic and Labrador Seas (Lord,
braided outwash plain (Collinson et al., 1994; Miller and Isbell, 1990; Shaw and Lesemann, 2003). In addition, times of elevated
this volume). The Middle–Upper Permian Buckley Formation concentration of suspended sediment promote hyperpycnal flows
consists of sandstone, siltstone, shale, and coal formed in high- in modern fjords (Syvitski et al., 1987). Only exceptional dis-
sinuosity fluvial systems (Knepprath et al., 2004). charges overcome the buoyancy effect of seawater in modern
The ichnology of postglacial deposits of the Mackellar For- examples, but high-discharge hyperpycnal flows may have been
mation has been discussed by Miller and Collinson (1994) and by extremely common during deglaciation. Also, because large dis-
Miller and Isbell (this volume). The ichnofauna consists of sim- charges reduce the salinity of the water body, the likelihood of
ple grazing trails (Cochlichnus, Mermia), simple feeding trace hyperpycnal flows is increased, providing a positive feedback. As
fossils (Planolites, Treptichnus), horizontal dwelling burrows a result, glacial melting led to the formation of freshwater bodies
(Palaeophycus), arthropod scratch marks, and bilobate trails, that were physically connected with the open sea. For example,
illustrating the freshwater Mermia ichnofacies. As in the case of the Holocene Yoldia Sea was freshwater in the northern Baltic Sea
other Gondwana ichnofaunas, vertical bioturbation is negligible Basin due to a high input of meltwater during deglaciation dur-
and all ichnotaxa are parallel to the bedding plane (Miller and ing most of its history (Virtasalo et al., 2006). A wide variety of
Isbell, this volume). Geochemical data (high C:S ratios) also sup- environmental stresses that affect benthic colonization may play
port freshwater conditions (Miller and Collinson, 1994). a role in these glacially influenced environments. These include
extreme dilution of salinity, high rates of sedimentation, variable
The Sydney Basin (Eastern Australia) degrees of substrate consolidation, oxygen-depleted conditions,
high water turbidity, and intense storm activity (Feder and Keiser,
The Sydney Basin (Fig. 1), located in New South Wales, 1980; Feder and Matheke, 1980; Farrow et al., 1983; Syvitski
eastern Australia, covers 64,000 km2 both onshore and offshore, et al., 1987; O’Clair and Zimmerman, 1987). In addition, seasonal
and includes up to 6000 m of Carboniferous–Triassic strata (Gos- light restriction and floating ice masses contributing to ice-rafted
tin and Herbert, 1973; Fielding et al., 2001). The basin is consid- debris rainfall may be important stress factors in polar areas.
ered a foreland basin formed in connection with the New England Integration of ichnologic evidence with sedimentologic,
Fold Belt, which developed as a volcanic arc during the Late Car- stratigraphic, and paleontologic data allows reconciling appar-
boniferous (McPhie, 1987). In the northern area of the basin, the ently inconsistent sets of data and sheds light on the nature of
bulk of sediment fill is included, from base to top, in the Lower peri-Gondwanic ecosystems (Fig. 12). A complex paleogeogra-
Permian Dalwood Group, the Lower to Upper Permian Maitland phy of fjords and deep, large coastal lakes is proposed for the Pan-
Group, and the Upper Permian Singleton Supergroup. How- thalassan margin of Gondwana. Although these peri-Gondwanic
ever, several less well known formations (Currabubula Forma- coasts were unique in some sense, comparisons with the modern
tion, Spion Kop Conglomerate, and Seaham Formation) occur coasts of southern Chile and Norway are suggested. All eight
below the Permian–Triassic column as a series of unconformity- basins discussed in this paper contain glacial diamictites, post-
bounded units that include Upper Carboniferous glacial deposits glacial fines, and distinctive ichnofaunas. Although the units dif-
of the second glacial episode (Isbell et al., 2003a; Birgenheier fer in the degree of marine connection, the common theme in all
et al., 2009). is the presence of freshwater ichnofaunas in direct association
Of these, the Seaham Formation of earliest Namurian to ear- with glacially influenced coasts affected by strong discharges of
liest Westphalian age overlies a polished and striated surface cut meltwater. At one extreme, the Itacuamí and Tarija Formations
on the underlying Paterson Volcanics. It consists of diamictite, of the Tarija Basin, the Agua Colorada of the Paganzo Basin,
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 165

Us
Figure 11. Trace fossils from the Sea-
ham Formation (Sydney Basin, eastern
Australia). Umfolozia sinuosa (Us) and
Cruziana problematica (Cp). Slab is
housed at the Invertebrate Paleontol-
Us ogy collection of Macquarie University,
Sydney. Scale bar is 1 cm.

Us Cp

the Fitzroy Tillite and Brenton Loch Formations of the Falk- feeders and the absence of vertical burrows of suspension feeders
land Basin, and the Mackellar Formation of the Transantarctic are consistent with high amounts of suspended fine-grained mate-
Basin contain only freshwater ichnofaunas, and no in situ marine rial (MacEachern et al., 2005). The fact that ichnofaunas occur
palynomorphs or body fossils have been found. Accordingly, throughout relatively thick intervals (e.g., ~30 m in the Guan-
they may represent deposition either in freshwater lakes isolated dacol Formation) suggests that freshwater conditions were tem-
from the sea or in fjords affected by strong meltwater discharge. porally persistent in some of the analyzed basins (Buatois et al.,
The other end member is represented by the Mafra and Rio do 2006). Also, freshwater ichnofaunas can be traced for ~250 km
Sul Formations of the Paraná Basin and the Ecca Group of the along dip in the Paganzo Basin, reflecting remarkable seaward
Karoo Basin, which contain intervals with relatively abundant migration of the salinity barrier. Although some of these settings
marine faunas and ichnofaunas indicative of marine conditions have been referred to as “brackish seas,” in fact they may be more
(depauperate Cruziana and Glossifungites ichnofacies). Interme- appropriately called “freshwater seas” because of the dominance
diate cases are represented by the Guandacol Formation of the of freshwater conditions due to extensive melting during postgla-
Paganzo Basin, and the El Imperial Formation of the San Rafael cial times.
Basin. The Guandacol Formation contains a single bed with lin- Recently, geochemical analyses were performed in some of
guliformean brachiopods near its base and discrete levels with the postglacial units of South Africa (Scheffler et al., 2006; Her-
acritarchs (Martínez, 1993). One level with acritarchs has been bert and Compton, 2007). According to Herbert and Compton
recorded in the El Imperial Formation (Pazos et al., 2007). (2007), the texture, low δ18O values, and radiogenic Sr isotope
Ichnologic information from late Paleozoic peri-Gondwanic ratios of early diagenetic calcite concretions of the Dwyka Group
settings indicates that freshwater conditions prevailed in coastal and phosphatic chert concretions of the Prince Albert Formation
areas during deglaciation because of extreme discharge of fresh- of the Ecca Group support an origin in glacial, freshwater sedi-
water due to melting of ice masses (Buatois et al., 2006). In some ments. In contrast, Scheffler et al. (2006) concluded from low
of these basins (e.g., Paganzo, Paraná), glacial melting led to the CIA, Rb/K, V/Cr and TOC contents that the upper Dwyka Group
formation of freshwater bodies that were physically connected was characterized by an alternation of freshwater and brackish-
with the open sea. Recent reconstructions indicate that ice sheet water conditions in the southern part of the basin. In addition,
volumes of late Paleozoic Gondwana were in the range of 49.1– on the basis of Rb/K ratios, they noted that more marine condi-
65.4 × 106 km3, although lower volumes have been estimated tions alternated with short-term brackish-water conditions during
for many ice sheets (Isbell et al., 2003a). Complete ablation of a deposition of the lower Prince Albert Formation.
single ice sheet would have resulted in a 100 m sea-level rise with There are relatively few studies documenting ichnofaunas
a concomitant reduction in coastal salinity. In these situations, in Quaternary glacial to postglacial settings. Most of these stud-
establishment of a marine benthos is inhibited due to reduced ies focus on Pleistocene glacial lakes (e.g., Gibbard and Stuart,
salinity, allowing colonization by a freshwater biota. In addition, 1974; Gibbard, 1977; Gibbard and Dreimanis, 1978; Walter and
the dominance of horizontal feeding traces of deposit and detritus Suhr, 1998; Gaigalas and Uchman, 2004; Benner and Ridge,
Mermia–Scoyenia Ichnofacies
Ma D
U C - Terrestrially derived palynomorphs
Un - Local presence of acritarchs and/or
linguliformeans in adjacent intervals
G - e.g., Guandacol and El Imperial formations
M

Depauperate Cruziana–Skolithos Ichnofacies

Ar Pa
Di
Th
Ch

Pl

- Terrestrially derived and marine palynomorphs


- Marine Fauna
- eg., Ecca Group, Mafra & Rio do Sul
formations

Mermia–Scoyenia Ichnofacies
M C
G D

Dp Glossifungites Ichnofacies
Un - Marine palynomorphs
- Diverse marine fauna in shale intervals
overlying ravinement surfaces
- e.g., Mafra Formation
- Terrestrially derived palynomorphs Th
Di
- No marine fossils
- e.g., Agua Colorada, Fritzroy Tillite & Rz

Mackellar formations
Figure 12. Reconstruction of postglacial coastal environments and associated ichnofaunas along the Panthalassan margin of Gondwana. M—Mermia, C—Cochlichnus, G—Gordia,
Un—Undichna, D—Diplichnites, Dp—Diplopodichnus, U—Umfolozia, Ma—Maculichna, Ar—Arenicolites, Di—Diplocraterion, Th—Thalassinoides, Pa—Palaeophycus, Pl—
Planolites, Ch—Chondrites, Rz—Rhizocorallium. Not all trace-fossil assemblages are present in every depositional system or basin. Changes in coastal physiography in response to
sea-level changes, and degree of freshwater discharge result in ichnofaunal changes through time and, accordingly, some trace-fossil assemblages may be replaced by others, reflect-
ing different paleosalinity levels. Although marine influence overall increases in a seaward direction, the presence of highly irregular shores with common lateral tributaries makes
linear proximal-distal extrapolations an oversimplification at best (e.g., freshwater coastal lakes between embayed fjord areas may occur in a distal position with respect to adjacent
marine tongues).
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 167

2007; Benner et al., 2009; Uchman et al., 2009). Pleistocene ACKNOWLEDGMENTS


glacial lake ichnofaunas are dominated by nonspecialized graz-
ing trails, arthropod trackways, and, to a lesser extent, fish Our ideas on ecologic aspects of the Gondwana glaciation
trails, closely resembling Gondwana fjord and lacustrine ichno- grew out of a series of projects undertaken in many late Paleozoic
faunas. Brackish-water ichnofaunas, consisting of bivalve vertical basins. We would like to thank Cecilia del Papa, Oscar Limarino,
burrows (Siphonichnus), U-shaped vertical burrows (e.g., Areni- and Ernesto Lavina for valuable discussions. Alfred Uchman
colites, Diplocraterion), gravel-lined polychaete burrows (Dio- and Ricardo Melchor provided valuable reviews of the manu-
patrichnus), and crustacean galleries (Thalassinoides), among script. Patricio Desjardins prepared Figure 11 and Terri Graham
other forms, have been documented in Cenozoic fjord deposits assisted with the reference list. Financial support for this study
of Alaska (Eyles et al., 1992). Monospecific suites of Teichichnus was provided by the Brazilian National Council for Scientific and
occur in Holocene fjord deposits of Norway (Corner and Fjal- Technological Development (CNPq) Grants 474345/2003-3 and
stad, 1993). A recent study of Holocene cores from the Baltic 304811/2004-1 awarded to Netto, and by the Canadian Natural
Sea by Virtasalo et al. (2006) demonstrated a vertical change of Sciences and Engineering Research Council (NSERC) Discov-
ichnofaunas in connection to a shift from freshwater to brackish- ery Grants 311727-05/08 and 311726-05/08 awarded to Mán-
water conditions. The trace-fossil assemblage of the Ancylus gano and Buatois, respectively.
Lake consists of Palaeophycus and Arenicolites, representing
an extremely impoverished ichnofauna. With the advent of the REFERENCES CITED
brackish-water Littorina Sea, this assemblage was replaced by an
assemblage consisting of Planolites, Arenicolites, Lockeia, and Adie, R.J., 1952, The position of the Falkland islands in a reconstruction of
Gondwanaland: Geological Magazine, v. 89, p. 401–410, doi: 10.1017/
Teichichnus. Although an increase in ichnodiversity occurs as a S0016756800068102.
result of glacial melting, ichnodiversity levels remain very low Aitken, A.E., 1990, Fossilization potential of Arctic fjord and continental shelf
as a result of brackish-water conditions during establishment of benthic macrofaunas, in Dowdeswell, J.A., and Scourse, J.E., eds., Gla-
cimarine Environments: Processes and Sediments: Geological Society
the Littorina Sea. This pattern of vertical distribution of trace fos- [London] Special Publication 53, p. 155–176.
sils in response to a transition from freshwater to brackish water Anderson, A., 1970, An analysis of supposed fish trails from interglacial sedi-
is similar to that observed in the Paganzo and Paraná Basins, ments in the Dwyka Series, near Vryheid, Natal, in Proceedings, Inter-
national Union of Geological Sciences Gondwana Symposium, 2nd,
albeit at different scales (Balistieri and Netto, 2002; Desjardins Pretoria, South Africa, p. 637–647.
et al., this volume). Pazos et al. (2007) noted that most models Anderson, A., 1975a, Turbidites and arthropod trackways in the Dwyka glacial
for brackish environments are based on nonglacial successions deposits (Early Permian) of southern Africa: Transactions, Geological
Society of South Africa, v. 78, p. 265–273.
and speculated that they may not be applicable to glacial settings. Anderson, A., 1975b, Limulid trackways in the Late Palaeozoic Ecca sediments
However, our review demonstrates that, although many other and their palaeoenvironmental significance: South African Journal of Sci-
environmental controls obviously play a role, salinity is also ence, v. 71, p. 249–251.
Anderson, A.M., 1976, Fish trails from the early Permian of South Africa:
of importance, and in fact brackish-water ichnofaunas are well Palaeontology, v. 9, p. 397–409.
known in late Paleozoic marginal-marine deposits of Gondwana Anderson, A., 1981, The Umfolozia arthropod trackways in the Permian
and in Quaternary fjord deposits (see also Buatois et al., 2005). Dwyka and Ecca Series of South Africa: Journal of Paleontology, v. 55,
p. 84–108.
Interestingly, although coastal deposits not influenced by Andreis, R., Leguizamon, R., and Archangelsky, S., 1986, El Paleovalle de
glaciation contain fully marine ichnofaunas in the transgressive Malanzán: Nuevos criterios para la estratigrafía del Neopaleozoico de la
maximum interval, transgressive and early highstand deposits in Sierra de Los Llanos, La Rioja, República Argentina: Boletín Academia
Nacional de Ciencias, v. 57, p. 2–122.
glacially influenced basins may contain freshwater ichnofaunas, Angiolini, L., Balini, M., Garzanti, E., Nicora, A., and Tintori, A., 2003, Gond-
as a result of extreme meltwater release. This should be taken wanan deglaciation and opening of Neotethys: The Al Khlata and Saiwan
into account if trace fossils are used in sequence-stratigraphic formations of Interior Oman: Palaeogeography, Palaeoclimatology, Palae-
oecology, v. 196, p. 99–123, doi: 10.1016/S0031-0182(03)00315-8.
studies of Gondwana successions. We propose that in the case of Arias, W.E., and Azcuy, C.A., 1986, El Paleozoico superior del cañón del río
Gondwana ichnofaunas the simple dichotomy between marine Atuel, provincia de Mendoza: Revista de la Asociación Geológica Argen-
and nonmarine settings is misleading, because these peculiar tina, v. 41, p. 262–269.
Azcuy, C.A., and Morelli, J.R., 1970, Geología de la comarca Paganzo-Amaná.
assemblages should first be understood in terms of their paleo- El Grupo Paganzo: Formaciones que lo componen y sus relaciones:
ecologic significance, and subsequently placed within a larger Revista de la Asociación Geológica Argentina, v. 25, p. 405–429.
paleoenvironmental context that takes into account the unusual Azcuy, C.L., and di Pasquo, M.M., 1999, Carbonífero y Pérmico de las Sierras
Subandinas, Cordillera Oriental y Puna, in Caminos, R., ed., Geología
depositional conditions associated with Gondwana glaciations Argentina: Anales Instituto de Geología y Recursos Minerales 29,
and subsequent glacial retreat. This pattern indicates the exis- p. 239–260.
tence of laterally persistent, albeit diachronous, peri-Gondwanan Azcuy, C.L., Carrizo, H.A., and Caminos, R., 1999, Carbonífero y Pérmico
de las Sierras Pampeanas, Famatina, Precordillera, Cordillera Frontal y
ichnofaunas that characterize melting of the late Paleozoic ice Bloque de San Rafael, in Caminos, R., ed., Geología Argentina: Anales
caps. Temporal recurrence of these ichnofaunas through the Late Instituto de Geología y Recursos Minerales 29, p. 261–317.
Carboniferous–Middle Permian indicates a common response of Balistieri, P.R.M.N., 2003, Paleoicnologia da porção superior do Grupo Itararé
na região de Mafra (SC): limitações paleoecológicas, paleoambientais e
benthic faunas under similar ecological conditions during degla- estratigráficas [Doctor of Science thesis]: São Leopoldo, Brazil, Universi-
ciation events. dade do Vale do Rio dos Sinos, 128 p.
168 Buatois et al.

Balistieri, P.R.M.N., and Netto, R.G., 2002, A Glossifungites suite in depos- y biología: Publicación Especial de la Asociación Paleontológica Argen-
its of the Itararé Group (Upper Carboniferous-Lower Permian of Paraná tina, v. 9, p. 119–130.
Basin) at Mafra region, north of Santa Catarina State, Brazil: Ichnotaxon- Buatois, L.A., and López Angriman, A.O., 1992, Trazas fósiles y sistemas
omy, and paleoecological and stratigraphical constraints: Acta Geologica deposicionales, Grupo Gustav, Cretácico de la isla James Ross, Antártida,
Leopoldensia, v. 55, p. 91–106. in Rinaldi, C.A., ed., Geología de la isla James Ross: Publicación del
Balistieri, P.R.M.N., Netto, R.G., and Lavina, E.L.C., 2002, Ichnofauna from Instituto Antártico Argentino, p. 239–262.
the Upper Carboniferous-Lower Permian rhythmites from Mafra, Santa Buatois, L.A., and Mángano, M.G., 1992, Abanicos sublacustres, abanicos
Catarina State, Brazil: Ichnotaxonomy: Revista Brasileira de Paleontolo- submarinos o plataformas glacimarinas? Evidencias icnológicas para una
gia, v. 4, p. 13–26. interpretación paleoambiental del Carbonífero de la cuenca Paganzo:
Balistieri, P.R.M.N., Netto, R.G., and Lavina, E.L.C., 2003a, Icnofauna de ritmi- Ameghiniana, v. 29, p. 323–335.
tos do topo da Formação Mafra (Permo-Carbonífero da bacia do Paraná) Buatois, L.A., and Mángano, M.G., 1993, Trace fossils from a Carboniferous
em Rio Negro, Estado do Paraná (PR), Brasil, in Buatois, L.A., and turbiditic lake: Implications for the recognition of additional nonmarine
Mángano, M.G., eds., Icnología: Hacia una convergencia entre geología ichnofacies: Ichnos, v. 2, p. 237–258, doi: 10.1080/10420949309380098.
y biología: Publicación Especial de la Asociación Paleontológica Argen- Buatois, L.A., and Mángano, M.G., 1994, Lithofacies and depositional pro-
tina, v. 9, p. 131–139. cesses from a Carboniferous lake of Gondwana, Sierra de Narváez, north-
Balistieri, P., Netto, R.G., Verde, M., and Goso, C.A.A., 2003b, Comparison west Argentina: Sedimentary Geology, v. 93, p. 25–49, doi: 10.1016/0037
between two trace fossil assemblages recorded in Upper Carboniferous- -0738(94)90027-2.
Lower Permian deposits from the Paraná Basin (Rio do Sul Formation, Buatois, L.A., and Mángano, M.G., 1995a, Postglacial lacustrine event sedi-
Brazil, and San Gregorio Formation, Uruguay): Latin American Sedimen- mentation in an ancient high mountain setting: The Carboniferous Lake
tological Congress, 3rd, Belém, Brazil, Abstracts, p. 172. Malanzán from western Argentina: Journal of Paleolimnology, v. 14,
Barrett, P.J., Elliott, D.H., and Lindsay, J.F., 1986, The Beacon Supergroup p. 1–22, doi: 10.1007/BF00682591.
(Devonian-Triassic) and Ferrar Group (Jurassic) in the Beardmore Glacier Buatois, L.A., and Mángano, M.G., 1995b, Sedimentary dynamics and evo-
area, Antarctica: Antarctic Research Series, v. 36, p. 339–428. lutionary history of a Late Carboniferous Gondwanic lake in northwest-
Beltan, L., 1978, Découverte d’une ichtyofaune dans le Carbonifère Supérieur ern Argentina: Sedimentology, v. 42, p. 415–436, doi: 10.1111/j.1365
d’Uruguay. Rapport avec les faunes ichtyologiques contemporaines des -3091.1995.tb00382.x.
autres regions du Gondwana: Annales de la Société Géologique du Nord, Buatois, L.A., and Mángano, M.G., 2003, Caracterización icnológica y
v. 97, p. 351–355. paleoambiental de la localidad tipo de Orchesteropus atavus, Huerta de
Beltan, L., 1981, Coccocephalichthys tessellatus n. sp. (Pisces, Actinopterygii) Huachi, provincia de San Juan, Argentina: Ameghiniana, v. 40, p. 53–70.
from the Upper Carboniferous of Uruguay, in Anais, Congreso Latino- Buatois, L.A., and Mángano, M.G., 2004, Ichnology of fluvio-lacustrine envi-
americano de Paleontología, 2nd, Porto Alegre, Brazil, v. 1, p. 95–106. ronments: Animal-substrate interactions in freshwater ecosystems, in
Benner, J., and Ridge, J., 2007, A review of Pleistocene glaciolacustrine ichnol- McIlroy, D., ed., The Application of Ichnology to Palaeoenvironmental
ogy and its potential as a high-resolution paleoecological record during and Stratigraphic Analysis: Geological Society [London] Special Publica-
tion 228, p. 311–333.
times of rapid climate change, with examples from the late Pleistocene
Buatois, L.A., and Mángano, M.G., 2007, Invertebrate ichnology of continental
of New England, USA, in Limnogeology: Tales of an Evolving Earth:
freshwater environments, in Miller, W., III, ed., Trace Fossils: Concepts,
International Limnogeology Congress, 4th, Barcelona, Spain, Programme
Problems, Prospects: Amsterdam, Elsevier, p. 285–323.
and Abstracts Book, p. 41–42.
Buatois, L.A., Mángano, M.G., Maples, C.G., and Lanier, W.P., 1998, Taxo-
Benner, J., Ridge, J., and Knecht, R.J., 2009, Timing of post-glacial reinhabita-
nomic reassessment of the ichnogenus Beaconichnus and additional
tion and ecological development of two New England, USA, drainages
examples from the Carboniferous of Kansas, U.S.A.: Ichnos, v. 5, p. 287–
based on trace fossil evidence: Palaeogeography, Palaeoclimatology,
302, doi: 10.1080/10420949809386427.
Palaeoecology, v. 272, p. 212–231, doi: 10.1016/j.palaeo.2008.10.029. Buatois, L.A., Gingras, M.K., MacEachern, J., Mángano, M.G., Zonneveld,
Benson, J.M., 1981, Late Carboniferous and Early Permian tillites north of J.-P., Pemberton, S.G., Netto, R.G., and Martin, A.J., 2005, Colonization
Newcastle, New South Wales, in Hambrey, M.J., and Harland, W.B., eds., of brackish-water systems through time: Evidence from the trace-fossil
Earth’s Pre-Pleistocene Glacial Record: Cambridge, UK, Cambridge Uni- record: Palaios, v. 20, p. 321–347, doi: 10.2110/palo.2004.p04-32.
versity Press, p. 480–484. Buatois, L.A., Netto, R.G., Mángano, M.G., and Balistieri, P., 2006, Extreme
Bercowski, F., Milana, J.P., and Peralta, S., 1990, La presencia de Cruziana freshwater release during the late Paleozoic Gondwana deglaciation and its
en la Formación Guandacol (Carbonífero) en la Precordillera central, impact on coastal ecosystems: Geology, v. 34, p. 1021–1024, doi: 10.1130/
departamento Jachal, San Juan, in Actas, Congreso Argentino de Paleon- G22994A.1.
tología y Bioestratigrafía, 5th, San Miguel de Tucumán, Serie Correlación Campbell, K.S.W., 1969, The Geology of New South Wales: Journal of the
Geológica, v. 7, p. 73–76. Geological Society of Australia, v. 16, p. 250–252.
Beri, A., 1987, Estudio preliminary del contenido palinológico de la perfo- Caputo, M.V., and Crowell, J.C., 1985, Migration of glacial centers across
ración n. 201 (Carbonífero Superior-Pérmico Inferior) del NE del Uru- Gondwana during Paleozoic Era: Geological Society of America
guay, in Actas, Simposio Argentino de Paleobotánica y Palinología, 7th, Bulletin, v. 96, p. 1020–1036, doi: 10.1130/0016-7606(1985)96<1020:
Buenos Aires, p. 33–36. MOGCAG>2.0.CO;2.
Beri, A., 1991, Palinologia no Neopaleozóico da bacia do Paraná na República Castro, J.C., Bortoluzzi, C.A., Caruso, F., Jr., and Krebs, A.S., 1994, Coluna
Oriental do Uruguai. Considerações bioestratigráficas e paleoecológicas White: Estratigrafia da Bacia do Paraná no Sul do Estado de Santa Catar-
(Master of Science thesis): Porto Alegre, Brazil, Universidade Federal do ina—Brasil: Florianópolis, Secretaria de Estado de Tecnologia, Energia e
Rio Grande do Sul, 96 p. Meio Ambiente, Série Textos Básicos de Geologia e Recursos Minerais
Beri, A., and Daners, G., 1994, Contenido palinológico de la Fm. San Gregorio de Santa Catarina, v. 1, 67 p.
en perforación Cerro Largo Sur, Volume 4, Pérmico Inferior, Uruguay: Catuneanu, O., 2004, Basement control on flexural profiles and the distribution
Jornadas de Paleontología de Uruguay, Montevideo, Resúmenes Amplia- of foreland facies, the Dwyka Group of the Karoo Basin, South Africa:
dos, p. 10–11. Geology, v. 32, p. 517–520, doi: 10.1130/G20526.1.
Birgenheier, L.P., Fielding, C.R., Rygel, M.G., Frank, T.D., and Roberts, J., Catuneanu, O., Hancox, J.P., and Rubidge, B.S., 1998, Reciprocal flexural
2009, Evidence for dynamic climate change on sub-106-year scales from behaviour and contrasting stratigraphies: A new basin development model
the late Paleozoic glacial record, Tamworth Belt, New South Wales, Aus- for the Karoo retroarc foreland system, South Africa: Basin Research,
tralia: Journal of Sedimentary Research, v. 79, p. 56–82, doi: 10.2110/jsr v. 10, p. 417–439, doi: 10.1046/j.1365-2117.1998.00078.x.
.2009.013. Catuneanu, O., Wopfner, H., Eriksson, P.G., Cairncross, B., Rubidge, B.S.,
Bromley, R.G., 1996, Trace fossils: Biology, taphonomy and applications: Lon- Smith, R.M.H., and Hancock, P.J., 2005, The Karoo basins of south-
don, Chapman and Hall, 361 p. central Africa: Journal of African Earth Sciences, v. 43, p. 211–253, doi:
Buatois, L.A., and del Papa, C.E., 2003, Trazas fósiles de la Formación Tarija, 10.1016/j.jafrearsci.2005.07.007.
Carbonífero Superior del norte argentino: Aspectos icnológicos de la Césari, S.N., and Gutiérrez, P.R., 2000, Palynostratigraphy of upper Paleozoic
transgresión postglacial en el oeste de Gondwana, in Buatois, L.A., and sequences in central-western Argentina: Palynology, v. 24, p. 113–146,
Mángano, M.G., eds., Icnología: Hacia una convergencia entre geología doi: 10.2113/0240113.
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 169

Closs, D., 1967, Orthocone cephalopods from the Upper Carboniferous of Eyles, C.H., Mory, A.J., and Eyles, N., 2003, Carboniferous-Permian facies
Argentina and Uruguay: Ameghiniana, v. 5, p. 123–129. and tectono-stratigraphic successions of the glacially influenced and
Closs, D., 1969, Intercalation of Goniatites in the Gondwana glacial beds of rifted Carnarvon Basin, Western Australia: Sedimentary Geology, v. 155,
Uruguay, in International Union of Geological Sciences Gondwana Sym- p. 63–86, doi: 10.1016/S0037-0738(02)00160-4.
posium, 1st, Mar del Plata, Argentina, p. 197–212. Eyles, N., 1993, Earth’s glacial record and its tectonic setting: Earth-Science
Collinson, J.W., Isbell, J.L., Elliot, D.H., Miller, M.F., Miller, J.M.G., and Reviews, v. 35, p. 1–248, doi: 10.1016/0012-8252(93)90002-O.
Veevers, J.J., 1994, Permian-Triassic Transantarctic Basin, in Veevers, Eyles, N., and Young, G.M., 1994, Geodynamic controls on glaciation in Earth
J.J., and Powell, C.McA., eds., Permian-Triassic Pangean Basins and history, in Deynoux, M., Miller, J.M.G., Domack, E.W., Eyles, N., Fair-
Foldbelts along the Panthalassan Margin of Gondwana: Geological Soci- child, I.J., and Young, G.M., eds., Earth’s Glacial Record: Cambridge,
ety of America Memoir 184, p. 173–222. UK, Cambridge University Press, p. 1–28.
Collinson, J.W., Hammer, W.R., Askin, R.A., and Elliot, D.H., 2004, Permian Eyles, N., Bonorino, G.G., Franca, A.B., Eyles, C.H., and Lopez Paulsen, O.,
to Triassic transition in the central Transantarctic Mountains, Antarctica: 1995, Hydrocarbon-bearing late Paleozoic glaciated basins of southern
Geological Society of America Abstracts with Programs, v. 36, p. 336. and central South America, in Tankard, A.J., Suárez Soruco, R., and Wel-
Cooper, M.R., and Kensley, B., 1984, Endemic South American Permian sink, H.J., eds, Petroleum Basins of South America: American Associa-
bivalve molluscs from the Ecca of South Africa: Journal of Paleontology, tion of Petroleum Geologists Memoir 62, p. 165–183.
v. 58, p. 1360–1363. Farrow, G.E., Syvitski, J.P.M., and Tunnicliffe, V., 1983, Suspended particulate
Corner, G.D., and Fjalstad, A., 1993, Spreite trace fossils (Teichichnus) in a loading on the macrobenthos in a highly turbid fjord: Knight Inlet, Brit-
raised Holocene fjord-delta, Breidvikeidet, Norway: Ichnos, v. 2, p. 155– ish Columbia: Canadian Journal of Fisheries and Aquatic Sciences, v. 40,
164, doi: 10.1080/10420949309380085. p. 273–288.
Correa da Silva, Z.C., 1978, Observações sobre o Grupo Tubarão no Rio Feder, H.M., and Keiser, G.E., 1980, Intertidal biology, in Colonell, J.M.,
Grande do Sul com especial destaque à estratigrafia da Formação Itararé: ed., Port Valdez, Alaska: Environmental Studies 1976–1979: Institute
Pesquisas, v. 9, p. 9–61. of Marine Sciences, University of Alaska, Occasional Publication 5,
da Silva, J., 1985, Bioestratigrafía preliminar del Paleozoico Superior del Uru- p. 143–233.
guay: Dirección Nacional de Geología y Minería Informe Interno, Mon- Feder, H.M., and Matheke, G.E.M., 1980, Subtidal benthos, in Colonell, J.M.,
tevideo, 14 p. ed., Port Valdez, Alaska: Environmental Studies 1976–1979: Institute
del Papa, C.E., and Martínez, L., 2001, Sedimentación lacustre glaci-dominada of Marine Sciences, University of Alaska, Occasional Publication 5,
en la Formación Tarija (Carbonífero), sierra de Aguaragüe, noroeste p. 235–324.
argentino: Revista de la Asociación Argentina de Sedimentología, v. 8, Fernandes, A.C.S., Carvalho, I.S., and Netto, R.G., 1987, Comentarios sobre os
p. 61–76. tracos fosseis do paleolago de Itu, Sao Paulo, in Proceedings, Simposio
Desjardins, P., Buatois, L.A., Limarino, C.O., and Cisterna, G., 2009, Latest Regional de Geologia, 6th, Rio Claro, Brazil, v. 1, p. 297–311.
Carboniferous-earliest Permian transgressive deposits in the Paganzo Fielding, C.R., Sliwa, R., Holcombe, R.J., and Jones, A.T., 2001, A new palaeo-
Basin of Western Argentina: Lithofacies and sequence stratigraphy of a geographic synthesis for the Bowen, Gunnedah and Sydney basins of
coastal plain to shallow-marine succession: Journal of South American eastern Australia, in Hill, K.C., and Bernecker, T., eds., Eastern Austral-
Earth Sciences, v. 28, p. 40–53, doi: 10.1016/j.jsames.2008.10.003. asian Basins Symposium 2001: A Refocused Energy Perspective for the
Desjardins, P.R., Buatois, L.A., Mángano, M.G., and Limarino, C.O., 2010, this Future: Petroleum Exploration Society of Australia Special Publication 1,
volume, Ichnology of the latest Carboniferous–earliest Permian transgres- p. 269–278.
sion in the Paganzo Basin of western Argentina: The interplay of ecology, Frakes, L.A., and Crowell, J.C., 1967, Facies and paleogeography of late Paleo-
sea-level rise, and paleogeography during postglacial times in Gondwana, zoic diamictite, Falkland Islands: Geological Society of America Bulletin,
v. 78, p. 37–58, doi: 10.1130/0016-7606(1967)78[37:FAPOLP]2.0.CO;2.
in López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial
França, A.B., 1994, Itararé Group: Gondwanan Carboniferous-Permian of the
Events and Postglacial Transgressions in Gondwana: Geological Society
Paraná Basin, Brazil, in Deynoux, M., Miller, J.M.G., Domack, E.W.,
of America Special Paper 468, doi: 10.1130/2010.2468(08).
Eyles, N., Fairchild, I.J., and Young, G.M., eds., Earth’s Glacial Record:
Dias-Fabrício, M.E., and Guerra-Sommer, M., 1989, Síntese dos estudos
Cambridge, UK, Cambridge University Press, p. 70–82.
icnológicos do Grupo Itararé no Rio Grande do Sul: Pesquisas, v. 22,
Gaigalas, A., and Uchman, A., 2004, Trace fossils from the Upper Pleistocene
p. 71–88.
varved clays S of 8 Kaunas, Lithuania: Geologija, v. 45, p. 16–26.
Díaz Martínez, E., Palmer, B.A., and Lema, J.C., 1993, The Carboniferous
Gandini, R., Netto, R.G., and Souza, P.A., 2007, Paleoicnologia e a palino-
sequence of the Northern Altiplano of Bolivia: From glacial-marine to logia dos ritmitos do Grupo Itararé na pedreira de Águas Claras (Santa
carbonate deposition, in Comptes Rendus, 12e Congrès International de Catarina, Brasil): Gaea, v. 3, p. 47–59.
la Stratigraphie et Geologie du Carbonifère et Permien, Buenos Aires, García, G.B., 1995, Palinología de la Formación El Imperial, Paleozoico Supe-
Argentina, v. 1, p. 203–222. rior, Cuenca San Rafael, Argentina, Parte I, Esporas: Ameghiniana, v. 32,
Dickins, J.M., 1961, Eurydesma and Peruvispira from the Dwyka beds of South p. 315–339.
Africa: Palaeontology, v. 4, p. 138–148. Gardiner, B.G., 1962, Namaichthys schroederi Guerich and other Palaeozoic
Dolianiti, E., 1945, Descoberta de fósseis na Formação Maricá, Estado do Rio fishes from South Africa: Palaeontology, v. 5, p. 9–21.
Grande do Sul: Mineração e Metalurgia, v. 9, p. 110. Gibbard, P.L., 1977, Fossil tracks from varved sediments near Lammi, south
Ekdale, A.A., 1985, Paleoecology of the marine endobenthos: Palaeogeography, Finland: Bulletin of the Geological Society of Finland, v. 49, p. 53–57.
Palaeoclimatology, Palaeoecology, v. 50, p. 63–81, doi: 10.1016/S0031 Gibbard, P.L., and Dreimanis, A., 1978, Trace fossils from late Pleistocene gla-
-0182(85)80006-7. cial lake sediments in southwestern Ontario, Canada: Canadian Journal of
Espejo, I., Andreis, R.R., and Mazzoni, M.M., 1996, Cuenca San Rafael, in Earth Sciences, v. 15, p. 1967–1976.
Archangelsky, S., ed., El Sistema Pérmico en la República Argentina y Gibbard, P.L., and Stuart, A.J., 1974, Trace fossils from proglacial lake sedi-
en la República Oriental del Uruguay: Córdoba, Argentina, Academia ments: Boreas, v. 3, p. 69–74.
Nacional de Ciencias de Córdoba, p. 177–185. Goso, C.A.A., 1995, Análise estratigráfica da Formação San Gregório (P) na
Eyles, C.H., and Eyles, N., 2000, Subaqueous mass flow origin for Lower borda leste da bacia Norte Uruguaia [M.S. thesis]: Rio Claro, Brazil, Uni-
Permian diamictites and associated facies of the Grant Group, Barbwire versidade Estadual Paulista, 215 p.
Terrace, Canning Basin, Western Australia: Sedimentology, v. 47, p. 343– Gostin, V.A., and Herbert, C., 1973, Stratigraphy of the upper Carboniferous
356, doi: 10.1046/j.1365-3091.2000.00295.x. and lower Permian sequence, southern Sydney Basin: Journal of the Geo-
Eyles, C.H., Vossler, S.M., and Lagoe, M.B., 1992, Ichnology of a glacially logical Society of Australia, v. 20, p. 49–70.
influenced continental shelf and slope; the late Cenozoic Gulf of Alaska Grill, H., 1997, The Permo-Carboniferous glacial to marine Karoo record in
(Yakataga Formation): Palaeogeography, Palaeoclimatology, Palaeoecol- southern Namibia; sedimentary facies and sequence stratigraphy: Berin-
ogy, v. 94, p. 193–221, doi: 10.1016/0031-0182(92)90119-P. geria, v. 19, 98 p.
Eyles, C.H., Eyles, N., and Franca, A.B., 1993, Glaciation and tectonics in an Guerra-Sommer, M., Mendez-Piccoli, A.E., and Dias-Fabricio, M.E., 1984,
active intracratonic basin; the late Palaeozoic Itararé Group, Paraná Basin, Icnofósseis em varvitos do Grupo Itararé, Permiano Inferior, bacia do
Brazil: Sedimentology, v. 40, p. 1–25, doi: 10.1111/j.1365-3091.1993 Paraná, RS, Brasil, in Memoria, Congreso Latinoamericano de Paleonto-
.tb01087.x. logia, 3rd, Mexico City, p. 130–139.
170 Buatois et al.

Gutiérrez, P.R., and Limarino, C.O., 2001, Palinología de la Formación Malan- Palaeoclimatology, Palaeoecology, v. 65, p. 115–131, doi: 10.1016/0031
zán (Carbonífero Superior), La Rioja, Argentina: Nuevos datos y consid- -0182(88)90116-2.
eraciones paleoambientales: Ameghiniana, v. 38, p. 99–118. Limarino, C.O., Césari, S.N., Net, L.I., Marenssi, S.A., Gutierrez, R.P., and
Herbert, C.T., and Compton, J.S., 2007, Depositional environments of the Lower Tripaldi, A., 2002, The Upper Carboniferous postglacial transgression
Permian Dwyka Diamictite and Prince Albert Shale inferred from the geo- in the Paganzo and Rio Blanco basins (northwestern Argentina): Facies
chemistry of early diagenetic concretions, southwest Karoo Basin, South and stratigraphic significance: Journal of South American Earth Sciences,
Africa: Sedimentary Geology, v. 194, p. 263–277, doi: 10.1016/j.sedgeo v. 15, p. 445–460, doi: 10.1016/S0895-9811(02)00048-2.
.2006.06.008. López Gamundí, O.R., 1986, Sedimentología de la Formación Tarija, Carboníf-
Hobday, D.K., and Tavener-Smith, R., 1975, Trace fossils in the Ecca of north- ero de la sierra de Aguaragüe, provincia de Salta: Revista de la Asociación
ern Natal and their palaeoenvironmental significance: Palaeontologica Geológica Argentina, v. 41, p. 334–355.
Africana, v. 18, p. 47–52. López Gamundí, O.R., 1989, Postglacial transgressions in Late Paleozoic
Isbell, J.L., Gelhar, G.A., and Seegers, G.M., 1997, Reconstruction of preglacial basins of western Argentina: A record of glacio-eustatic sea level rise:
topography using a postglacial flooding surface; Upper Paleozoic depos- Palaeogeography, Palaeoclimatology, Palaeoecology, v. 71, p. 257–270,
its, Central Transantarctic Mountains, Antarctica: Journal of Sedimentary doi: 10.1016/0031-0182(89)90054-0.
Research, v. 67, p. 264–273. López Gamundí, O.R., 1997, Glacial-postglacial transition in the Late Paleozoic
Isbell, J.L., Miller, M.F., Babcock, L.E., and Hasiotis, S.T., 2001, Ice-marginal basins of southern South America, in Martini, I.P., ed., Late Glacial and
Postglacial Environmental Changes: Quaternary, Carboniferous-Permian,
environment and ecosystem prior to initial advance of the late Palaeozoic
and Proterozoic: New York, Oxford University Press, p. 147–168.
ice sheet in the Mount Butters area of the central Transantarctic Moun-
López-Gamundí, O.R., 2010, this volume, Transgressions related to the demise
tains, Antarctica: Sedimentology, v. 48, p. 953–970, doi: 10.1046/j.1365
of the late Paleozoic Ice Age: Their sequence stratigraphic context, in
-3091.2001.00403.x.
López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial
Isbell, J.L., Miller, M.F., Wolfe, K.L., and Lenaker, P.A., 2003a, Timing of Events and Postglacial Transgressions in Gondwana: Geological Society
late Paleozoic glaciation in Gondwana: Was glaciation responsible for of America Special Paper 468, doi: 10.1130/2010.2468(01).
the development of northern hemisphere cyclothems? in Chan, M.A., López Gamundí, O., and Martínez, M., 2000, Evidence of glacial abrasion in
and Archer, A.W., eds., Extreme Depositional Environments: Mega End the Calingasta-Uspallata and western Paganzo basins, mid-Carboniferous
Members in Geologic Time: Geological Society of America Special Paper of western Argentina: Palaeogeography, Palaeoclimatology, Palaeoecol-
370, p. 5–24. ogy, v. 159, p. 145–165, doi: 10.1016/S0031-0182(00)00044-4.
Isbell, J.L., Lenaker, P.A., Askin, R.A., Miller, M.F., and Babcock, L.E., 2003b, Lord, A.R., 1990, The Pleistocene-Holocene transition in southwestern Swe-
Reevaluation of the timing and extent of late Paleozoic glaciation in Gond- den and the recognition of deglaciation effects in adjacent seas, in
wana: Role of the Transantarctic Mountains: Geology, v. 31, p. 977–980, Dowdeswell, J.A., and Scourse, J.E., eds., Glacimarine Environments,
doi: 10.1130/G19810.1. Processes and Sediments: Geological Society [London] Special Publica-
Johnson, E.W., Briggs, D.E.G., Suthren, R.J., Wright, J.L., and Tunnicliff, S.P., tion 53, p. 317–328.
1994, Non-marine arthropod traces from the subaerial Ordovician Bor- MacEachern, J.A., and Gingras, M., 2007, Recognition of brackish-water trace
rowdale Volcanic Group, English Lake District: Geological Magazine, fossil assemblages in the Cretaceous western interior seaway of Alberta,
v. 131, p. 395–406, doi: 10.1017/S0016756800011146. in Bromley, R., Buatois, L.A., Mángano, M.G., Genise, J., and Melchor,
Johnson, M.R., van Vuuren, C.J., Visser, J.N.J., and Cole, D.I., Wickens, H. R., eds., Sediment-Organism Interactions, a Multifaceted Ichnology:
de V., Christie, A.D.M., and Roberts, D.L., 1997, The Foreland Karoo Society of Economic Paleontologists and Mineralogists Special Publica-
Basin, South Africa, in Selley, R.C., ed., Sedimentary Basins of the World: tion 88, p. 149–194.
African Basins: Amsterdam, Elsevier, p. 269–317. MacEachern, J.A., and Pemberton, S.G., 1994, Ichnological aspects of incised
Johnson, S.D., Flint, S., Hinds, D., and deVille Wickens, H., 2001, Anatomy, valley fill systems from the Viking Formation of the Western Canada
geometry and sequence stratigraphy of basin floor to slope turbidite sys- Sedimentary Basin, Alberta, Canada, in Boyd, R., Zaitlin, B.A., and
tems, Tanqua Karoo, South Africa: Sedimentology, v. 48, p. 987–1023, Dalrymple, R., eds., Incised Valley Systems—Origin and Sedimentary
doi: 10.1046/j.1365-3091.2001.00405.x. Sequences: Society of Economic Paleontologists and Mineralogists Spe-
Keighley, D.G., and Pickerill, R.K., 1996, Small Cruziana, Rusophycus, and related cial Publication 51, p. 129–157.
ichnotaxa from Eastern Canada; the nomenclature debate and systematic MacEachern, J.A., Raychaudhuri, I., and Pemberton, S.G., 1992, Stratigraphic
ichnology: Ichnos, v. 4, p. 261–285, doi: 10.1080/10420949609380136. applications of the Glossifungites ichnofacies: Delineating discontinuities
Key, R.M., Tidi, J., McGeorge, I., Aitken, G., Cadmann, A., and Anscombe, J., in the rock record, in Pemberton, S.G., ed., Applications of Ichnology to
1998, The lower Karoo Supergroup geology of the southeastern part of Petroleum Exploration: Society of Economic Paleontologists and Miner-
the Gemsbok Sub-basin of the Kalahari Basin, Botswana: South African alogists, Core Workshop 17, p. 57–84.
Journal of Geology, v. 101, p. 225–236. MacEachern, J.A., Bann, K.L., Bhattacharya, J.P., and Howell, C.D., Jr., 2005,
Kling, S.A., and Reif, W.E., 1969, The Paleozoic history of amphidisc and Ichnology of deltas: Organism responses to the dynamic interplay of riv-
hemidisc sponges: New evidence from the Carboniferous of Uruguay: ers, waves, storms, and tides, in Giosan, L., and Bhattacharya, J.P., eds.,
River Deltas—Concepts, Models and Examples: Society of Economic
Journal of Paleontology, v. 43, p. 1429–1434.
Paleontologists and Mineralogists Special Publication 83, p. 49–85.
Kneller, B., Milana, J.P., Buckee, C., and al Ja’aidi, O., 2004, A depositional
Mángano, M.G., and Buatois, L.A., 2004, Ichnology of Carboniferous tide-
record of deglaciation in a paleofjord (Late Carboniferous [Pennsylva-
influenced environments and tidal flat variability in the North American
nian] of San Juan Province, Argentina): The role of catastrophic sedimen-
Midcontinent, in McIlroy, D., ed., The Application of Ichnology to Palae-
tation: Geological Society of America Bulletin, v. 116, p. 348–367, doi:
oenvironmental and Stratigraphic Analysis: Geological Society [London]
10.1130/B25242.1. Special Publication 228, p. 157–178.
Knepprath, N.E., Miller, M.F., and Isbell, J.L., 2004, Dense Permian polar Marenssi, S.A., Tripaldi, A., Limarino, C.O., and Caselli, A.T., 2005, Facies
forests with large trees; upper Buckley Formation, central Transantarc- and architecture of a Carboniferous grounding-line system from the
tic Mountains: Geological Society of America Abstracts with Programs, Guandacol Formation, Paganzo Basin, northwestern Argentina: Gond-
v. 36, p. 92. wana Research, v. 8, p. 187–202, doi: 10.1016/S1342-937X(05)71117-5.
Lermen, R., 2006, Assinaturas icnológicas em depósitos glaciogênicos do Marques-Toigo, M., 1970, Anabaculites nov. gen., a new miospore genus from
Grupo Itararé no RS [Master of science dissertation]: São Leopoldo, Uni- San Gregório Formation of Uruguay: Ameghiniana, v. 2, p. 79–82.
versidade do Vale do Rio dos Sinos, 84 p. Marques-Toigo, M., 1973a, Ammonoids x pollens and the Carboniferous or
Limarino, C.O., 1988, Paleoambientes sedimentarios y paleogeografía de la Permian age of San Gregorio Formation of Uruguay, Paraná Basin: Anais
sección inferior del Grupo Paganzo en el Sistema del Famatina: Anales da Academia Brasileira de Ciencias, v. 44, suppl., p. 237–241.
Academia Nacional de Ciencias Exactas: Físicas y Naturales, v. 39, Marques-Toigo, M., 1973b, Estudo palinológico de concreções calcárias da
p. 145–178. Formação San Gregório, NE da República Oriental do Uruguai—Bacia
Limarino, C.O., and Césari, S., 1988, Paleoclimatic significance of the lacus- do Paraná [Master of Science thesis]: Porto Alegre, Brazil, Universidade
trine Carboniferous deposits in northwest Argentina: Palaeogeography, Federal do Rio Grande do Sul, 103 p.
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 171

Marques-Toigo, M., 1974, Some new species of spores and pollens of Lower of Geological Sciences Gondwana Symposium, 5th, Wellington, New
Permian age from San Gregorio Formation in Uruguay: Anais da Aca- Zealand, p. 47–52.
demia Brasileira de Ciencias, v. 46, p. 601–616. Netto, R.G., and Goso, C.A.A., 1998, Icnología de la Formación San Gregorio
Marques-Toigo, M., Dias-Fabrício, M.E., Guerra-Sommer, M., Cazzulo- (Pérmico Inferior), en el área de los Cerros Guazunambí, Cuenca Norte
Klepzig, M., and Piccoli, A.E.M., 1989, Afloramentos da área de Trom- Uruguaya, in Actas, Congreso Uruguayo de Geología, 2nd, Punta del Este,
budo Central, Permiano Inferior, Santa Catarina: Palinologia, icnologia Uruguay, v. 1, p. 188–189.
e sedimentologia, in Anais, Congresso Brasileiro de Paleontologia, 11th, Netto, R.G., and Rossetti, D.F., 2003, Ichnology and salinity fluctuations: A
Curitiba, Brazil, v. 1, p. 125–150. case study from the Early Miocene (Lower Barreiras Formation) of São
Marshall, J.E.A., 1994, The Falkland Islands and the early fragmentation of Luís Basin, Maranhão, Brazil: Revista Brasileira de Paleontologia, v. 6,
Gondwana; implications for hydrocarbon exploration in the Falkland Pla- p. 5–18.
teau: Marine and Petroleum Geology, v. 11, p. 631–636, doi: 10.1016/0264 Netto, R.G., Buatois, L.A., Mángano, M.G., and Balistieri, P., 2007, Gyrolithes
-8172(94)90073-6. as a multipurpose burrow: An ethologic approach: Revista Brasileira de
Martin, H., and Wilczewski, N., 1970, Palaeoecology, conditions of deposition Paleontologia, v. 10, p. 157–168, doi: 10.4072/rbp.2007.3.03.
and the palaeogeography of the marine Dwyka Beds of South West Africa, Netto, R.G., Balistierin, P.R.M.N., Lavina, E.N.C., and Silveira, D.M., 2009,
in Haughton, S.H., ed., Trans-Karoo Excursion: International Union of Ichnological signatures of shallow freshwater lakes in the glacial Itararé
Geological Sciences Gondwana Symposium, 2nd, Pretoria, Symposium Group (Mafra Formation, Upper Carboniferous–Lower Permian of Paraná
Guidebook, v. 3, p. 225–232. Basin, S Brazil): Palaeogeography, Palaeoclimatology, Palaeoecology,
Martínez, M., 1993, Hallazgo de fauna marina en la Formación Guandacol v. 272, p. 240–255.
(Carbonífero) en la localidad de Agua Hedionda, San Juan, Precordillera Nogueira, M.S., and Netto, R.G., 2001a, Icnofauna da Formação Rio do Sul
Nororiental, Argentina, in Comptes Rendus, 12e Congrès International de (Grupo Itararé, Permiano da bacia do Paraná) na Pedreira Itaú-Itauna, Santa
la Stratigraphie et Geologie du Carbonifère et Permien, Buenos Aires, Catarina, Brasil: Acta Geologica Leopoldensia, v. 52/53, p. 397–406.
Argentina, v. 2, p. 291–296. Nogueira, M.S., and Netto, R.G., 2001b, A presença de Cruziana nos sedimen-
Martínez Machiavello, I.C., 1963, Microesporomorfos tipos contenidos em el tos da Formação Rio do Sul (Grupo Itararé, Permo-Carbonífero da bacia
glacial em la base del Sistema de Gondwana em Uruguay. Boletim da do Paraná) na pedreira Itaú-Itaúna, Santa Catarina, Brasil: Acta Geologica
Universidade Federal do Paraná: Geologia, v. 10, p. 1–14. Leopoldensia, v. 52/53, p. 387–396.
Martins, E.A., 1948, Fósseis marinhos na Série Maricá, Estado do Rio Grande O’Brien, P.E., Lindsay, J.F., Knauer, K., and Sexton, M.J., 1998, Sequence stra-
do Sul: Mineração e Metalurgia, v. 12, p. 237–239. tigraphy of a sandstone-rich Permian glacial succession, Fitzroy Trough,
Martins, E.A., 1951, Aviculopecten cambahyensis n. sp. do Permo-Carbonífero Canning Basin, Western Australia: Australian Journal of Earth Sciences,
do R. G. S: Boletim do Museu Nacional: Geologia, v. 13, p. 1–7. v. 45, p. 533–545, doi: 10.1080/08120099808728410.
Martins, E.A., and Sena Sobrinho, M., 1950, Novos fósseis e a idade da Forma- O’Clair, C.E., and Zimmerman, S.T., 1987, Biogeography and ecology of inter-
ção Maricá, Rio Grande do Sul: Boletim do Museu Nacional do Rio de tidal and shallow subtidal Communities, in Hood, D.W., and Zimmer-
Janeiro, v. 8, p. 1–15. man, S.T., eds., The Gulf of Alaska: Physical Environment and Biological
Resources: National Oceanic and Atmospheric Administration, Anchor-
Mason, T.R., Stanistreet, I.G., and Tavener-Smith, R., 1983, Spiral trace fossils
age, Alaska, p. 305–344.
from the Permian Ecca Group of Zululand: Lethaia, v. 16, p. 241–247,
Oliveira, E.P., 1930, Fósseis marinhos na série Itararé no Estado de Santa Catar-
doi: 10.1111/j.1502-3931.1983.tb01149.x.
ina: Anais da Academia Brasileira de Ciencias, v. 2, p. 17–21.
McLachlan, I.R., and Anderson, A., 1973, A review of the evidence for marine
Ottone, E.G., and Azcuy, C.L., 1986, El perfil de la Quebrada de La Delfina,
conditions in southern Africa during Dwyka times: Palaeontologia Afri-
provincia de San Juan: Revista de la Asociación Geológica Argentina,
cana, v. 15, p. 37–64.
v. 41, p. 124–136.
McPhie, J., 1987, The Hianana Volcanics; remnants of a Late Permian tuff ring and
Paz, C.P., Netto, R.G., and Balistieri, P., 2002, Biomecânica de decápodes e
lava flow, Coombadjha volcanic complex, northeastern New South Wales:
variações morfológicas das assinaturas icnológicas, de acordo com a
Australian Journal of Earth Sciences, v. 34, p. 417–433, doi: 10.1080/ natureza e a consistência do substrato: Paleontologia em Destaque, v. 40,
08120098708729423. p. 22.
Mendes, J.C., 1952, Invertebrés du system de Gondwana au Brèsil, in Report, Paz, C.P., Netto, R.G., and Balistieri, P., 2003, Decapod biomechanics and
International Geological Congress, 19th, Algiers, p. 302–307. morphological variations of the ichnological signatures according to the
Mezzalira, S., 1956, Novas ocorrências de camadas marinhas permo- nature and consistency of the substrate: Latin American Sedimentological
carboníferas no estao de São Paulo: Boletim da Sociedade Brasileira de Congress, 3rd, Belém, Brazil, p. 179.
Geologia, v. 5, p. 61–69. Paz, C.P., Netto, R.G., and Balistieri, P., 2004, Paleoecologic and paleoenvi-
Milani, E.J., 1997, Evolução tectono-estratigráfica da bacia do Paraná e seu rela- ronmental implications of distinct preservation of Diplichnites gouldi:
cionamento com a geodinâmica fanerozóica do Gondwana sul-ocidental Ichnia 2004, International Congress on Ichnology, 1st, Trelew, Argentina,
[Doctor of Science thesis]: Porto Alegre, Brazil, Universidade Federal do Abstracts, p. 66.
Rio Grande do Sul, 255 p. Pazos, P.J., 2000, Trace fossils and facies in glacial to postglacial deposits from
Miller, G.D., 1986, The sediments and trace fossils of the Rough Rock Group the Paganzo basin (Late Carboniferous), central Precordillera, Argentina:
on Cracken Edge, Derbyshire: Mercian Geologist, v. 10, p. 189–202. Ameghiniana, v. 37, p. 23–38.
Miller, J.M.G., 1989, Glacial advance and retreat sequences in Permo-Carbon- Pazos, P.J., 2002a, Palaeoenvironmental framework of the glacial-postglacial
iferous section, central Transantarctic Mountains: Sedimentology, v. 36, transition (late Paleozoic) in the Paganzo-Calingasta Basin (southern South
p. 419–430, doi: 10.1111/j.1365-3091.1989.tb00617.x. America) and the Great Karoo-Kalahari Basin (southern Africa): Ichno-
Miller, M.F., and Collinson, J.W., 1994, Late Paleozoic post-glacial inland sea logical implications: Gondwana Research, v. 5, p. 619–640, doi: 10.1016/
filled by fine-grained turbidites: Mackellar Formation, Central Transant- S1342-937X(05)70634-1.
arctic Mountains, in Deynoux, M., Miller, J.M.G., Domack, E.W., Eyles, Pazos, P.J., 2002b, The Late Carboniferous glacial to postglacial transition: Facies
N., Fairchild, I.J., and Young, G.M., eds., Earth’s Glacial Record: Cam- and sequence stratigraphy, Western Paganzo Basin, Argentina: Gondwana
bridge, UK, Cambridge University Press, p. 215–233. Research, v. 5, p. 467–487, doi: 10.1016/S1342-937X(05)70736-X.
Miller, M.F., and Isbell, J.L., 2010, this volume, Reconstruction of a high-latitude, Pazos, P.J., di Pasquo, M., and Amenabar, C.R., 2007, Trace fossils of the gla-
post-glacial lake: Mackellar Formation (Permian), Transantarctic Moun- cial to postglacial transition in the El Imperial Formation (Upper Car-
tains, in López-Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic boniferous), San Rafael Basin, Argentina, in Bromley, R., Buatois, L.A.,
Glacial Events and Postglacial Transgressions in Gondwana: Geological Mángano, M.G., Genise, J., and Melchor, R., eds., Sediment-Organism
Society of America Special Paper 468, doi: 10.1130/2010.2468(09). Interactions: A Multifaceted Ichnology: Society of Economic Paleontolo-
Mitchell, C., Taylor, G.K., Cox, K.G., and Shaw, J., 1986, Are the Falkland gists and Mineralogists Special Publication 88, p. 137–147.
Islands a rotated microplate?: Nature, v. 319, p. 131–134, doi: 10.1038/ Pemberton, S.G., and Wightman, D.M., 1992, Ichnological characteristics of
319131a0. brackish water deposits, in Pemberton, S.G., ed., Applications of Ichnol-
Mones, A., and Figueiras, A., 1980, A geo-paleontological synthesis of the ogy to Petroleum Exploration: Society of Economic Paleontologists and
Gondwana formations of Uruguay, in Proceedings, International Union Mineralogists, Core Workshop 17, p. 141–167.
172 Buatois et al.

Pemberton, S.G., Spila, M., Pulham, A.J., Saunders, T., MacEachern, J.A., Savage, N.M., 1971, A varvite ichnocoenosis from the Dwyka Series of Natal:
Robbins, D., and Sinclair, I.K., 2001, Ichnology and sedimentology of Lethaia, v. 4, p. 217–233, doi: 10.1111/j.1502-3931.1971.tb01290.x.
shallow to marginal marine systems, Ben Nevis and Avalon Reservoirs, Scheffler, K., Buehmann, D., and Schwark, L., 2006, Analysis of late Palaeo-
Jeanne d’Arc Basin: Geological Association of Canada, Short Course zoic glacial to postglacial sedimentary successions in South Africa by
Notes, v. 15, 343 p. geochemical proxies; response to climate evolution and sedimentary envi-
Peralta, S.H., and Milana, J.P., 1999, Trazas fósiles del Carbonífero Medio- ronment: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 240,
Superior de la Precordillera de San Juan, su correlación con un evento p. 184–203, doi: 10.1016/j.palaeo.2006.03.059.
glacial, in Actas Congreso Geológico Argentino, 14th, Salta, Argentina, Schneider, R.L., Muehlmann, H., Tommazi, E., Medeiros, R.A., Daemon, R.F.,
v. 1, p. 363–366. and Nogueira, A.A., 1974, Revisão estratigráfica da bacia do Paraná,
Pinto, I.D., 1947, Novos fósseis na Formação Maricá: Ciências e Letras, v. 1, in Anais, Congresso Brasileiro de Geologia, 27th, Porto Alegre, Brazil,
p. 9. Sociedade Brasileira de Geologia, v. 1, p. 41–66.
Pinto, I.D., 1949, Contribuição ao conhecimento de novos fósseis na Formação Scott, E.D., and Bouma, A.H., and Wickens, H. de V., 2000, Influence of tec-
Maricá (Afl. Budó): Faculdade de Filosofia, Universidade Federal do Rio tonics on submarine fan deposition, Tanqua and Laingsburg subbasins,
Grande do Sul, Porto Alegre, Brazil, Publication 2, p. 1–6. South Africa, in Bouma, A.H., and Stone, J., eds., Fine-Grained Turbid-
Pinto, I.D., 1955, Série Maricá, Camaquan e Formação Teixeira Soares no Rio ite Systems: American Association of Petroleum Geologists Memoir 72,
Grande do Sul. Histórico, idade e correlação: Boletim do Instituto de p. 47–56.
Ciências Naturais, v. 2, p. 1–18. Sepkoski, J.J., 2002, A compendium of fossil marine animal genera: Bulletins
Pinto, I.D., 2000, Insetos fósseis, in Holz, M., and De Ros, L.F., eds., Paleonto- of American Paleontology, v. 363, p. 560.
logia do Rio Grande do Sul: Universidade Federal do Rio Grande do Sul/ Shaw, J., and Lesemann, J.E., 2003, Subglacial outburst floods and extreme
Centro de Investigações do Gondwana, Porto Alegre, Brazil, p. 131–140. sedimentary events in the Labrador Sea, in Chan, M.A., and Archer, A.W.,
Pinto, I.D., and Purper, I., 2000, Escolecodontes—Dentes de vermes, in Holz, eds., Extreme Depositional Environments: Mega End Members in Geo-
M., and De Ros, L.F., eds., Paleontologia do Rio Grande do Sul: Univer- logic Time: Geological Society of America Special Paper 370, p. 25–41.
sidade Federal do Rio Grande do Sul/Centro de Investigações do Gond- Smith, R.M.H., Eriksson, P.G., and Botha, W.J., 1993, A review of the stratigra-
wana, Porto Alegre, Brazil, p. 126–130. phy and sedimentary environments of the Karoo-aged basins of Southern
Pinto, I.D., and Sedor, F.A., 2000, A new Upper Carboniferous blattoid from Africa: Journal of African Earth Sciences, v. 16, p. 143–169, doi: 10.1016/
Mafra Formation, Itararé Group, Paraná Basin, Brazil: Pesquisas, v. 27, 0899-5362(93)90164-L.
no. 2, p. 45–48. Stanistreet, I.G., Le Blanc Smith, G., and Cadle, A.B., 1980, Trace fossils as
Piper, D.J.W., Shaw, J., and Skene, K.I., 2007, Stratigraphic and sedimento- sedimentological and palaeoenvironmental indices in the Ecca Group
logical evidence for late Wisconsinan sub-glacial outburst floods to Lau- (Lower Permian) of the Transvaal: Transactions, Geological Society of
rentian Fan: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 246, South Africa, v. 83, p. 333–344.
p. 101–119, doi: 10.1016/j.palaeo.2006.10.029.
Stanley, S.M., and Powell, M.G., 2003, Depressed rates of origination and
Powell, M.G., 2005, Climatic basis for sluggish macroevolution during the late
extinction during the late Paleozoic ice age: A new state for the global
Paleozoic ice age: Geology, v. 33, p. 381–384, doi: 10.1130/G21155.1.
marine ecosystem: Geology, v. 31, p. 877–880, doi: 10.1130/G19654R.1.
Preciozzi, F., Spoturno, J., Heinzen, W., and Rossi, P., 1988, Memoria expli-
Starck, D., and del Papa, C., 2006, The northwestern Argentina Tarija Basin:
cativa y carta geológica del Uruguay (1:500.000): Dirección Nacional de
Stratigraphy, depositional systems, and controlling factors in a glaciated
Geología y Minería, Montevideo, 90 p.
basin: Journal of South American Earth Sciences, v. 22, p. 169–184, doi:
Ramos, V.A., 1988, The tectonics of the Central Andes; 30° to 33° S latitude,
10.1016/j.jsames.2006.09.013.
in Clark, S.P., ed., Processes in Continental Lithospheric Deformation:
Geological Society of America Special Paper 218, p. 31–54. Starck, D., Gallardo, E., and Schulz, A., 1993, Neopaleozoic stratigraphy of the
Richter, M., 2000, Peixes fósseis do Rio Grande do Sul, in Holz, M., and De Sierras Subandinas Occidentales and Cordillera Oriental Argentina; with
Ros, L.F., eds., Paleontologia do Rio Grande do Sul: Universidade Fed- comments on the southern border of the Tarija basin, in Comptes Rendus,
eral do Rio Grande do Sul/Centro de Investigações do Gondwana, Porto 12e Congrès International de la Stratigraphie et Geologie du Carbonifère
Alegre, Brazil, p. 162–175. et Permien, Buenos Aires, Argentina, v. 2, p. 353–372.
Rilett, M.H.P., 1963, A fossil cephalopod from the middle Ecca beds in the Klip Syvitski, J.P.M., Burell, D.C., and Skei, J.M., 1987, Fjords: Processes and Prod-
river coalfield near Dundee, Natal: Transactions of the Royal Society of ucts: New York, Springer, 379 p.
South Africa, v. 37, p. 73–74. Tavener-Smith, R., and Mason, T.R., 1985, The Vryheid Formation (Middle
Rocha-Campos, A.C., 1970, Moluscos permianos da Formação Rio Bonito Ecca) at Tugela Ferry: A postscript to earlier work: Suid-Afrikaanse Tyd-
(Subgrupo Guatá) Santa Catarina: Boletim da Divisão de Geolgia e Min- skrif vir Wetenskap, v. 81, p. 275–279.
eralogia, Departmento Nacional da Produção Mineral, v. 251, 58 p. Taylor, G.K., and Shaw, J., 1989, The Falkland Islands; new palaeomagnetic
Rocha-Campos, A.C., 1981a, Late Palaeozoic tillites of the Sergipe-Alagoas data and their origin as a displaced terrane from Southern Africa: Geo-
Basin, Rondônia and Mato Grosso, Brazil, in Hambrey, M.J., and Har- physical Monograph, v. 50, p. 59–72.
land, W.B., eds., Earth’s Pre-Pleistocene Glacial Record: Cambridge, UK, Teichert, C., and Rilett, M., 1974, Revision of Permian Ecca Series cephalo-
Cambridge University Press, p. 838–841. pods, Natal, South Africa: University of Kansas Paleontological Contri-
Rocha-Campos, A.C., 1981b, Late Palaeozoic “Serie Tilitica” of Mozambique, butions, v. 74, p. 1–8.
in Hambrey, M.J., and Harland, W.B., eds., Earth’s Pre-Pleistocene Gla- Tomazelli, L.J., and Soliani, E., Jr., 1982, Evidências de atividade glacial no
cial Record: Cambridge, UK, Cambridge University Press, p. 52–54. Paleozóico Superior no Rio Grande do Sul, in Anais, Congresso Brasil-
Rubidge, B.S., Hancox, P.J., and Catuneanu, O., 2000, Sequence analysis of eiro de Geologia, 32nd, Salvador, Brazil, Sociedade Brasileira de Geolo-
the Ecca-Beaufort contact in the southern Karoo of South Africa: South gia, v. 4, p. 1378–1391.
African Journal of Geology, v. 103, p. 81–96, doi: 10.2113/103.1.81. Trewin, N.H., 2000, The ichnogenus Undichna, with examples from the Perm-
Ruedmann, R., 1929, Fossils from the Permian tillite of São Paulo, Brazil and ian of the Falkland Islands: Palaeontology, v. 43, p. 979–997, doi: 10.1111/
their bearing on the origin of tillite: Geological Society of America Bul- 1475-4983.00158.
letin, v. 40, p. 243. Trewin, N.H., MacDonald, D.I.M., and Thomas, C.G.C., 2002, Stratigraphy and
Santos, P.R. dos, 1987, Fácies e evolução paleogeográfica do Subgrupo Itararé/ sedimentology of the Permian of the Falkland Islands: Lithostratigraphic
Grupo Aquidauana (neopaleozóico) na bacia do Paraná, Brazil [Doctor and palaeoenvironmental links with South Africa: Journal of the Geologi-
of science thesis]: São Paulo, Instituto de Geociências, Universidade de cal Society [London], v. 159, p. 5–19, doi: 10.1144/0016-764900-089.
São Paulo, 128 p. Tripsanas, E.K., Bryant, W.R., Slowey, N.C., Bouma, A.H., Karageorgis, A.P.,
Santos, P.R. dos, Rocha-Campos, A.C., and Canuto, J.R., 1996, Patterns of and Berti, D., 2007, Sedimentological history of Bryant Canyon area, north-
Late Paleozoic deglaciation in the Paraná Basin, Brazil: Palaeogeography, west gulf of Mexico, during the last 135 kyr (marine isotope stages 1–6);
Palaeoclimatology, Palaeoecology, v. 125, p. 165–184, doi: 10.1016/S0031 a proxy record of Mississippi River discharge: Palaeogeography, Palaeo-
-0182(96)00029-6. climatology, Palaeoecology, v. 246, p. 137–161, doi: 10.1016/j.palaeo
Savage, N.M., 1970, A preliminary note on arthropod trace fossils from the .2006.10.032.
Dwyka series, in Proceedings, International Union of Geological Sci- Trosdtorf, I., Jr., Rocha-Campos, A.C., dos Santos, P.R., and Tomio, A., 2005,
ences Gondwana Symposium, 2nd, Pretoria, South Africa, p. 627–635. Origin of Late Paleozoic, multiple, glacially striated surfaces in northern
Ichnology of late Paleozoic postglacial transgressive deposits in Gondwana 173

Paraná Basin (Brazil): Some implications for the dynamics of the Paraná Visser, J.N.J., 1990, Glacial bedforms at the base of the Permo-Carboniferous
glacial lobe: Sedimentary Geology, v. 181, p. 59–71, doi: 10.1016/j.sedgeo Dwyka Formation along the western margin of the Karoo Basin, South
.2005.07.006. Africa: Sedimentology, v. 37, p. 231–245, doi: 10.1111/j.1365-3091.1990
Turner, B.R., Stanistreet, I.G., and Whateley, M.K.G., 1981, Trace fossil and .tb00957.x.
palaeoenvironments in the Ecca Group of the Nongoma Graben, northern Visser, J.N.J., 1992, Deposition of the Early to Late Permian Whitehill For-
Zululand, South Africa: Palaeogeography, Palaeoclimatology, Palaeo- mation during a sea level highstand in a juvenile foreland basin: South
ecology, v. 36, p. 113–123, doi: 10.1016/0031-0182(81)90053-5. African Journal of Geology, v. 95, p. 181–193.
Uchman, A., Gaigalas, A., and Kazakauskas, V., 2009, Trace fossils from Late Visser, J.N.J., 1995, Post-glacial Permian stratigraphy and geography of south-
Pleistocene lacustrine varve sediments in eastern Lithuania: Palaeogeogra- ern and central Africa: Boundary conditions for climatic modelling:
phy, Palaeoclimatology, Palaeoecology, v. 272, p. 199–211, doi: 10.1016/ Palaeogeography, Palaeoclimatology, Palaeoecology, v. 118, p. 213–243,
j.palaeo.2008.08.003. doi: 10.1016/0031-0182(95)00008-3.
van Dijk, D.E., Hobday, D.K., and Tankard, A.J., 1978, Permo-Triassic lacus- Visser, J.N.J., 1996, Controls on Early Permian shelf deglaciation in the Karoo
trine deposits in the Eastern Karoo Basin, Natal, South Africa, in Matter, Basin of South Africa: Palaeogeography, Palaeoclimatology, Palaeoecol-
A., and Tucker, M.E., eds., Modern and Ancient Lake Sediments: Interna- ogy, v. 125, p. 129–139, doi: 10.1016/S0031-0182(96)00027-2.
tional Association of Sedimentologists Special Publication 2, p. 225–239. Visser, J.N.J., 1997, A review of the Permo-Carboniferous glaciation in Africa,
Veevers, J.J., and Powell, C.M., 1987, Late Paleozoic glacial episodes in Gond- in Martini, I.P., ed., Late Glacial and Postglacial Environmental Changes:
wanaland reflected in transgressive-regressive depositional sequences in Quaternary, Carboniferous-Permian, and Proterozoic: New York, Oxford
Euramerica: Geological Society of America Bulletin, v. 98, p. 475–487, University Press, p. 169–191.
doi: 10.1130/0016-7606(1987)98<475:LPGEIG>2.0.CO;2. Visser, J.N.J., and Young, G.M., 1990, Major element geochemistry and paleo-
Vergel, M.M., Buatois, L.A., and Mángano, M.G., 1993, Primer registro Pal- climatology of the Permo-Carboniferous glacigene Dwyka Formation
inológico en el Carbonífero Superior del margen norte de la Cuenca and post-glacial mudrocks in Southern Africa: Palaeogeography, Palaeo-
Paganzo, Los Jumes, Catamarca, Argentina, in Comptes Rendus, 12e climatology, Palaeoecology, v. 81, p. 49–57, doi: 10.1016/0031-0182
Congrès International de la Stratigraphie et Geologie du Carbonifère et (90)90039-A.
Permien, Buenos Aires, Argentina, v. 1, p. 213–227. Walter, H., and Suhr, P., 1998, Lebesspuren aus kaltzeitlichen Bändersedi-
Vesely, F.F., and Assine, M.L., 2006, Deglaciation sequences in the Permo- menten des Quartärs: Abhandlungen des Staatlichen Museums für Miner-
Carboniferous Itararé Group, Paraná Basin, southern Brazil: Journal of alogie und Geologie zu Dresden, v. 43/44, p. 311–328.
South American Earth Sciences, v. 22, p. 156–168, doi: 10.1016/j.jsames Wickens, H. de V., 1996, Die stratigraphie en sedimentologie van die Ecca
.2006.09.006. Groep wes van Sutherland: Council for Geosciences, Pretoria, Bulletin
Viljoen, J.H.A., 1994, Sedimentology of the Collingham Formation, Karoo 107, 49 p.
Supergroup: South African Journal of Geology, v. 97, p. 167–183. Wopfner, H., and Casshyap, S.M., 1997, Transition from freezing to subtropi-
Virtasalo, J.J., Kotilainen, A.T., and Gingras, M.K., 2006, Trace fossils as indi- cal climates in the Permo-Carboniferous of Afro-Arabia and India, in
cators of environmental change in Holocene sediments of the Archipelago Martini, I.P., ed., Late Glacial and Postglacial Environmental Changes:
Sea, northern Baltic Sea: Palaeogeography, Palaeoclimatology, Palaeo- Quaternary, Carboniferous-Permian, and Proterozoic: New York, Oxford
ecology, v. 240, p. 453–467, doi: 10.1016/j.palaeo.2006.02.010. University Press, p. 192–212.
Visser, J.N.J., 1983, Glacial-marine sedimentation in the Late Paleozoic Karoo Ybert, J.P., and Marques-Toigo, M., 1970, Polarisaccites nov. gen: Revue Pol-
Basin, Southern Africa, in Molnia, B.F., ed., Glacial-Marine Sedimenta- len et Spores, v. 12, p. 469–481.
tion: New York, Plenum, p. 667–701.
Visser, J.N.J., 1987, The palaeogeography of part of south-western Gond-
wana during the Permo-Carboniferous glaciations: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 61, p. 205–219, doi: 10.1016/0031
-0182(87)90050-2. MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

Ichnology of the latest Carboniferous–earliest Permian


transgression in the Paganzo Basin of western Argentina:
The interplay of ecology, sea-level rise, and paleogeography
during postglacial times in Gondwana

Patricio R. Desjardins
Luis A. Buatois
M. Gabriela Mángano
Department of Geological Sciences, University of Saskatchewan, Saskatoon, Saskatchewan, S7N 5E2, Canada

Carlos O. Limarino
CONICET (Consejo Nacional de Investigaciones Científicas y Técnicas) and Departamento de Ciencias Geológicas,
Universidad de Buenos Aires, Pabellón 2, Ciudad Universitaria, C1428EHA, Buenos Aires, Argentina

ABSTRACT

The late Paleozoic climatic evolution of Gondwana can be traced by analyzing


the benthic ecology of its coastal environments as revealed by their ichnologic con-
tent. Latest Carboniferous–earliest Permian transgressive deposits occur in the lower
member of the Tupe Formation, Paganzo Group, western Argentina. Three different
trace-fossil assemblages are present as part of two complete depositional sequences.
Trace-Fossil Assemblage 1 consists of Treptichnus pollardi and Helminthopsis abeli,
which occur in fine-grained heterolithic facies. This assemblage characterizes a
subaqueous freshwater substrate in a flood-plain environment. Trace-Fossil Assem-
blage 2, consisting of Halopoa isp., Palaeophycus crenulatus and Planolites montanus,
occurs in thin-bedded, tabular sandstone. The tracemakers inhabited a low-energy
distal-bay environment dominated by background sedimentation with sporadic storm
episodes. The trace fossils represent the activity of post-storm colonizers. Trace-Fossil
Assemblage 3 is monospecific, comprising only Rhizocorallium commune preserved at
the interface between a sandstone bed and the overlying mudstone. The tracemakers
inhabited the overlying muddy substrate in a low-energy distal-bay environment and
burrowed down into the sediment, expanding laterally at the top of the underlying
sandstone layer. Trace-Fossil Assemblages 2 and 3 do not resemble typical marine
ichnocoenoses and can be considered a depauperate Cruziana ichnofacies, suggest-
ing brackish-water conditions in a restricted marine embayment. The ichnofauna
associated with the latest Carboniferous–earliest Permian transgression of the Tupe
Formation is compared with that in the older (early Late Carboniferous) postglacial
transgression recorded in the underlying Guandacol Formation. The latter reflects

Desjardins, P.R., Buatois, L.A., Mángano, M.G., and Limarino, C.O., 2010, Ichnology of the latest Carboniferous–earliest Permian transgression in the Paganzo
Basin of western Argentina: The interplay of ecology, sea-level rise, and paleogeography during postglacial times in Gondwana, in López-Gamundí, O.R., and
Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana: Geological Society of America Special Paper 468, p. 175–192, doi:
10.1130/2010.2468(08). For permission to copy, contact [email protected]. ©2010 The Geological Society of America. All rights reserved.

175
176 Desjardins et al.

freshwater conditions related to an extreme meltwater influx coming from retreating


glaciers in fjord environments. In contrast, the latest Carboniferous–earliest Permian
transgression in the Paganzo Basin occurred in a confined, brackish-water embay-
ment, but away from the direct influence of meltwater discharges.

INTRODUCTION (López Gamundí, 1997): (1) Late Devonian–Early Carbonifer-


ous (Solimões and Amazonas Basins in Brazil, and Lake Titi-
Postglacial transgressions occur as a response to sea-level rise caca region of Bolivia); (2) early Late Carboniferous (Paganzo
induced by the melting of ice masses. Thus, a salinity gradient is and Calingasta-Uspallata Basins of western Argentina and Tarija
established within postglacial seas, controlled by the distance to Basin of northwest Argentina and southern Bolivia); and (3) Late
freshwater discharge areas near to ice masses. However, salinity Carboniferous–Early Permian (various basins in Brazil, Uruguay,
levels at a particular coastal environment can be also controlled by Paraguay, Antarctica, South Africa, India, and Australia). Each of
the local paleogeography. Salinity exerts a significant control on the these periods was followed by a postglacial sea-level rise, the
distribution, abundance, and type of organisms in marginal-marine establishment of fjords and lakes, and a subsequent climatic
settings (Remane and Schlieper, 1971; McLusky, 1989). Accord- amelioration (e.g., Limarino and Césari, 1988; López Gamundí,
ingly, the ecology of the associated biota is highly influenced by 1989; Buatois and Mángano, 1994, 1995a; Limarino et al., 2002).
freshwater discharges coming from the melting ice (Buatois et al., The Paganzo Basin is a pericratonic foreland basin fringed on
2006, this volume). Trace fossils have proved to be a useful tool in its western margin by the Protoprecordillera, a positive element
determining paleosalinity levels (e.g., Pemberton and Wightman, that divided the drainage between the Calingasta-Uspallata and
1992; MacEachern and Pemberton, 1994; Mángano and Buatois, Paganzo Basins. However, marine deposits within the Paganzo
2004; Buatois et al., 2005; MacEachern and Gingras, 2007). Dur- Basin indicate the existence of a breach in the Protoprecordillera
ing the late Paleozoic, Gondwana underwent three glacial and during the Late Carboniferous–Early Permian (López Gamundí
deglaciation events recorded in different basins of South America, et al., 1992).
South Africa, Antarctica, Australia, and India. These events had a Our studied sections are located in the surroundings of Cuesta
major impact on coastal environments; deglaciation was accom- de Huaco, San Juan province (Fig. 1), where three formations are
panied by sea-level rises that triggered transgressive phases over recognized in the Paganzo Group: Guandacol, Tupe, and Patquía
Gondwana (López Gamundí, 1989; Limarino et al., 2002). (Fig. 2) (Limarino et al., 1986). The Guandacol Formation records
The Paganzo Group of western Argentina in Cuesta de a transition from a glacial to a postglacial period (Limarino et al.,
Huaco and adjacent areas provides an opportunity to analyze and 2002; Pazos, 2002a). Coarse-grained diamictite formed in sub-
compare the ichnology of two transgressive episodes. The first aqueous outwash environments (Limarino et al., 2002; Marenssi
episode, recorded within the Guandacol Formation, is a trans- et al., 2005) are replaced upward by mudstone and thinly lami-
gression influenced by freshwater discharges from ice melting nated deposits containing dropstones. The upper interval of the
in a fjord environment (Limarino et al., 2002; Buatois and Mán- Guandacol Formation mainly consists of sandstone and records
gano, 2003; Buatois et al., 2006). The second transgressive epi- the progradation of a Gilbert-type delta (Limarino et al., 2002).
sode recorded, in the Tupe Formation, is contemporaneous with The Tupe Formation unconformably overlies the Guandacol
deglaciation phases in eastern basins in Gondwana, but does not Formation and consists of fluvial deposits punctuated by a short
show any direct influence or connection to melting ice masses. marine interval at its lower part (Cuerda, 1965; Limarino et al.,
This paper analyzes the transgressive deposits within the 1986; Pazos, 1994; Desjardins et al., 2009). In particular, Desjar-
lower member of the Tupe Formation and focuses on (1) ich- dins et al. (2009) suggested the existence of a punctuated trans-
nologic aspects of the latest Carboniferous–earliest Permian gression, in which a transgressive coastal plain was interrupted
transgression; (2) environmental and paleogeographic controls by the establishment of a braided-fluvial system, which later
affecting the ecology of this transgression, and (3) the compari- evolved into a marginal-marine embayment. This marine interval
son of the ichnofauna of the latest Carboniferous–earliest Perm- has a limited distribution in the basin (Fig. 3), in contrast to the
ian transgression with that of the older, postglacial transgression older (Namurian–early Westphalian) postglacial transgression.
(early Late Carboniferous) recorded in the Guandacol Formation The Patquía Formation consists of conglomerate, sandstone, and
in the same region. evaporite recording fluvial, eolian, and playa-lake environments
in an arid to semiarid setting (Limarino et al., 1986).
GEOLOGIC SETTING
DEPOSITIONAL SETTING AND SEQUENCE
Like other late Paleozoic successions of Gondwana, the STRATIGRAPHY OF THE STUDIED SECTIONS
Paganzo Group records a complex paleoclimatic evolution, as a
response to the movement of the supercontinent over high lati- The lower interval of the Tupe Formation was studied in
tudes. Three glacial events have been identified in Gondwana detail in the localities of Quebrada La Delfina and Mina La
Paganzo Basin of western Argentina 177

Argentina Qu N
eb
N ra
da
La
De
San lfi
Juan na

?
San Juan
1 Quebrada La Delfina
2 Mina La Cienaga
C H I L E

40
1
2 Huaco
150
150 San Jose de
Jachal

River
436
412 40
San Agustín del
Valle Fertil 510

Seco
Calingasta 12
San Juan
141
412
er
Riv
La Cienaga
40
a co
Hu
20 del Vallecito

Key 40

San Juan Fm

Guandacol Fm Mina
La Cienaga
Patquia Fm

Tupe Fm
1 km

Figure 1. Location of the studied outcrops (Mina La Ciénaga and Quebrada la Delfina) and simplified geology of the
Cuesta de Huaco area.

Ciénaga (Figs. 1 and 4). Sixteen facies grouped in seven facies delta-front deposits included within the Guandacol Formation.
associations have been identified and described within two A detailed sedimentologic and sequence-stratigraphic analysis of
complete depositional sequences (Tables 1 and 2). The envi- these deposits has been presented by Desjardins et al. (2009).
ronmental framework of MacEachern and Gingras (2007) for
bay successions is adopted in this study. These authors distin- Sequence 1
guished restricted bays that have limited or intermittent con-
nection to the open sea, from open bays that have virtually Lowstand Systems Tract
unrestricted connection to the open sea. Transgressive deposits Sheet-like, fining-upward sandstone packages, commonly
in the Tupe Formation are interpreted to record deposition in a displaying lateral-accretion surfaces (Facies Association I) are
restricted bay. interpreted as high-sinuosity channels. Facies Association I
The contact between the Guandacol Formation and the Tupe also includes thinly interbedded sandstone, siltstone, mudstone,
Formation corresponds to a sequence boundary (SB). Coastal- and coal facies, representing flood plains, peat, and associated
plain deposits of the Tupe Formation unconformably overly water bodies.
178 Desjardins et al.

Transgressive Systems Tract


Patquia Fm Toward the top of the deposits included within Facies Asso-
ciation I, an increase in coal thickness from 1–4 cm to 1 m is
Fluvial - Eolian observed. The thicker coal bed at the top indicates an increase in
Playa Lake accommodation for peat production. The thicker coal beds may
reflect a rise in the water table within a retreating shoreline. As
in Mina La Ciénaga, coastal-lagoon deposits cap a coastal-plain
p

interval, indicating that the upper parts of Facies Association I and


Facies Association II are part of the transgressive systems tract.
u

The transgressive surface is located at the base of the first thick


coal bed included in Facies Association I (Fig. 4). The establish-
ment of water bodies toward the top of this systems tract (Facies
o

Association II) indicates that the water table rise exceeded the
sediment influx and peat production (Bohacs and Suter, 1997).
Anastomosing Rivers The maximum flooding surface is located within the coastal-
F m
r

lagoon facies of Mina La Ciénaga (Fig. 4).


The trace fossils present in sequence 1 occur toward the top
G

of the lowstand systems tract and the base of the transgressive


systems tract. This ichnofauna includes Helminthopsis abeli and
Treptichnus pollardi (Trace-Fossil Assemblage 1).
T u p e

Marine Embayment
Highstand Systems Tract
Fine- to medium-grained, thickening-upward tabular sand-
o

stone packages interbedded with chloritic mudstone (Facies Asso-


Braided Plain ciation II) record progradation of clastic wedges into a water body.
The chloritic composition indicates that the lagoon was periodi-
z

cally connected with the adjacent embayment, because chlorite


Coastal Plain dominates in marine intervals that have the Precordillera as their
n

main source area (Net et al., 2002). Sandstone deposits are inter-
preted as river-fed quasi-steady turbidity currents. Association with
small mountain rivers and the brackish-water nature of the lagoon
Delta Front
a

promoted formation of hyperpycnal flows. This systems tract is


Guandacol Fm

bounded at the top by a sharp unconformity of regional extent and


sequence-boundary hierarchy, which represents a change from
g

Prodelta coastal-plain to braided alluvial-plain settings (Fig. 4).

Sequence 2
a

Lowstand Systems Tract


Pebbly and coarse-grained to medium-grained amalgamated
P

Fjord Embayment
sandstone facies (Facies Association III), representing multi-
story channels and longitudinal bars, are preserved with variable
Subaqueous Outwash thicknesses (0.4–2 m) in the studied localities. Lowstand deposits
are relatively thin in Mina La Ciénaga compared to those in Que-
mudstone
brada La Delfina, indicating that sediment bypass took place at
cg.
VF F M C VC
sandstone
this locality. However, the Quebrada La Delfina lowstand depos-
its are up to 50 m thick, suggesting differential accommodation
Dropstone Trough cross- Planar cross-
stratification stratification potential along depositional strike (Fig. 4). A 1-m-thick coal bed
Current ripple Wave ripple toward the top of this facies association indicates an increase in
Diamictite
cross-lamination cross-lamination accommodation and water table (Fig. 4).
Planar-lamination Coal
Transgressive Systems Tract
Figure 2. General stratigraphy and sedimentary environments of The lower portion of this systems tract shows different facies
the Paganzo Group in the western margin of the Paganzo Basin. associations in the two localities. The transgressive surface in
Paganzo Basin of western Argentina 179

Argentina 67° W Fluvio-deltaic


deposits
N
N Shallow- marine
and bay deposits
?
Shoreline at maximum
transgression during
the earliest Permian

Río del
Peñon

EM
SIN

ST
Las Gredas

SY
NCO BA

29° S

NA
nt
e

ATI
ayme
-marin

FAM
La Rioja
emb

Guandacol
Open
RÍO BLA

ricted

Study
sta
Rest

Sañ ierra

Cortaderas area
oga

Huaco
S

Malimán

?
C H I L E

S I E R R A S PA M P E A N A S
PAGANZO BASIN Figure 3. Paleogeographic map of the
Olta Paganzo Basin, and spatial distribu-
Malanzán tion of the latest Carboniferous–earliest
Sierra Permian transgression.
de
Malimán

SIERRA
San Juan DE
? PIE DE
PA L O

32° 30' S

0 100
Mendoza
Carboniferous Permian

Latest Carboniferous - earliest Permian


transgression

Namurian postglacial
transgression

Quebrada La Delfina is expressed by the presence of thick coal alluvial-plain deposits. The bay-margin facies are grouped
beds associated with other fine-grained flood-plain deposits, together in parasequences 1, 2, and 3. Bay-margin deposits (Facies
which are replaced upward by deltaic facies (Facies Associa- Association V) include a wide range of medium- to fine-grained
tion IV). sandstone facies (Sl, Swr, Sh, Sr [see Table 1 for definition of
In Mina La Ciénaga, the transgressive surface is of wave- abbreviations]) and heterolithic facies. Interbedded mudstone
ravinement nature, and bay-shoreline deposits overlie the braided and fine-grained sandstone facies (Facies Association VI), which
Mina Quebrada
La Ciénaga 6 km La Delfina

FA VII
FA VII

LST SB
Sequence boundary

P7 FS
TFA 2
TFA 3
HST P6
FS

FA VI
FA VI

P5 TFA 2
TFA 2 FS
Maximum flooding surface
P4
Sequence 2

FS
P3 FS

FA V
P2 FS
TST
FA V

P1

Sequence 2
FA IV
Transgressive surface FS
FA III

LST
Sequence boundary
FA II

Maximum flooding surface HST


Sequence 1

TST

FA III
Transgressive sur
face
FA I

TFA 1

LST
Sequence
boundary
Mudstone vf f m c vc cg
Sandstone

Legend
Current ripple
Brachiopods cross-lamination

Gastropods Trough cross-

Sequence 1
stratification
TFA 1
Bivalves Planar cross-

FA I
stratification
Palaeophycus crenulatus;
Parallel lamination
Planolites montanus
Treptichnus pollardi; Wave ripple
Helminthopsis abeli cross-lamination

Halopoa isp. Low-angle 10 m


cross-stratification
Rhizocorallium commune Coal
Mudstone vf f m c vc cg
Sandstone

Figure 4. Correlation between the two studied sections, and distribution of facies associations, sequence stratigraphic surfaces, parasequences,
systems tracts, and trace-fossil assemblages within Sequences 1 and 2. FA—facies association; TFA—trace-fossil assemblage; P—parasequence;
LST—lowstand systems tract; TST—transgressive systems tract; HST—highstand systems tract; SB—sequence boundary; FS—flooding sur-
face. The panel is oriented roughly parallel to the coastline.
TABLE 1. FACIES CHART (BASED ON DESJARDINS ET AL., 2009)
Lithofacies Lithology and sedimentary structures Sedimentary processes and depositional Sedimentary Distribution
conditions environments
Gt Trough cross-stratified, gravelly very coarse-grained sandstone. Each Tractive bed-load deposition from Braided and Facies association: III, VII
set generally comprises a sharp and erosive base. Clay chips are unidirectional currents as channel fills anastomosing
common. rivers

Gp Planar cross-stratified, gravelly very coarse-grained sandstone. Each Tractive bed-load deposition from Braided and Facies association: III, VII
set is limited by a sharp and planar base and top. Clay chips are unidirectional currents as longitudinal anastomosing
common. bars and/or channel fills rivers

Gh Matrix-supported conglomerate. Well-rounded clasts scattered in a Deposition from a concentrated density Braided river Facies association: III, VII
medium- to coarse-grained sandstone. flow

St Trough cross-stratified, moderate- to well-sorted, medium- to coarse- Tractive bed-load deposition from Braided river, coastal Facies association: I, III,
grained sandstone. Sharp and erosive bases; sharp or gradational unidirectional currents plain, and bay IV, V, VII
tops. margin

Sp Planar cross-stratified, moderate- to well-sorted, very fine to medium- Tractive bed-load deposition from Braided river, coastal Facies association: I, III,
grained sandstone. Erosive and sharp bases; sharp tops. unidirectional currents in channels plain IV, VII
Sl Horizontally laminated, well-sorted, very fine to medium-grained Bed-load + suspension-load deposition Braided river, coastal Facies association: I, II,
sandstone. Sharp to gradational bases and tops. Locally from unidirectional currents plain, coastal III, IV, VII
interbedded with thin layers of siltstone and claystone. lagoon, and lake

Sr Climbing ripple cross-laminated, very well-sorted, fine- to very fine- Continuous traction + suspension fallout Coastal plain, coastal Facies association: I, II,
grained sandstone. Gradational and sharp bases and tops. Locally deposition from turbulent currents lagoon, and delta III, IV, VII
interbedded within thin layers of siltstone and claystone.

Sh Planar to low-angle cross-stratified, well-sorted, medium- to fine- Sediment reworking by waves in the surf Bay shoreline Facies association: V
grained sandstone. Sharp bases and tops. Parting lineation is zone under an upper-flow regime
present.

Swr Wave ripple cross-laminated and low-angle cross-laminated, very well- Sediment deposit and/or reworked by Bay margin and distal Facies association: V, VI
sorted, very fine to fine-grained sandstone. Sharp bases and tops. oscillatory currents bay

Srd Very well-sorted, normally graded, fine- to very fine-grained Traction + suspension fallout from Distal bay Facies association: VI
sandstone. Sharp erosive base and dome-shaped tops. turbidity currents and high-energy
waves

Sm Massive, well-sorted, medium- to very fine-grained sandstone. Continuous traction + suspension fallout Coastal plain, delta, Facies association: I, II,
Gradational and sharp bases and tops. deposition from turbulent currents coastal lagoon IV

Fl Thin horizontally laminated mudstones, siltstone, and very fine-grained Suspension fallout, in times alternating Coastal plain, alluvial Facies association: III, IV,
sandstone. Sharp bases and tops. with traction + suspension fallout from plain, delta VII
turbulent currents

Fr Thin horizontally laminated to massive rooted mudstone and siltstone. Soil development on alluvial and coastal Coastal plain, alluvial Facies association: I, VII
Sharp bases and tops. plain deposits plain

Fc Thinly laminated chloritic mudstone and siltstone. Sharp bases and Suspension fallout Coastal lagoon Facies association: II
tops.
Fm Massive and thinly laminated marine mudstone. Sharp bases and Suspension fallout Distal bay Facies association: V, VI
tops.

C Coal and carbonaceous mudstone. Sharp bases and tops. Organic material accumulation in Coastal plain Facies association: I, III
setting with very low clastic sediment
input
TABLE 2. CHARACTERISTICS OF FACIES ASSEMBLAGES, LOCATION WITHIN THE SEQUENCE-STRATIGRAPHIC FRAMEWORK, AND ENVIRONMENTAL SIGNIFICANCE
Facies associations Lithofacies Remarks Systems tracts Sedimentary Locality
environment
I: St, Sp, Sr, Fining-upward cycles LST to TST Coastal plain Mina La Ciénaga
Interbedded medium- to fine-grained Sm, Sl, Fl, Fr, -Extensive, periodically inundated flood-plain Quebrada La Delfina
sandstone-siltstone-coal lithofacies C deposits
-Waterlogged soils
-Thickening-upward coal
-Ichnofauna: Helminthopsis abeli, Treptichnus
pollardi

II: Sl, Sm, Sr, Fc Chlorite mineralogy of mudstone suggests a marine TST to HST Coastal lagoon Mina La Ciénaga
Interbedded chlorite mudstone and connection associated to river-
sandstone lithofacies -Thickening-upward sandstone lithofacies (fluvial fed quasi-steady
progradation) turbidity currents

III: St, Sp, Sr, Sl, Laterally continuous, amalgamated lenticular package LST to TST Alluvial braided plain Mina La Ciénaga
Pebbly and coarse to medium-grained Gt, Gp, Gh, Fl, -Sharp-based, normally graded sandstone Quebrada la Delfina
amalgamated sandstone lithofacies C packages scouring into relatively thick coal
and coal lithofacies lithofacies
IV: Sl, Sr, Sp, Sm, Laminated mudstone containing small plant fragments TST River-dominated delta Quebrada La Delfina
Interbedded medium- to fine-grained St, Fl -Tabular fine-grained sandstone packages
sandstone-mudstone lithofacies -Amalgamated sandstone bodies containing
numerous reactivation surfaces

V: Sh, Swr, St, Retrogradational stacking pattern of shallow-marine TST Bay shoreline and bay Mina La Ciénaga
Medium- to very fine-grained Fm deposits margin Quebrada La Delfina
sandstone and heterolithic lithofacies -Dominance of wave-induced features
VI: Fm, Swr, Srd Thick mudstone layers TST to HST Distal bay Mina La Ciénaga
Interbedded mudstone and fine- -Tabular, normally graded, and wave-rippled Quebrada La Delfina
grained bioturbated sandstone sandstone packages
lithofacies -Scarce marine fauna of the Tivertonia jachalensis–
Streptorhynchus inaequiornatus Biozone
-Ichnofauna: Rhizocollarium commune, Halopoa
isp., Paleophycus crenulatus, Planolites montanus

VII: Gh, Gt, Gp, Erosionally based, laterally discontinuous sandstone LST? Anastomosing river Mina La Ciénaga
Interbedded pebbly and coarse- Sp, St, Sr, Sl, bodies Quebrada La Delfina
grained sandstone-siltstone lithofacies Sm, Fl, Fr -Abundance of large (up to 1 m wide) clay chips
-Thick siltstone beds
Note: Based on Desjardins et al. (2009). LST—lowstand systems tract; HST—highstand systems tract; TST—transgressive systems tract. Lithofacies abbreviations are defined in
Table 1.
Paganzo Basin of western Argentina 183

are interpreted as distal-bay deposits, make up parasequences 4, all trace fossils occurring within a particular facies or association
5, 6, and 7. The tabular sandstone facies are interpreted as the of facies. The integration of ichnologic and sedimentologic data
product of a high-density flow, which probably originated either provides the opportunity for refining interpretations of sedimen-
during a flooding stage of a river that flowed into the bay or from tary environments and depositional evolution within a sequence-
a storm-induced gravity flow. stratigraphic framework.
A maximum flooding surface is located at a mudstone hori-
zon containing a fragmentary and scarce marine fauna of the Trace-Fossil Assemblage 1
Tivertonia jachalensis–Streptorhynchus inaequiornatus Biozone
(Cisterna et al., 2005, 2006). In Quebrada La Delfina, brachio- Trace-Fossil Assemblage 1 consists of Helmintopsis abeli
pods, bivalves, gastropods, and a few crinoids occur within this and Treptichnus pollardi in flood-plain deposits. The assemblage
horizon. In Mina La Ciénaga, the only body fossils observed are is present at the base and top of very fine-grained sandstone (Sr)
present in a coquina composed of large gastropods. Within the within the fine-grained heterolithic facies (Facies Association I).
distal-bay tempestites, four ichnotaxa were identified and grouped Four slabs containing different specimens were collected and
into two different assemblages. Halopoa isp., Palaeophycus analyzed. Within this facies, paleosols with distinctive rhizoliths
crenulatus, and Planolites montanus are included in Trace-Fossil are present.
.
Assemblage 2, and Rhizocorallium commune compose Trace- Helminthopsis abeli Ksia˛zkiewicz, 1977 (Fig. 5A): Simple,
Fossil Assemblage 3. These trace fossils occur only in a sand- irregularly meandering and winding horizontal trails. The mean-
stone facies (Srd [see Table 1 for definition of abbreviation]). ders are irregular in shape and size. Diameter is 0.3–0.4 mm. The
trace fill is similar to the host rock. All specimens are preserved
Highstand Systems Tract as positive epirelief in mudstone facies. Helminthopsis is inter-
Poorly represented shallowing-upward distal-bay deposits preted as the product of the combined activity of locomotion and
occur above the maximum flooding surface. As a consequence of feeding (grazing traces) by arthropods and nematodes (Buatois
a base-level drop, a sequence boundary is located at the top of the and Mángano, 1993a).
distal-bay deposits, revealing the passage to continental environ- Treptichnus pollardi Buatois and Mángano, 1993b (Fig. 5B):
ments in a semiarid climate (Facies Association VII). Simple burrow systems consisting of straight, unlined, unorna-
mented, horizontal burrow segments. Small pits occur mostly
COMPOSITION AND CHARACTERISTICS OF within burrow segments and locally at the angle of juncture,
THE ICHNOFAUNA representing the bedding-plane expression of vertical shafts.
Diameter is 0.4–0.5 mm. Length of individual segments is 1.2–
Three different trace-fossil assemblages were identified in the 5.3 mm. Pit diameter is 1.1–1.6 mm. Specimens preserved as
stratigraphic sections. Each of these assemblages is composed of positive and negative epirelief in mudstone facies. Treptichnus is

Figure 5. Trace-Fossil Assemblage 1.


(A) Helminthopsis abeli. (B) Treptich-
nus pollardi.
184 Desjardins et al.

a feeding trace that in freshwater environments may be produced trace fossils. Diameter is 2.4–2.7 mm and constant within a
by worms or insect larvae (Buatois and Mángano, 1993b; Bua- specimen. The trace fill differs from the host rock in grain size
tois et al., 1998a). and color. The trace surfaces are typically smooth. Specimens
The environmental implications of this trace-fossil assem- are preserved as full relief. The ichnotaxonomy of Planolites has
blage are consistent with our interpretation for the associated been addressed by Pemberton and Frey (1982), who noted that
facies. The tracemakers inhabited a subaqueous freshwater sub- the diagnostic characteristic of P. montanus is the curved to con-
strate in a flood-plain environment characterized by good oxy- torted shape of the traces. These authors interpreted Planolites as
genation, and low-energy and sedimentation rates, periodically feeding structures of deposits feeders, most likely polychaetes.
disturbed by overbank flooding. No evidence of marine connec- Other authors (e.g., Fillion and Pickerill, 1990; Knaust, 2007),
tion has been detected for this facies assemblage. Other ichnofau- however, considered that other phyla could also be involved.
nas from identical environmental settings of the Tupe Formation In particular, Knaust attributed P. montanus to the activity of
show similarities with the assemblage observed in the studied bivalves and other mollusks.
localities. Buatois and Mángano (2002) reported the occurrence Halopoa isp. (Figs. 6D–6F): Horizontal to oblique burrows
of poorly specialized grazing and feeding trails (Helminthoid- comprising inflated-shape segments covered with longitudinal
ichnites tenuis, Planolites isp.) and other ichnotaxa, such as irregular striations. The segments are unbranched. Burrow seg-
Archaeonassa fossulata, Didymaulichnus lyelli, and Palaeophy- ments are 10.8–21.1 mm wide and 8.9–22.8 mm long. Striae
cus tubularis, in deposits of the Tupe Formation in the vicinity of are 0.8–0.9 mm. Burrows are preserved as full relief on both
Huerta de Huachi. As in the case of Trace-Fossil Assemblage 1, tops and bases of sandstone beds. Halopoa has been revised by
the ichnofauna documented by Buatois and Mángano (2002) is Uchman (1998, 2001), who recognized three different ichno-
.
characterized by the dominance of very simple forms, superficial species: H. imbricata Torell, 1870; H. annulata Ksia˛zkiewicz,
or very shallow trace fossils, mostly restricted to bedding-plane 1977; and H. storeana Uchman, 2001. Halopoa imbricata is
surfaces, allowing preservation of the primary sedimentary fabric. unbranched, and has long and continuous furrows and wrinkles.
Dominance of bedding-plane trace fossils and little disturbance of Jensen (1997), however, placed Halopoa imbricata in Palaeo-
primary fabric are typical of Paleozoic flood-plain deposits (Bua- phycus as P. imbricatus. Halopoa annulata is branched and
tois et al., 1998b). Flood-plain water-body trace-fossil assem- contains perpendicular constrictions. Halopoa storeana has
blages show an expected low ichnodiversity in comparison with wrinkles arranged in a distinct, oriented plait-like design. The
their equivalents from lacustrine basins. The depauperate assem- studied specimens display diagnostic features of Halopoa (e.g.,
blages in flood-plain deposits have been linked to the less stable predominately horizontal structures, inflated morphology, and
conditions and the temporary nature of their ponds (Buatois and longitudinal tension-related striations). However, determination
Mángano, 2002). at ichnospecific level is not possible. The fragmentary nature
of the analyzed specimens prevents evaluations of the overall
Trace-Fossil Assemblage 2 architecture. Two possible tracemakers may have produced
Halopoa: (1) worm-like organisms that could expand its body
Trace-Fossil Assemblage 2 consists of Palaeophycus crenu- hydraulically when moving through the sediment, and (2) crus-
latus, Planolites montanus, and Halopoa isp. This assemblage taceans that could push against the walls with their carapace
occurs in relatively thin and tabular distal-bay sandstone pack- (Seilacher, 2007).
ages (Sr) of Facies Association VI. The trace-fossil assemblage described is characterized by
Palaeophycus crenulatus Buckman, 1995 (Figs. 6A and 6B): low ichnodiversity but high abundance. The degree of bioturba-
Straight to sinuous, horizontal, unbranched, thinly lined, cylin- tion is locally intense, mostly due to the presence of high-density
drical, smooth-walled, endichnial burrows having distinctive assemblages of Halopoa isp. The assemblage consists dominantly
annulations. Overcrossing is very common (Fig. 6A). Burrow of horizontal trace fossils produced by deposit and suspension
diameter is 4.8–6.2 mm. Burrow fill matches the host rock and feeders. The tracemakers presumably inhabited a low-energy
is massive. Annulations have a spacing of 0.9–1.6 mm (Fig. 6B). distal-bay environment dominated by background sedimentation
Pemberton and Frey (1982) and Mángano et al. (2002) reviewed punctuated by storm episodes. However, the trace fossils are pre-
Palaeophycus. Palaeophycus crenulatus is characterized by a served in thin-bedded tempestites, most likely representing the
lined open burrow with millimeter-scale annulations. The annula- activity of post-storm colonizers. Although Palaeophycus and
tions indicate that the burrow lining was constructed as a series Planolites are facies-crossing forms that occur in both continen-
of ring-shaped structures. Palaeophycus is interpreted as the tal and marine settings, Halopoa seems to be restricted to marine
dwelling structure of predaceous or suspension-feeding animals conditions (e.g., Seilacher, 1955; Crimes and McCall, 1995;
.
(Pemberton and Frey, 1982). The polychaete Glycera has been Ksia˛zkiewicz, 1977; Jensen, 1997; Uchman, 1998; Buatois et al.,
proposed as a modern analogue for the Palaeophycus tracemaker 2001; Mángano et al., 2002). The more intense bioturbation in
by Osgood (1970). the Halopoa beds is also a significant difference with respect to
Planolites montanus Richter 1937 (Fig. 6C): Horizontal to Trace-Fossil Assemblage 1; freshwater ichnofaunas show no dis-
subhorizontal, subcylindrical sinuous and undulating, unlined turbance of the primary sedimentary fabric.
Figure 6. Trace-Fossil Assemblage 2. (A) Palaeophycus crenulatus. (B) Close-up of one specimen showing distinctive
annulations. (C) Planolites montanus. (D) Halopoa isp. (E) Halopoa isp. (F) Close-up of the striae of Halopoa isp.
186 Desjardins et al.

Trace-Fossil Assemblage 3 Recently, Knaust (2007) reexamined the type area and distin-
guished R. commune in addition to the other three ichnospecies.
This is a monospecific trace-fossil assemblage that consists Rhizocorallium jenense is distinguished from the other ichnospe-
of Rhizocorallium commune. It occurs at the top of one bed of cies by consisting of a straight, short, U-shaped spreiten-burrow
distal-bay tempestites (Facies Association VI) that can be traced commonly oblique to the bedding plane. However, the burrow
laterally for tens of meters. as a whole can be locally vertically retrusive. Rhizocorallium
Rhizocorallium commune Schmid, 1876 (Figs. 7A–D): irregulare is characterized by long sinuous, bifurcating or pla-
Horizontal to subhorizontal, straight, relatively short U-shaped nispiral U-shaped spreiten-burrows, whereas R. uliarense pres-
burrows displaying a protrusive spreite. The tube diameter is ents a diagnostic trochospiral U-shaped spreiten-burrow. Rhizo-
3.0–9.1 mm, burrow width is 31.0–107.2 mm, and length is corallium commune is mainly horizontally oriented and exhibits
90.1–240.2 mm (terminology following that of Schlirf, 2000). a clear spreite.
The spreite is continuous, having the same material as the host Crustaceans have been assigned as the most likely producers
rock and abundant fecal pellets. The arms of the causative burrow of Rhizocorallium by several authors (Weigelt, 1929; Seilacher,
may be parallel to each other or widen downward in some speci- 1955, 1967; Fürsich, 1974; Schlirf, 2000). This idea is supported
mens. The pellets are cylindrical in shape and are 0.9–1.5 mm by a relatively wide causative tube, the common preservation
long, and 0.4–0.8 mm wide. The ichnogenus Rhizocorallium has of scratch marks (Fürsich, 1974), and the presence of cylindri-
been reviewed by Fürsich (1974). According to this author, three cal or rod-shaped pellets, which are commonly associated with
different ichnospecies can be recognized: R. jenense Zenker, the activity of crustaceans (Ekdale et al., 1984; Bromley, 1996).
1836, R. irregulare Mayer, 1954, and R. uliarense Firtion, 1958. However, the wide range of morphologies and their occurrences

Figure 7. Trace-Fossil Assemblage 3. Rhizocorallium commune. (A–C) General view. Coin in C is 2.5 cm. (D) Close-up
showing rod-like pellets organized in a spreiten structure.
Paganzo Basin of western Argentina 187

in different paleoenvironments suggested either a combination of DISCUSSION


different deposit-feeding crustaceans or a mixture of crustaceans
and worm-like animals (Knaust, 2007). Although the crustacean Trace fossil have largely been used to infer salinity condi-
origin is the more widely accepted, it is still debated whether tions in paleoenvironmental reconstructions (e.g., Wightman
Rhizocorallium represents the activity of suspension- or deposit- et al., 1987; Pemberton and Wightman, 1992; MacEachern and
feeding organisms (Fürsich, 1974; Knaust, 2007; Seilacher, Pemberton, 1994; Buatois et al., 1997, 2005; Mángano and Bua-
2007). tois, 2004; MacEachern and Gingras, 2007). In the specific case
Specimens from the Tupe Formation are characterized by of late Paleozoic Gondwanic successions, a number of studies
relatively short but dominantly horizontal burrows, and the gen- have dealt with the relationships between ichnofauna and paleo-
eral architecture closely resembles that of R. commune. Density salinity in South American basins (e.g., Buatois and Mángano,
of specimens is moderate over a single bedding plane (Fig. 7D). 1992, 1993a, 1995a, 1995b, 2002, 2003; Nogueira and Netto,
A wide range of sizes is observed, suggesting that the surface 2001; Pazos, 2000, 2002b; Balistieri et al., 2002, 2003; Buatois
was available for colonization for a considerable amount of time. and del Papa, 2003; Mángano et al., 2003; Buatois et al., 2006,
Specimens do not crosscut each other. Some large specimens this volume).
(Figs. 7A and 7D) display downward expansion, suggesting Buatois et al. (2006, this volume) described and compared
growth of the producer. From the horizontal orientation, abun- the ichnology of postglacial transgressions in different basins
dance of pellets, and the tightly constructed spreite, a deposit- of Gondwana. They concluded that postglacial transgressions
feeding mode of life is inferred. commonly contain freshwater ichnofaunas of the Mermia and
Trace-Fossil Assemblage 3 represents the activity of a ben- Scoyenia ichnofacies that developed as a response to extreme
thic community that inhabited the muddy substrate of a low- freshwater discharges caused by the melting of ice masses (see
energy environment, burrowed down into the sediment, and also Buatois and Mángano, 2003). In the case of the early Late
expanded laterally at the top of the underlying sandstone layer. Carboniferous transgression in the Paganzo Basin recorded by
Specimens most likely penetrated into the fine-grained substrate the Guandacol Formation and coeval units, they noted the pres-
from the overlying flooding surface. MacEachern and Gingras ence of a freshwater assemblage, characterized by nonspecial-
(2007) indicated that distal-bay deposits accumulate in the ized grazing traces, arthropod trackways, and fish trails (Figs. 8
deepest parts of the bays or the most sheltered areas and tend and 9A). Paleoenvironmental reconstructions for these deposits
to be mud-dominated, reflecting overall low-energy conditions. suggest that deglaciation led to the establishment of large fresh-
Rhizocorallium may be regarded as a facies-crossing ichnotaxon, water bodies to the east and a marine incursion from the west
although it is more common in shallow- to marginal-marine (e.g., that flooded valleys, creating a series of fjords along the coast
Seilacher, 1955; Farrow, 1966; Hakes, 1976; El-Asa’ad, 1987; (Limarino et al., 2002; Marenssi et al., 2005; Buatois et al., 2006,
Fraaye and Werver, 1990; Mángano et al., 2002) than in deep- this volume). The presence of acritarchs in some beds, mostly in
marine settings (e.g., Uchman, 1991). The only well-documented the western region, supports a marine connection with intermit-
occurrence of Rhizocorallium in continental environments cor- tent periods of brackish-water conditions. However, acritarchs
responds to firm-grounds and specimens are very different from do not occur in beds containing freshwater ichnofaunas, and
those analyzed here (Fürsich and Mayr, 1981). moreover, trace fossils are associated with terrestrially derived

A B C
Figure 8. Freshwater trace-fossil assemblage present in the deposits of the early Late Carboniferous postglacial transgression.
(A) Mermia carickensis. (B) Maculichna carboniferous. (C) Orchesteropus atavus.
A. 5

7 6

4
3 2
1

1. Helminthopsis tenuis
2. Undichna britannica
3. Helminthoidichnites tenuis
4. Mermia carickensis
5. Treptichnus bifurcus
6. Circulichnis montanus
7. Maculichna carboniferus

B. Trace-Fossil Assemblage 1
Trace-Fossil Assemblage 2
2

1
Trace-Fossil Assemblage 3 1
3

2
1. Helminthopsis abeli
2. Treptichnus pollardi
1 1. Planolites montanus
2. Halopoa isp.
3. Palaeophycus crenulatus

1. Rhizocorallium commune

Figure 9. Postglacial transgressive scenarios. (A) Fjord environment influenced by extreme freshwater discharge from
retreating glaciers. A freshwater ichnofauna occurs in the transgressive deposits (Guandacol Formation). (B) Coastal
environment without direct influence of freshwater influx from melting ice masses. A freshwater ichnofauna is present in
coastal lakes and temporally inundated flood plains (Trace-Fossil Assemblage 1). A brackish-water ichnofauna (Trace-
Fossil Assemblages 2 and 3) occurs in distal-bay facies (Tupe Formation).
Paganzo Basin of western Argentina 189

palynomorphs (Buatois and Mángano, 2003; Buatois et al., 2006, ing stressed environmental conditions. The horizon containing a
this volume). Grazing trails are also present in the Trace-Fossil marine fauna is interpreted to contain the maximum flooding sur-
Assemblage 1, but associated with flood-plain pond deposits face, and suggests that normal or near-normal marine conditions
toward the base of the Tupe Formation, recording the activity of in this area of the basin were only reached during times of maxi-
a freshwater biota. mum transgression. Deposits of the latest Carboniferous–earliest
The described Trace-Fossil Assemblages 2 and 3, pres- Permian transgression in less-restricted areas (e.g., Quebrada La
ent in the latest Carboniferous–earliest Permian transgressive Herradura) contain a more abundant and diverse fauna than in the
deposits of the Tupe Formation, provide further evidence of localities analyzed in this article.
paleosalinity during sea-level rises. The presence of Halopoa Studies in late Paleozoic and Mesozoic marginal-marine
and Rhizocorallium clearly indicates marine conditions, and environments suggest that brackish-water trace-fossil assem-
represents a significant departure with respect to the early Late blages are characterized by (1) low ichnodiversity, (2) forms typi-
Carboniferous postglacial transgression in the area (Fig. 9B). cally found in marine environments, (3) mixture of vertical and
This is consistent with the presence of the marine fauna of the horizontal trace fossils from the Skolithos and Cruziana ichnofa-
Tivertonia jachalensis–Streptorhynchus inaequiornatus Bio- cies, (4) dominance of infaunal traces rather than epifaunal trails,
zone. The discrepancy between the ichnofaunas of these two (5) simple structures produced by trophic generalists, (6) vari-
coastal settings is mainly due to differences in the salinity levels able abundances, (7) presence of monospecific associations, and
of their transgressive seas. The distance of the depositional area (8) small sizes (Wightman et al., 1987; Pemberton and Wight-
with respect to the melting ice masses that triggered the trans- man, 1992; MacEachern and Pemberton, 1994; Mángano and
gressions created a salinity gradient. The Guandacol Formation Buatois, 2004; Buatois et al., 2005). Some of these features are
contains an ichnofauna dominated by simple trails and arthro- present in Trace-Fossil Assemblages 2 and 3. In particular, Trace-
pod trackways, which developed as a consequence of the high Fossil Assemblage 2 is characterized by a few ichnotaxa, and is
freshwater influx coming from the melting ice. The Tupe Forma- dominated by simple and relatively small infaunal trace fossils.
tion bay deposits recorded a transgression that occurred far away The bioturbation degree is variable, being locally intense where
from any glacial center. If the coastal areas transgressed are far Halopoa isp. is present. Trace-Fossil Assemblage 3 is a monospe-
from any glacial center, a marine signature is clearly expressed in cific assemblage. The overall simple architecture and restricted
the associated ichnofauna. The transgressive deposits recorded in distribution of these structures also suggest stressed conditions.
the Tupe Formation lack any evidence of glacial influence (e.g., The envisaged brackish-water scenario is consistent with the
dropstones), supporting previous schemes of paleoclimatic evo- presence of abundant microbial mat textures in coeval deposits
lution that suggest climatic amelioration in the Paganzo Basin identified farther west in the Uspallata region by Buatois et al.
during the Carboniferous–Permian transition (López Gamundí (2002). Formation of microbial mats is commonly inhibited in
et al., 1992). However, coeval transgressive deposits in the Phanerozoic open-marine environments due to disruption by
Paraná Basin display clear glacial signatures and suggest that the infaunal organisms. Buatois et al. (2002) noted that some marine
latest Carboniferous–earliest Permian transgression in Argentina intervals in the coeval Santa Elena Formation lack bioturbation
is a glacioeustatic response to deglaciation in the eastern basins and contain a wide variety of structures indicative of microbial
(Buatois et al., 2002). binding (e.g., Manchuriophycus-like cracks, load-casted ripples)
Several ichnologic features support the notion that the lat- that suggest anomalous conditions, probably reflecting dilution
est Carboniferous–earliest Permian transgression was not char- of seawater.
acterized by fully marine conditions in the study area. Overall, The origin of brackish-water conditions during the latest
trace fossils are exceedingly rare throughout the section stud- Carboniferous–earliest Permian is attributed to the local paleo-
ied. Although the scarcity of bedding planes may be detrimen- geography of the basin. The basin was fringed in all direc-
tal to trace-fossil identification, examination of vertical sections tions by topographic highs, including the Protoprecordillera to
reveals little or no bioturbation but preservation of primary sedi- the west, which severely limited water circulation between the
mentary structures in most of the beds. The trace fossils do not more marine settings of the west and the embayment on the east
resemble typical offshore ichnocoenoses, which are commonly (Fig. 3). A breach on the Protoprecodillera served as a connec-
characterized by the archetypal Cruziana ichnofacies (Pem- tion between the two basins, but at the same time this topographic
berton et al., 1992). Offshore environments under fully marine high allowed the Paganzo embayment to remain brackish. The
salinity conditions are commonly highly bioturbated and con- latest Carboniferous–earliest Permian transgressive deposits of
tain a high-diversity ichnofauna. In contrast, ichnodiversity in the Tupe Formation cannot be traced along depositional dip for
the marine deposits of the Tupe Formation is very low. Interest- long distances (i.e., more than 100 km). The inherited paleo-
ingly, the notable exception is the Halopoa-bearing deposits. geography from the glacial times favored the development of
In addition, providing that bedding-plane views are available, embayments, and the patchy distribution of these deposits along
trace-fossil suites consist of only one or two ichnotaxa. The trace- the entire basin supports the notion of embayments rather that an
fossil assemblages present in the marine intervals of the Tupe open sea. The scarcity of storm deposits, particularly in the bay-
Formation represent a depauperate Cruziana ichnofacies, reflect- margin successions, is consistent with deposition in a restricted
190 Desjardins et al.

bay. The presence of the Protoprecordillera would have protected stone layer. These two assemblages represent a depau-
the embayment from high-energy open-marine waves. Finally, perate Cruziana ichnofacies, suggesting brackish-water
another peculiarity of this embayment is the absence of tidal sig- conditions.
natures, which suggests a microtidal regime in the late Paleozoic 4. In a postglacial transgressive scenario, the proxim-
seaway. In addition to basin restriction, late Paleozoic Gondwanic ity to retreating ice masses is reflected in the nature of
seas only locally may have reached fully marine salinity condi- the ichnofauna. The postglacial transgression recorded
tions because of the overall high discharges of meltwater during in the underlying Guandacol Formation developed
postglacial episodes. in a fjord environment subject to extreme freshwater
Comparisons with Quaternary analogues are not straightfor- influxes, and the associated ichnofauna corresponds to
ward mainly because ichnologic studies in glacial to postglacial the freshwater Mermia and Scoyenia ichnofacies. The
settings are scarce. Virtasalo et al. (2006) documented a pattern latest Carboniferous–earliest Permian transgression of
of vertical distribution of trace fossils in Holocene cores from the Tupe Formation represents the eustatic response to
the Baltic Sea that resembles that of the Paganzo Basin, albeit at deglaciation of distal ice masses covering eastern regions
a smaller scale. An increase in ichnodiversity was noted toward of Gondwana, but did not evolve under the direct influ-
the end of the lacustrine phase in relation to a marine influence. ence of extreme meltwater discharges. The marine ichno-
Glacial-fed freshwater Ancylus Lake deposits are replaced by fauna of the Tupe Formation represents the activity of
those from the brackish-water Littorina Sea. The trace-fossil an infaunal community in a brackish-water, restricted
assemblage (Palaeophycus assemblage) of the Ancylus Lake embayment.
consists of Palaeophycus and Arenicolites, representing an
extremely depauperate ichnofauna. With the establishment of ACKNOWLEDGMENTS
brackish-water conditions, the Palaeophycus assemblage was
replaced by an assemblage consisting of Planolites, Arenicolites, Financial support for this study was provided by the Argen-
Lockeia, and Teichichnus (Planolites assemblage). However, ich- tinean Agency of Scientific Research, Natural Sciences and Engi-
nodiversity levels remain very low as a result of brackish-water neering Research Council (NSERC) Discovery Grants awarded
conditions and restricted oxygen availability. to Mángano and Buatois, as well as by University of Saskatch-
ewan startup funds awarded to Buatois. We are grateful to Dirk
CONCLUSIONS Knaust and Joonas Virtasalo for providing useful reviews. Alfred
Uchman is thanked for comments on ichnotaxonomic aspects,
1. Ichnofaunas from the latest Carboniferous–early Perm- Gabriela Cisterna for providing information on the associated
ian Tupe Formation reflect the shift from a nonmarine invertebrate fauna, and Ramiro Salvatierra for fieldwork assis-
freshwater community (Trace-Fossil Assemblage 1) to tance. Special thanks to Brian R. Pratt for his assistance in photo-
a marine community under brackish-water conditions graphing the specimens, and his comments on the manuscript.
(Trace-Fossil Assemblages 2 and 3).
2. Helminthopsis abeli and Treptichnus pollardi (Trace- REFERENCES CITED
Fossil Assemblage 1) occur in subaqueous freshwater
substrates of small water bodies on temporally inundated Balistieri, P.R.M.N., Netto, R.G., and Lavina, E.L.C., 2002, Ichnofauna from
the Upper Carboniferous–Lower Permian rhythmites from Mafra, Santa
flood plains of a transgressive coastal plain. This assem- Catarina State, Brazil: Ichnotaxonomy: Revista Brasileira de Paleontolo-
blage represents a depauperate Mermia ichnofacies. The gia, v. 4, p. 13–26.
lower diversity of the flood-plain–pond ichnofaunas in Balistieri, P.R.M.N., Netto, R.G., and Lavina, E.L.C., 2003, Icnofauna de ritmi-
tos do topo da Formação Mafra (Permo-Carbonífero da Bacia do Paraná)
comparison with their equivalents of the Mermia ichnofa- em Rio Negro, Estado do Paraná (PR), Brasil, in Buatois, L.A., and
cies of lacustrine and fjord environments is linked to the Mángano, M.G., eds., Icnología: Hacia una convergencia entre geología
less stable conditions and temporary nature of the flood- y biología: Publicación Especial de la Asociación Paleontológica Argen-
tina, v. 9, p. 131–139.
plain water bodies. Bohacs, K., and Suter, J., 1997, Sequence stratigraphic distribution of coaly
3. Halopoa isp., Palaeophycus crenulatus, and Planolites rocks: Fundamental controls and paralic examples: American Association
montanus (Trace-Fossil Assemblage 2) characterize a of Petroleum Geologists Bulletin, v. 81, p. 1639–1697.
Bromley, R.G., 1996, Trace Fossils: Biology, Taphonomy and Applications:
low-energy distal-bay environment dominated by back- London, Chapman and Hall, 361 p.
ground sedimentation. Their preservation in thin-bedded Buatois, L.A., and Mángano, M.G., 1992, Abanicos sublacustres, abanicos
tempestites reflects the activity of post-storm coloniz- submarinos o plataformas glacilacustres? Evidencias icnológicas para
una interpretación paleoambiental del Carbonífero de la cuenca Paganzo:
ers. Rhizocorallium commune (Trace-Fossil Assemblage Ameghiniana, v. 2, p. 323–335.
3) results from the activity of infaunal deposit feeders, Buatois, L.A., and Mángano, M.G., 1993a, Trace fossils from a Carboniferous
which inhabited the muddy distal-bay environment dur- turbiditic lake: Implications for the recognition of additional nonmarine
ing maximum transgression. The tracemakers burrowed ichnofacies: Ichnos, v. 2, p, 237–258.
Buatois, L.A., and Mángano, M.G., 1993b, The ichnotaxonomic status of
down from the maximum flooding surface, and expanded Plangtichnus and Treptichnus: Ichnos, v. 2, p. 217–224, doi: 10.1080/
laterally at the interface with the underlying tabular sand- 10420949309380095.
Paganzo Basin of western Argentina 191

Buatois, L.A., and Mángano, M.G., 1994, Lithofacies and depositional processes Cisterna, G.A., Sterren, A.F., and Archbold, N.W., 2006, A review of the Tiver-
from a Carboniferous lake, Sierra de Narváez, northwest Argentina: Sedi- tonia jachalensis–Streptorhynchus inaequiornatus Biozone in La Delfina
mentary Geology, v. 93, p. 25–49, doi: 10.1016/0037-0738(94)90027-2. Creek, San Juan province, Argentina: Ameghiniana, v. 43, p. 487–491.
Buatois, L.A., and Mángano, M.G., 1995a, Sedimentary dynamics and evolu- Crimes, T.P., and McCall, G.J.H., 1995, A diverse ichnofauna from Eocene–
tionary history of a Late Carboniferous Gondwanic lake at Northwestern Miocene rocks of the Makran Range (S.E. Iran): Ichnos, v. 3, p. 231–258,
Argentina: Sedimentology, v. 42, p. 415–436, doi: 10.1111/j.1365-3091 doi: 10.1080/10420949509386394.
.1995.tb00382.x. Cuerda, A., 1965, Estratigrafía de los depósitos neopaleozoicos de la Sierra de
Buatois, L.A., and Mángano, M.G., 1995b, Postglacial lacustrine event sedimen- Maz (Provincia de la Rioja), in Actas, 2° Jornadas de Geología Argentina,
tation in an ancient mountain setting: Carboniferous Lake Malanzán (west- Salta, Argentina, v. 3, p. 79–94.
ern Argentina): Journal of Paleolimnology, v. 14, p. 135–140, doi: 10.1007/ Desjardins, P.R., Buatois, L.A., Limarino, C.O., and Cisterna, G.A., 2009, Lat-
BF00682591. est Carboniferous–earliest Permian transgressive deposits in the Paganzo
Buatois, L.A., and Mángano, M.G., 2002, Trace fossils from Carboniferous Basin of Western Argentina: Lithofacies and sequence stratigraphy of a
floodplain deposits in western Argentina: Implications for ichnofacies coastal plain to shallow-marine succession: Journal of South American
models of continental environments: Palaeogeography, Palaeoclimatology, Earth Sciences, v. 28, p. 40–53, doi: 10.1016/j.jsames.2008.10.003.
Palaeoecology, v. 183, p. 71–86, doi: 10.1016/S0031-0182(01)00459-X. Ekdale, A.A., Bromley, R.G., and Pemberton, S.G., 1984, Ichnology—The use
Buatois, L.A., and Mángano, M.G., 2003, Caracterización icnológica y of trace fossils in sedimentology and stratigraphy: Society of Economic
paleoambiental de la localidad tipo de Orchesteropus atavus, Huerta Paleontologists and Mineralogists, Short Course 15, 317 p.
de Huachi, provincia de San Juan, Argentina: Implicancias en el debate El-Asa’ad, G.M.A., 1987, Mesozoic trace fossils from Central Saudi Arabia.
sobre los ambientes de sedimentación en el Carbonífero de Precordillera: Arabian Gulf Journal of Scientific Research: Mathematical and Physical
Ameghiniana, v. 40, p. 53–70. Sciences, v. A5, p. 205–224.
Buatois, L.A., and del Papa, C.E., 2003, Trazas fósiles de la Formación Tarija, Farrow, G.E., 1966, Bathymetric zonation of Jurassic trace fossils from the
Carbonífero Superior del norte argentino: Aspectos icnológicos de la coast of Yorkshire, England: Palaeogeography, Palaeoclimatology, Palae-
transgresión postglacial en el oeste de Gondwana, in Buatois, L.A., and oecology, v. 2, p. 103–151, doi: 10.1016/0031-0182(66)90011-3.
Mángano, M.G., eds., Icnología: Hacia una convergencia entre geología Fillion, D., and Pickerill, R.K., 1990, Ichnology of the Lower Ordovician Bell
y biología: Publicación Especial de la Asociación Paleontológica Argen- Island and Wabana Groups of eastern Newfoundland: Palaeontographica
tina, v. 9, p. 119–130. Canadiana, v. 7, p. 1–119.
Buatois, L.A., Mángano, M.G., Maples, C.G., and Lanier, W.P., 1997, The para- Firtion, F., 1958, Sur la presence d’ichnites dans le Portlandien de l’Ile d’Oléron
dox of nonmarine ichnofaunas in tidal rhythmites: Integrating sedimento- (Chanrente maritime): Annales Universitatis Saraviens (Naturwiss), v. 7,
logic and ichnologic data from the Late Carboniferous of eastern Kansas, p. 107–112.
USA: Palaios, v. 12, p. 467–481, doi: 10.2307/3515384. Fraaye, R.H.B., and Werver, O.P., 1990, Trace fossils and their environmen-
Buatois, L.A., Mángano, M.G., Maples, C.G., and Lanier, W.P., 1998a, Ich- tal significance in Dinantian carbonates of Belgium: Palaöntologische
nology of an Upper Carboniferous fluvio-estuarine paleovalley: The Zeitschrift, v. 64, p. 367–377.
Tonganoxie Sandstone, Buildex Quarry, eastern Kansas, USA: Journal of Fürsich, F.T., 1974, Ichnogenus Rhizocorallium: Palaöntologische Zeitschrift,
Paleontology, v. 72, p. 152–180. v. 48, p. 16–28.
Buatois, L.A., Mángano, M.G., Genise, J.F., and Taylor, T.N., 1998b, The ich- Fürsich, F.T., and Mayr, H., 1981, Non-marine Rhizocorallium (trace fossil)
nologic record of the invertebrate invasion of nonmarine ecosystems: from the Upper Freshwater Molasse (Upper Miocene) of southern Ger-
Evolutionary trends in ecospace utilization, environmental expansion, many: Neues Jahrbuch für Geologie und Paläontologie. Monatshefte,
and behavioral complexity: Palaios, v. 13, p. 217–240, doi: 10.2307/ v. 181, p. 321–333.
3515447. Hakes, W.G., 1976, Trace fossils and depositional environment of four clastic
Buatois, L.A., Mángano, M.G., and Sylvester, Z., 2001, A deep-marine ichno- units, Upper Pennsylvanian megacyclothems, northeast Kansas: Univer-
fauna from the Eocene Tarcau Sandstones of Eastern Carpathians, Roma- sity of Kansas Paleontological Contributions, v. 63, p. 1–46.
nia: Ichnos, v. 8, p. 23–62, doi: 10.1080/10420940109380172. Jensen, S., 1997, Trace fossils from the Lower Cambrian Mickwitzia sandstone,
Buatois, L.A., Netto, R.G., and Mángano, M.G., 2002, Estructuras vinculadas south-central Sweden: Fossils and Strata, v. 42, p. 1–111.
a tapetes microbiales en depósitos siliciclásticos Post-Vendianos: Eviden- Knaust, D., 2007, Invertebrate trace fossils and ichnodiversity in shallow-
cias de ecosistemas extremos en el Paleozoico Superior Gondwánico, in marine carbonates of the German Middle Triassic (Muschelkalk), in
Actas, 15° Congreso Geológico Argentino, El Calafate, Argentina, v. 3, Bromley, R., Buatois, L.A., Mángano, M.G., Genise, J., and Melchor, R.,
p. 140–141. eds., Sediment-Organism Interactions: A Multifaceted Ichnology: Society
Buatois, L.A., Gingras, M.K., MacEachern, J., Mángano, M.G., Zonneveld, of Economic Paleontologists and Mineralogists Special Publication 88,
J.-P., Pemberton, S.G., Netto, R.G., and Martin, A.J., 2005, Colonization .p. 223–240.
of brackish-water systems through time: Evidence from the trace-fossil Ksia˛z kiewicz, M., 1977, Trace fossils in the flysch of the Polish Carpathians:
record: Palaios, v. 20, p. 321–347, doi: 10.2110/palo.2004.p04-32. Paleontologia Polonica, v. 36, p. 1–200.
Buatois, L.A., Netto, R.G., Mángano, M.G., and Balistieri, P.R.M.N., 2006, Limarino, C.O., and Césari, S., 1988, Paleoclimatic significance of the lacus-
Extreme freshwater release during the late Paleozoic Gondwana deglacia- trine Carboniferous deposits in northwest Argentina: Palaeogeography,
tion and its impact on coastal settings: Geology, v. 34, p. 1021–1024, doi: 10 Palaeoclimatology, Palaeoecology, v. 65, p. 115–131, doi: 10.1016/0031
.1130/G22994A.1. -0182(88)90116-2.
Buatois, L.A., Netto, R.G., and Mángano, M.G., 2010, this volume, Ichnol- Limarino, C.O., Sessarego, H., Césari, S., and Lopéz Gamundí, O., 1986,
ogy of late Paleozoic postglacial transgressive deposits in Gondwana: El Perfil de la cuesta de Huaco, estratotipo de referencia (Hipoestratotipo)
Reconstructing salinity conditions in coastal ecosystems affected by del Grupo Paganzo en la Precordillera Central: Anales Academia Nacio-
strong meltwater discharge, in López-Gamundí, O.R., and Buatois, nal de Ciencias Exactas, Físicas y Naturales, v. 38, p. 81–108.
L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions Limarino, C.O., Césari, S.N., Net, L.I., Marenssi, A., Gutiérrez, R.P., and Tri-
in Gondwana: Geological Society of America Special Paper 468, doi: paldi, A., 2002, The Upper Carboniferous postglacial transgression in
10.1130/2010.2468(07). the Paganzo and Río Blanco basins (northwestern Argentina): Facies and
Buckman, J.O., 1995, A comment on annulate forms of Palaeophycus Hall stratigraphic significance: Journal of South American Earth Sciences,
1847: With particular reference to P. “annulatus” sensu Pemberton and v. 15, p. 445–460, doi: 10.1016/S0895-9811(02)00048-2.
Frey 1982, and the erection of P. crenulatus ichnosp. nov: Ichnos, v. 4, López Gamundí, O.R., 1989, Postglacial transgressions in Late Paleozoic
p. 131–140, doi: 10.1080/10420949509380120. basins of western Argentina: A record of glacioeustatic sea level rise:
Cisterna, G.A., Gutiérrez, P.R., Sterren, A.F., Desjardins, P.R., and Balarino, L., Palaeogeography, Palaeoclimatology, Palaeoecology, v. 71, p. 257–270,
2005, The marine interval of the Tupe Formation in western Paganzo doi: 10.1016/0031-0182(89)90054-0.
Basin and its implication in the definition of the Carboniferous–Permian López Gamundí, O.R., 1997, Glacial-postglacial transition in the Late Paleozoic
boundary in South America, in Gondwana 12 Conference, Geological basin of southern South America, in Martini, I.P., ed., Late Glacial and
and Biological Heritage of Gondwana, Mendoza, Argentina: Academia Postglacial Environmental Changes: Quaternary, Carboniferous–Permian,
Nacional de Ciencias, Argentina, Abstracts, p. 105. and Proterozoic: New York, Oxford University Press, p. 147–168.
192 Desjardins et al.

López Gamundí, O.R., Limarino, C.O., and Césari, S.N., 1992, Late Paleo- South America) and the Great Karoo–Kalahari Basin (Southern Africa):
zoic paleoclimatology of central west Argentina: Palaeogeography, Ichnological implications: Gondwana Research, v. 5, p. 619–640, doi:
Palaeoclimatology, Palaeoecology, v. 91, p. 305–329, doi: 10.1016/0031 10.1016/S1342-937X(05)70634-1.
-0182(92)90074-F. Pemberton, S.G., and Frey, R.W., 1982, Trace fossil nomenclature and the Plan-
MacEachern, J.A., and Gingras, M., 2007, Recognition of brackish-water trace olites-Palaeophycus dilemma: Journal of Paleontology, v. 56, p. 843–881.
fossil assemblages in the Cretaceous western interior seaway of Alberta, Pemberton, S.G., and Wightman, D.M., 1992, Ichnological characteristics of
in Bromley, R., Buatois, L.A., Mángano, M.G., Genise, J., and Melchor, brackish water deposits, in Pemberton, S.G., ed., Applications of Ichnol-
R., eds., Sediment-Organism Interactions: A Multifaceted Ichnology: ogy to Petroleum Exploration: Society of Economic Paleontologists and
Society of Economic Paleontologists and Mineralogists Special Publica- Mineralogists, Core Workshop 17, p. 141–167.
tion 88, p. 149–194. Pemberton, S.G., MacEachern, J.A., and Frey, R.W., 1992, Trace fossils facies
MacEachern, J.A., and Pemberton, S.G., 1994, Ichnological aspects of incised models: Environmental and allostratigraphic significance, in Walker, R.G.,
valley fill systems from the Viking Formation of the Western Canada and James, N.P., eds., Facies Models: Response to Sea Level Change:
Sedimentary Basin, Alberta, Canada, in Boyd, R., Zaitlin, B.A., and Waterloo, Geological Association of Canada, p. 47–72.
Dalrymple, R., eds., Incised Valley Systems—Origin and Sedimentary Remane, A., and Schlieper, C., 1971, Biology of Brackish Water: New York,
Sequences: Society of Economic Paleontologists and Mineralogists Spe- John Wiley and Sons, 372 p.
cial Publication 51, p. 129–157. Richter, R., 1937, Marken und spuren aus allen Zeiten, Parts 1 and 2: Sencken-
Mángano, M.G., and Buatois, L.A., 2004, Ichnology of Carboniferous tide- bergiana, v. 19, p. 150–169.
influenced environments and tidal flat variability in the North American Schlirf, M., 2000, Upper Jurassic trace fossils from the Boulonnais (northern
Midcontinent, in McIlroy, D., ed., The Application of Ichnology to Palae- France): Beringeria, v. 34, p. 145–213.
oenvironmental and Stratigraphic Analysis: Geological Society [London] Schmid, E.E., 1876, Der Muschelkalk des ostlichen Thuringen: Jena, Fromann,
Special Publication 228, p. 157–178. 20 p.
Mángano, M.G., Buatois, L.A., West, R.R., and Maples, C.G., 2002, Ichnol- Seilacher, A., 1955, Spuren und Fazies im Unterkambrium; in Beiträge zur
ogy of a Pennsylvanian equatorial tidal flat—The Stull Shale Member Kenntnis des Kambriums in der Salt Range (Pakistan), in Schindewolf,
at Waverly, eastern Kansas: Kansas Geological Survey Bulletin, v. 245, O.H., and Seilacher, A., eds., Akademie der Wissenschaften und der Lit-
133 p. eratur zu Mainz, mathematisch-naturwissenschaftliche Klasse, Abhand-
Mángano, M.G., Buatois, L.A., Limarino, C.O., Tripaldi, A., and Caselli, A., lungen, v. 10, p. 11–143.
2003, El icnogénero Psammichnites Torell, 1870 en la Formación Hoyada Seilacher, A., 1967, Bathymetry of trace fossils: Marine Geology, v. 5, p. 413–
Verde, Carbonífero Superior de la cuenca Calingasta-Uspallata: Ameghin- 428, doi: 10.1016/0025-3227(67)90051-5.
iana, v. 40, p. 601–608. Seilacher, A., 2007, Trace Fossil Analysis: Berlin, Springer, 226 p.
Marenssi, S.A., Tripaldi, A., Limarino, C.O., and Caselli, A.T., 2005, Facies Torell, O., 1870, Petrificata Suecana Formationis Cambricae: Lunds Univer-
and architecture of a Carboniferous grounding-line system from the sitets Årsskrift 6: Afd, v. 28, p. 1–14.
Guandacol Formation, Paganzo Basin, northwestern Argentina: Gond- Uchman, A., 1991, Trace fossils from stress environments in Cretaceous-
wana Research, v. 8, p. 187–202, doi: 10.1016/S1342-937X(05)71117-5. Palaeogene of Polish Outer Carpathians: Annales Societatis Geologorum
Mayer, G., 1954, Neue Beobachtungen an Lebensspuren aus dem unteren Poloniae, v. 61, p. 207–220.
Hauptmuschelkalk (Trochitenkalk) von Wiesloch: Neues Jahrbuch für Uchman, A., 1998, Taxonomy and ethology of flysch trace fossils: Revision of
Geologie und Paläontologie. Monatshefte, v. 99, p. 223–229. the Marian Ksiazkiewicz collection and studies of complementary mate-
McLusky, D.S., 1989, The Estuarine Ecosystem: New York, Blackie, 150 p. rial: Annales Societatis Geologorum Poloniae, v. 64, p. 105–208.
Net, L.I., Alonso, M.S., and Limarino, C.O., 2002, Source rock and environ- Uchman, A., 2001, Eocene flysch trace fossils from the Hecho Group of the
mental control on clay mineral associations, Lower Section of Paganzo Pyrenees, northern Spain: Beringeria, v. 28, p. 3–41.
Group (Carboniferous): Northwest Argentina: Sedimentary Geology, Virtasalo, J.J., Kotilaine, A.T., and Gingras, M.K., 2006, Trace fossils as indica-
v. 152, p. 183–199, doi: 10.1016/S0037-0738(02)00068-4. tors of environmental change in Holocene sediments of the Archipelago
Nogueira, M.S., and Netto, R.G., 2001, Icnofauna da Formação Rio do Sul Sea, northern Baltic Sea: Palaeogeography, Palaeoclimatology, Palaeo-
(Grupo Itararé, Permiano da Bacia do Paraná) na Pedreira Itaú-Itauna, ecology, v. 240, p. 453–467, doi: 10.1016/j.palaeo.2006.02.010.
Santa Catarina, Brasil: Acta Geologica Leopoldensia, v. 24, no. 52/53, Weigelt, J., 1929, Fossile Grabschächete brachyurer Decapoden als Lokalge-
p. 397–406. schiebe in Pommer und das Rhizocorallium-Problem: Z. f: Geschiebe-
Osgood, R.G., 1970, Trace fossils of the Cincinnati area: Palaeontographica forschung, v. 5, p. 1–42.
Americana, v. 6, p. 277–444. Wightman, D.M., Pemberton, S.G., and Singh, C., 1987, Depositional modelling
Pazos, P., 1994, Sedimentología de la transgresión estefaniana en el área of the Upper Mannville (Lower Cretaceous), east-central Alberta: Impli-
de Huaco, dpto. Jachal. Provincia de San Juan, in Actas, 5° Reunión cations for the recognition of brackish water deposits, in Tillman, R.W.,
Argentina de Sedimentología, San Miguel de Tucumán, Argentina, v. 1, and Weber, K.J., eds., Reservoir Sedimentology: Society of Economic
p. 77–82. Paleontologists and Mineralogists Special Publication 40, p. 189–220.
Pazos, P., 2000, Trace fossils and facies in glacial to postglacial deposits from Zenker, J.C., 1836, Historisch-topographisiches Taschenbuch von Jena and
the Paganzo basin (Late Carboniferous), central Precordillera, Argentina: seiner Umgebung besonders in seiner naturwissenschaftlicher und medi-
Ameghiniana, v. 37, p. 23–38. cinischer Beziehung: Jena, J. C. Zenker, 338 p.
Pazos, P.J., 2002a, The Late Carboniferous glacial to postglacial transition: Facies
and sequence stratigraphy, Western Paganzo Basin, Argentina: Gondwana
Research, v. 5, p. 467–487, doi: 10.1016/S1342-937X(05)70736-X.
Pazos, P.J., 2002b, Palaeoenvironmental framework of the glacial-postglacial
transition (Late Paleozoic) in the Paganzo-Calingasta Basin (Southern MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA


The Geological Society of America
Special Paper 468
2010

Reconstruction of a high-latitude, postglacial lake: Mackellar


Formation (Permian), Transantarctic Mountains

Molly F. Miller
Department of Earth and Environmental Sciences, Vanderbilt University, Nashville, Tennessee 37235, USA

John L. Isbell
Department of Geosciences, University of Wisconsin, Milwaukee, Wisconsin 53201, USA

ABSTRACT

The Lower Permian Mackellar Formation is well exposed in a 10,000 km2 out-
crop belt in the Nimrod, Beardmore, and Shackleton Glacier areas of the Transant-
arctic Mountains. This formation directly overlies glacial deposits and provides a
unique glimpse of high paleolatitude conditions during the last icehouse to green-
house transition. The unit records deposition in the Mackellar Lake or Inland Sea
(MLIS), a fresh-water body at ~80° S paleolatitude that was broadly analogous to
Glacial Lake Agassiz and was filled by fine-grained turbidites. Low total organic car-
bon (TOC) content and predominant vitrinite and inertinite are consistent with a
low influx of organic matter from a sparsely vegetated, recently deglaciated terrain.
A widespread but low-diversity ichnofauna and variable (although low overall) lev-
els of bioturbation suggest oxic conditions and a bottom fauna restricted to areas of
low sedimentation. Integration of sedimentologic, organic geochemical, and paleobio-
logic information with results of climate models and characteristics of modern lakes
enhances reconstruction of parameters that controlled the functioning of the lake
as an ecosystem. Regression equations relating mean annual temperature and mean
depth of modern lakes to the number of ice-free days applied to the MLIS indicate
ice cover from two to five months a year. Estimates of the depth of mixing and depth
to the thermocline, based on maximum length, maximum width, and area, suggest a
mixing depth of ~50 m and a thermocline of ~20 m. The MLIS probably was stratified
during the summer and was dimictic, with overturns occurring after fall cooling and
after ice melt; mixing was enhanced by turbidity currents. Productivity was low, as
recorded by the low TOC, but organic matter fixed in the surface water of the lake
may have been degraded and not recorded in the sediments. In spite of its high paleo-
latitude, the MLIS as reconstructed was dynamic and biologically active; the same
probably was true of other Permian postglacial lakes.

Miller, M.F., and Isbell, J.L., 2010, Reconstruction of a high-latitude, postglacial lake: Mackellar Formation (Permian), Transantarctic Mountains, in López-
Gamundí, O.R., and Buatois, L.A., eds., Late Paleozoic Glacial Events and Postglacial Transgressions in Gondwana: Geological Society of America Special Paper
468, p. 193–207, doi: 10.1130/2010.2468(09). For permission to copy, contact [email protected]. ©2010 The Geological Society of America. All rights
reserved.

193
194 Miller and Isbell

INTRODUCTION compare them to features of modern lakes in order to reconstruct


the ecosystem of the Makellar Lake or Inland Sea (MLIS). The
Upper Paleozoic sedimentary rocks record the history of focus herein is the outcrop between the Nimrod Glacier and the
the last large-scale period of glaciation and subsequent warm- Shackleton Glacier (Fig. 2). During the Early Permian, this area
ing, thus providing the most recent record of the transition from lay at 75° S to 85° S, similar to its present location (Powell and
icehouse to greenhouse conditions (e.g., Rees et al., 2002; Pow- Li, 1994).
ell, 2005). On a worldwide basis, this record is best exposed in Permian to Jurassic sedimentary rocks in the central Trans-
postglacial lake deposits that occur on diverse Gondwana con- antarctic Mountains form a sequence several kilometers thick of
tinents, including South America (Buatois and Mangano, 1995; continental glacial, lacustrine, fluvial, and coal-bearing deposits
López-Gamundí, 1997; Limarino et al., 2002;Trosdtorf et al., (Table 1). The Mackellar Formation consists of interbedded shale
2005; Archanjo et al., 2006), Africa (Visser, 1994; Scheffler and sandstone and overlies the earliest Permian Pagoda Forma-
et al., 2003), Australia (Dickins, 1996; Lindsay, 1997; Eyles tion, which is composed of glacial deposits, primarly glaciolacus-
et al., 2003; Jones and Fielding, 2004), India (Maejima et al., trine or glaciomarine, and secondarily subglacial deposits (Elliot,
2004), and Antarctica (Collinson et al., 1994; Miller and Collin- 1975; Barrett et al., 1986; Miller, 1989; Isbell et al., 2003). It is
son, 1994; Isbell et al., 2003). overlain by braided stream deposits of the Fairchild Formation,
The well exposed Mackellar Formation of the central Trans- and coal-bearing fluvial deposits of the Buckley Formation. In
antarctic Mountains allows reconstruction of the postglacial situ stumps within the Buckley Formation reflect dense forests
events and history that characterized southern polar paleolatitudes of Glossopteris trees that averaged >15 m tall (Knepprath et al.,
immediately following the final melting of ice sheets in Antarc- 2004), attesting to the climate amelioration that occurred after
tica during the earliest Permian. The Mackellar Formation was glaciation in spite of the continued high paleolatitude (Powell
deposited in a large lake or series of smaller time-transgressive and Li, 1994).
lakes that extended >1000 km through the present-day Transant-
arctic Mountains from north of the Nimrod Glacier to the Ells- MACKELLAR FORMATION: DEPOSITIONAL
worth Mountains (Fig. 1); the Mackellar body of water thus is SETTING AND PROCESSES
broadly comparable to large Quaternary lakes of North America
such as glacial Lake Agassiz. In this paper we integrate sedimen- The Mackellar Formation records the filling of a postglacial
tological, geochemical, and biological data to delineate the char- basin occupied by the Mackellar Lake or Inland Sea (MLIS). Its
acteristics of the lake recorded by the Mackellar Formation and origin, setting, and size are comparable to those of glacial Lake
Agassiz, although the near absence of shoreline features, paucity
of internally correlative units, and poor age control preclude the
detailed reconstruction that has elucidated the history of Lake
Agassiz (e.g., Teller and Clayton, 1983; Teller and Bluemle,
1983; Teller, 2001; Teller et al., 2005). Excellent exposures of
the Mackellar Formation in many areas of the Transantarctic
Mountains allow interpretation of the depositional setting and
processes, which are inferred to resemble those of Lake Agas-
siz. Major loci of deposition in Lake Agassiz were large fans of
sediment at the mouths of spillways (Fenton et al., 1983). These
fans record bed-load deposition from underflow currents (Fenton
et al., 1983).
The Mackellar Formation consists primarily of facies inter-
preted as underflow deposits that are analogous to underflow
deposits of Lake Agassiz and that accumulated in diverse com-
ponents of what has been described as a lacustrine turbidite
sequence (Table 2; Fig. 3; Miller and Collinson, 1994). In the
Beardmore Glacier area the unit is dominated by the interbedded
sandstone and shale facies, with the associated shale facies and
massive sandstone facies providing important information about
the sedimentary processes. The shale facies records quiet-water
conditions. In contrast, the interbedded sandstone and shale
facies reflects quiet-water conditions interrupted by turbidity-
current events that deposited upward-fining sandstones (Figs. 4
Figure 1. Extent of outcrop of the Mackellar Formation and its correla- and 5; Tb-d of Lowe, 1982). In the Beardmore Glacier area, indi-
tives from the Nimrod Glacier area to the Ellsworth Mountains. vidual upward-fining beds are organized into upward-coarsening
Mackellar Formation, Transantarctic Mountains 195

170°W 180° 170°E 160°E 150°E

85°S 84°S

84°S
83°S

83°S

82°S

0 50 100 km
82°S

180° 170°E 160°E

Figure 2. Outcrop locations within the study area. The shoreline of the Mackellar Lake or Inland Sea (MLIS) is inferred
to have extended from north of the Nimrod Glacier away from the Ross Ice Shelf and south toward Mount Bowers.

TABLE 1. STRATIGRAPHIC UNITS AND DEPOSITIONAL ENVIRONMENTS, BEARDMORE GLACIER AREA


Formation Rock type Thickness Inferred environment
(fossil) (m)
Buckley Formation Sandstone, shale, coal (Glossopteris) 750 Braided stream systems, lowland lakes
Fairchild Formation Sandstone 250 Braided stream
Mackellar Formation Shale, sandstone 100 Lake, inland sea
Pagoda Formation Diamictite, sandstone 200 Glacial, glacial basinal
Note: From Barrett et al. (1986), Isbell and Collinson (1988), Miller (1989), and Isbell et al. (2003).

sequences a few to tens of meters thick (Fig. 4). These packages remobilization of poorly sorted glacial debris from the underly-
are interpreted as crevasse splays of turbidite channels. The larg- ing Pagoda Formation. The burrowed sandstone facies records
est scale upward-coarsening sequences are capped by the mas- shallow shoreline deposition near the lake margin but away from
sive sandstone facies. This facies is composed of beds up to 3 m centers of deposition.
thick of homogeneous, structureless fine-grained sandstone that The shoreline of the MLIS is inferred to have swung east
fills cross-cutting channels and grades upward into sandstone from north of the Nimrod Glacier and south along the present-day
with planar lamination and climbing ripple lamination (Fig. 6; boundary between the Transantarctic Mountains and the ice land-
Miller and Collinson, 1994). We infer that this sandstone facies scape of the Polar Plateau (Fig. 2). In this configuration the Moore
was deposited in subaqueous channels that served as the major Mountains were closest to the locus of deposition, and the Shack-
conduits of the turbidite system. At two localities the large-scale, leton Glacier area (Mount Butters, Ramsey Glacier) was least
cross-stratified sandstone facies records Gilbert-type deltas feed- affected by sediment influx. The rate of deposition was very high
ing the turbidite system, and the rare diamictite facies reflects in the Moore Mountains (Fig. 2), as evidenced by omnipresent
TABLE 2. MACKELLAR FORMATION FACIES AND FACIES INTERPRETATION
Facies/abundance Grain size Structures/geometry Scale Trends/association Interpretation
Shale Clay, silt Horizontal lamination, regular to variable Few meters Most common at base of upward- Deposition from suspension
Widespread, but minor thickness, normal grading coarsening sequences and low-density turbidity
currents

Interbedded Sand: fine to very Normal grading, horizontal lamination, Tens of Gradationally overlies shale. Includes Deposition from high- and
sandstone and shale fine climbing ripple lamination, parting meters hierarchy of upward-coarsening low-density turbidity
Abundant; dominant Mud: clay to lineation, linguoid and 2-D ripples, local sequences, largest capped by massive currents
facies coarse silt bioturbation sandstone facies

Massive sandstone Fine sand, with Massive, but grades up into planar and Few meters Caps upward-coarsening sequences Deposition from high and
Present at all localities minor coarser climbing ripple lamination. Fills broad low (minor) turbidity
component channels, <5 m deep, hundreds of m currents
wide

Burrowed sandstone Fine sand Bioturbation; symmetrical ripple marks 10 m Near top, surrounded by interbedded Deposited above wave
Rare sandstone and shale facies base, away from loci of
deposition

Large-scale cross- Fine sand; minor Normally graded foresets with climbing 10 m Near top, grades into Fairchild Fm. Possible Gilbert-type delta
stratified sandstone clay, silt ripple lamination
Rare

Diamictite Silt-sand matrix Fills broad, shallow channel at one ~3 m Surrounded by interbedded sandstone Deposited by debris flow
Rare with clasts up to locality and shale facies locally derived from
3m remnant highs of glacial
diamictite
Note: From Miller and Collinson (1994). 2-D—two-dimensional.
Figure 3. Depositional environment of
the Mackellar Formation within the con-
text of the underlying and overlying units
(Table 1). Letters indicate inferred loca-
tions within the MLIS (Mackeller Lake
or Inland Sea) of facies well exposed at
outcrops shown in Figure 2. A—Moore
Mountains; B—Mount Weeks; C—
Tillite Glacier; D—Mount Butters and
Ramsey Glacier; E—Mount Bowers.

A
Figure 4. (A) Upward-fining sandstone in interbedded sandstone and
shale facies, Tillite Glacier; 15 cm ruler for scale. (B) Upward-coarsening
sequence in interbedded sandstone and shale facies, Tillite Glacier,
recording progradation of crevasse splay of turbidite channel.

B
A
Figure 5. (A) Interbedded sandstone and shale facies at Mount But-
ters, lacking well-defined upward coarsening; interpreted as depos-
ited away from crevasse splay. (B) Fine sandstone with planar beds
and climbing ripple lamination, indicating a high flow regime and B
rapid sedimentation near sediment source; Moore Mountains.

B
A
Figure 6. (A) Massive sandstone–filled channel (sandstone facies) that grades upward into climbing ripple lamination, Moore
Mountains. Person in center for scale. (B) Permian marine turbidite, intensely bioturbated in both vertical and horizontal
dimensions. Bluff, New Zealand. The penetrative and pervasive bioturbation contrasts with that of the Mackellar Formation
(Figs. 7 and 8). Ruler is 1 cm long.
Mackellar Formation, Transantarctic Mountains 199

climbing ripple lamination and the absence of bioturbation deira, 2002; Miller et al., 2005) using pattern recognition meth-
(Fig. 5B). In contrast, the Mackellar Formation at Mount Butters ods (ichnofabric indicies [ii] of Droser and Bottjer, 1986; Bed-
and the Ramsey Glacier in the Shackleton Glacier region consists ding Plane Bioturbation Indices [BPBI] of Miller and Smail,
of shale and interbedded sandstones that are not organized into 1997). Overall, there is relatively little bioturbation on horizontal
upward-coarsening sequences (Fig. 5A). We interpret this region surfaces of the interbedded sandstone and shale facies (Fig. 8),
as the most distal and relatively unaffected by the crevasse splays with 89% of observations showing no bioturbation (BPBI = 1)
of the turbidite system. Mount Bowers, near the Polar Plateau, and only 6.3% of observations showing BPBI = 5 (60%–100%
is inferred to have been relatively close to the paleoshoreline; a bioturbated). There is virtually no vertical bioturbation.
basin-margin setting is consistent with paleocurrent data, suggest- Although the sparse bioturbation is widely distributed,
ing transverse flow into the basin (Miller and Collinson, 1994, important differences in the amount of bioturbation on bedding
Fig. 2) and with inferred basin morphology (Isbell, 1999). plane surfaces in different areas of the lake system (Fig. 8). In the
Moore Mountains (Fig. 2) there is no bioturbation recorded on
BIOGENIC STRUCTURES AND DISTRIBUTION bedding plane surfaces. This area is closest to the lake margin,
OF BIOTURBATION which was north of the present-day Nimrod Glacier, and there
is abundant evidence of rapid sedimentation that probably pre-
Trace fossils are relatively rare in the Mackellar Formation cluded colonization by infaunal animals (Fig. 5B). Tillite Gla-
and consist primarily of simple horizontal endostratal trails assign- cier, with well-developed upward coarsening sequences, records
able to the ichnogenera Paleophycus or Planolites (Fig. 7). Rarely deposition close to turbidite channels in areas affected by crevasse
the traces formed looped patterns on bedding planes, resembling splays (Figs. 2 and 4). Crevasse splay deposition and prograda-
the trace fossil Mermia (Buatois and Mangano, 1995); the dis- tion are recorded by upward coarsening sequences and would be
tinctive traces Cochlichnus and Treptichnus also occur in rare inimical to infaunal animals, resulting in the observed low lev-
horizons (Fig. 7). Scratch marks are discernible on the margins els of bioturbation. Extensive bioturbation would be most likely
of some traces; some of these are bilobed and were produced by to have occurred in the distal parts of the lake, far from sources
animals with skeletalized appendages (e.g., arthropods). Scratch of sediment (e.g., in the Shackleton area), and also in nearshore
marks occur on parts of otherwise structureless traces, implying areas that were protected from rapid sedimentation. In modern
preservational control over their distribution. The widespread, if lakes, shallow littoral zones have the highest diversity of benthic
uneven, distribution of diverse simple traces suggests that many, animals, particularly in areas shielded from wave activity. The
if not all, of the trace fossils were produced by arthropods. Some presence of symmetrical ripples at Mount Weeks indicates wave
of the small traces may have been produced by conchostracans, activity in a nearshore zone, but the relatively extensive bioturba-
which have been reported from proglacial lake deposits of the tion implies that this shoreline zone was sufficiently protected
Pagoda Formation at Mount Butters (Fig. 2; Babcock et al., to be colonized by active burrowers. Levels of bioturbation at
2002).These conchostracans are known to have produced bilobed Mount Bowers are surprisingly high, with 21.6% of observations
trails (Tasch, 1964). recorded as BPBI = 5. This may reflect longer periods between
Bioturbation on bedding plane and vertical surfaces of the sedimentation events, as well as greater colonization by infau-
Mackellar Formation in the Shackleton and Beardmore Glacier nal animals. Mount Bowers may have been close to the original
areas has been assessed (Miller et al., 2002; Miller and Laban- shoreline (Fig. 2) and subject to enhanced deposition of organic

A B
Figure 7. (A, B) Horizontal bioturbation in interbedded sandstone and shale facies, Mount Bowers. The bioturbation does not penetrate more
than a few millimeters vertically. This degree of bioturbation is unusual for the Mackellar Formation. Ruler in mm.
200 Miller and Isbell

100

90

Mt. Bowers
80
Shackleton Gl
Mt. Weeks
70 Tillite Gl
Moore Mtns Figure 8. Bioturbation on bedding planes
at diverse outcrops of the Mackellar
% of observations

60
Formation within the study area. BPBI
(Bedding Plane Bioturbation Indicies,
50 Miller and Smail, 1997) are as follows:
category 1, no bioturbation; BPBI 2,
0%–10%; BPBI 3, 10%–40%; BPBI
40 4, 40%; BPBI 5, 60%–100%. Number
of observations: Mount Bowers, 1102;
Shackleton Glacier area, 1188; Mount
30
Weeks, 250; Tillite Glacier, 559; Moore
Mountains, 1220. Gl—Glacier.
20

10

0
1 2 3 4 5
BPBI

matter. If so, the additional nutritional supply may have stimu- tinctly different from those of marine shales (Berner and Raiswell,
lated colonization by infaunal animals, resulting in more obser- 1984; Miller and Collinson, 1994); (3) the absence of marine
vations of intensely disrupted sediments (BPBI = 5; Fig. 8). trace fossils, including penetrative forms typical of marine tur-
The distribution of bioturbation indicates that bioturbation, bidites (Fig. 6B); and (4) the absence of the Eurydesma fauna of
a proxy for infaunal activity, is least in areas of high sedimenta- marine body fossils that is characteristic of Permian high south-
tion, either near the shoreline or close to turbidity channels, and ern paleolatitude marine deposits (Banks and Clarke, 1987).
reaches highest levels farthest from loci of deposition toward the The Mackellar Formation and correlative units can be traced
distal part of the lake (e.g., Mount Butters and Ramsey Glacier in >1000 km through the Transantarctic Mountains through the
the Shackleton Glacier area, Fig. 2). Bioturbation is also enhanced Ohio Range to the Ellsworth Mountains (Fig. 1). The Polarstar
near the lake margins, in regions sheltered from sediment influx Formation in the Ellsworth Mountains, which is thicker than the
(e.g., Mount Butters and Mount Bowers) where organic matter Mackellar Formation and records sedimentation in a subsiding
may be less diluted by clastic sediment than elsewhere in the co-eval basin, contains marginal-marine trace fossils (e.g., Phy-
system, and where shallow depth and associated factors enhance codes) but lacks marine body fossils, including the cold-water
benthic diversity (Kalff, 2002). The generally low level of bio- bivalve Eurydesma (Collinson et al., 1992). The Discovery Ridge
turbation and its distribution mirrors that in modern lakes (White Formation of the Ohio Range is also correlative with the Mackel-
and Miller, 2008). lar and Polarstar Formations (Collinson et al., 1994). The salin-
ity conditions under which this formation was deposited are not
PALEOSALINITY clear, but marginal-marine trace fossils occur in the overlying
unit, suggesting a similar origin for the Discovery Ridge Forma-
The Mackellar Formation in the Beardmore and Shackleton tion (Bradshaw et al., 1984; Aitchison et al., 1988). Salinity con-
Glacier areas was deposited in fresh-water or slightly brackish- ditions inferred from these correlative units are consistent with
water conditions. This interpretation is based on three lines of a model of an inland sea from the (fresh-water) Mackellar For-
evidence: (1) the presence of low-diversity ichnofauna of sim- mation in the present-day Transantarctic Mountains to marginal-
ple crawling traces, including some forms (e.g., Cochlichnus, marine conditions in the present-day Ohio Range and Ellsworth
Fig. 7B) that are known to be fresh-water ichnogenera and occur Mountains. Although putative marine trace fossils indicate a
in lacustrine deposits of the (slightly) younger Buckley Forma- marine influence in the Ohio Range and Ellsworth Mountains,
tion; (2) high carbon/sulfur ratios of the shales, which are dis- the absence of the indisputably marine Eurydesma body fossil
Mackellar Formation, Transantarctic Mountains 201

assemblage suggests that salinity conditions were not normal The relationship between thermal alteration, as indicated
marine. The emerging scenario is one of an inland sea with a by vitrinite reflectance, and TOC in the Mackellar Formation, is
restricted opening to the paleo–Pacific Ocean in the Ellsworth not clear (Miller et al., 1988; Horner and Krissek, 1991, their
Mountains area, becoming increasingly fresh-water–dominated table 5). The mean Ro for 36 samples analyzed by Horner and
down the (present-day) Transantarctic Mountains with fully Krissek (1991, their table 5) is 2.73, and the mean TOC of 107
fresh-water conditions in the present-day Shackleton and Beard- samples is 0.322%. Samples from Mount Bowers (Fig. 2) have
more Glacier areas. a much higher TOC than samples from other areas; all of the
four Mount Bowers samples analyzed by Miller et al. (1988)
ORGANIC GEOCHEMISTRY have TOC >1.20%, and the mean TOC of the nine samples from
Mount Bowers analyzed by Horner and Krissek (1991, their table
Samples from the Mackellar Formation were analyzed for 5) is 1.31%. In contrast, the mean TOC of Horner and Krissek’s
total organic carbon (TOC), type of organic matter, and vitrin- (1991, their table 5) 98 samples from other localities is 0.23%.
ite reflectance by DGSI using standard protocols (Miller et al., The high TOC content at Mount Bowers is not due solely to
1988); the results of TOC and vitrinite reflectance analyses are reduced thermal alteration, as indicated by vitrinite reflectance.
given in Table 3. Visually determined kerogen types from 10 The mean for five samples from Mount Bowers is 2.78 Ro versus
samples indicate the presence of vitrinite (30%), inertinite (51%), a mean of 2.72 Ro for 31 samples from other localities (Horner
and amorphous organic matter (19%) (Miller et al., 1988); simi- and Krissek, 1991, their table 5); the sample with the highest
lar percentages of kerogen types (30% vitrinite, 50% inertinite, TOC (2.34%) has an Ro of 2.86.
20% amorphous) were found by Horner and Krissek (1991, their Thermal alteration does not affect all organic matter equally.
fig. 2). Approximately 10% of TOC may be lost during maturation if it is
Vitrinite is derived from terrestrial plant material (Hutton type III, but up to 80% may be lost if it is type I. If organic matter
et al., 1994). Inertinite is derived from organic matter that was at Mount Bowers was transported land plant material (type III)
burned soon after deposition, extensively oxidized by bacterial rather than algal material (type I), less would have been destroyed
degradation, or thermally altered. Distinguishing between these by heating, resulting in anomalously high TOC. Higher terrestrial
origins is possible, particularly if heating has not been too intense organic input into the Mount Bowers area than elsewhere in the
(e.g., Diessel, 1992; Mukhopadhyay, 1992), but this determina- MLIS is consistent with its inferred position near the shoreline;
tion was not undertaken for the Mackellar Formation samples. higher plant material is even carried into deep basins of African
In general, the Mackellar samples have little amorphous organic rift lakes (Talbot, 1988). Macroscopic plant debris should be
matter, suggesting deposition in an oxic or suboxic environment present if there were a large influx of plant material, but none
in which oxygen-consuming bacteria degraded organic matter. has been found in the Mackellar Formation, including at Mount
The widespread occurrence of biogenic structures implies oxy- Bowers or in the Moore Mountains, which are also interpreted as
genated bottom conditions. The few samples high in amorphous near the paleoshoreline (Fig. 2).
organic matter are low in TOC. This is consistent with deposi- Organic geochemical parameters of samples from the
tion under oxic conditions; under low-oxygen conditions, more Mackellar Formation do not provide unequivocal evidence about
organic matter would be preserved. the productivity of the MLIS. High productivity would be dem-
The Mackellar Formation is intruded by mafic Jurassic sills. onstrated by high TOC, by kerogen derived from algal sources,
Horner and Krissek (1991, their figs. 15 and 19) demonstrated and by evidence of anoxia, particularly in the profundal zone.
that vitrinite reflectance is raised and TOC lowered up to a dis- None of these indicators is present, but their absence does not
tance equivalent to the sill thickness from the contact with a sill. conclusively demonstrate that productivity in the MLIS was low.
This is consistent with studies suggesting that effects of intrusion Organic matter produced in the water column can be degraded
on the organic matter diminish significantly in short distances by aerobic bacteria. Comparison of algal species in the water
from the contact (Clayton and Bostick, 1986). At Mount Bow- column and sediments in modern lakes has verified that some
ers, vitrinite reflectance drops from 3.9 Ro to 1.71 Ro in <100 m phytoplankton are not preserved in sediments (Livingstone,
stratigraphically; this is interpreted to reflect increasing distance 1984); in Lake Michigan, only ~2% of the organic matter that
from a Jurassic sill of unknown thickness presumed to be beneath is fixed in the epilimnion is preserved in the sediment (Eadie
the exposure (Miller et al., 1988). et al., 1992). Bacterial degradation of organic matter would be
promoted by oxygen-rich water; the widespread distribution of
trace fossils in the Mackellar Formation suggests that the bot-
TABLE 3. TOTAL ORGANIC CARBON (TOC) AND tom waters of the MLIS were oxygen rich. Amorphous kerogen
VITRINITE REFLECTANCE (Ro) OF SAMPLES FROM THE could be derived from highly degraded algal organic matter, but
MACKELLAR FORMATION, BEARDMORE GLACIER AREA
it is not abundant and is restricted to samples with very low TOC.
Number of Mean Standard Variance
samples deviation Although there is no direct evidence regarding nutrient levels in
Ro 12 2.88 0.73 0.53 the MLIS, the geologic setting (i.e., its position overlying Perm-
TOC 20 0.61% 0.63 0.39 ian glacial deposits and Devonian quartz-rich sandstones), and
202 Miller and Isbell

absence of palynomorphs and plant macrofossils, suggest a low TABLE 4. CHARACTERISTICS OF MACKELLAR LAKE OR
influx of nutrients. INLAND SEA BASED ON LITHOLOGIC, ICHNOLOGIC,
SEDIMENTOLOGIC, AND ORGANIC GEOCHEMICAL DATA
Productivity in the lake might have been much higher than FROM THE MACKELLAR FORMATION
the low TOC would suggest if the organic matter had been altered Setting Cratonic, postglacial, ~80°S
or destroyed by bacterial degradation in well-oxygenated water Depositional processes Deposition dominated by turbidity
or sediments and/or by heating associated with intrusion of Juras- currents. Evidence for waves,
sic sills. Available data neither support nor rule out this scenario. storms restricted to rare symmetrical
The simplest interpretation of the organic geochemical data is ripples. No hummocky cross-
stratification
that (1) the MLIS received a limited influx of terrestrial organic
Salinity Fresh water
matter (more in the Mount Bowers area) from streams travers-
ing a recently deglaciated terrain (evidence is low TOC, vitrin- Total organic carbon in Low (mostly <0.5%)
sediments
ite, absence of plant fossils); (2) that the in situ productivity was
Extent Outcrop belt ~500 km × 100 km in
relatively low, perhaps reflecting low nutrient input; and (3) that area studied; correlative units
the organic matter was affected to varying degrees by bacterial extend an additional 1000 km
degradation and thermal alteration (evidence is high vitrinite Water depth >20 m–30 m; constrained by
reflectance) caused by Jurassic intrusions. thickness of upward-coarsening
sequences
LIMNOLOGY OF THE MACKELLAR LAKE OR
INLAND SEA (MLIS)

Limnologists study modern lakes as systems with complexly depth) data from 59 North American lakes located from 41° N to
interacting biological, physical, and chemical components con- 75° N was used to constrain the IFD of the MLIS (Shuter et al.,
strained and controlled by geologic setting and climate. Ancient 1983; Kalff, 2002). Climate modeling for the Permian indicates
lake systems increasingly are interpreted as reflecting the inter- mean summer temperatures at >60° S for both the Sakmarian
play between tectonics, climate, and the array of diverse interact- and Wordian (Early and Middle Permian, respectively) of about
ing biotic and abiotic factors (e.g., Cohen, 2003; Bohacs et al., –2 °C, and a mean winter temperature of –30 °C for a mean
2003; Gierlowski-Kordesch and Kelts, 2000). Important charac- annual temperature of –14 °C (Gibbs et al., 2002). Climate simu-
teristics of lakes required for understanding their functioning as lation for the southern polar region in the latest Permian (after
ecosystems include the (1) tectonic, geomorphic, and climatic significant warming had occurred) yields a mean summer tem-
setting; (2) size, depth, and morphology; (3) mixing regime (e.g., perature of 12 °C and a mean winter temperature of –20 °C for a
whether or not it is stratified, and, if so, the frequency and extent mean annual temperature of –8 °C (Kiehl and Shields, 2005). The
of mixing); (4) productivity; and (5) oxygen availability. The depth of the MLIS is known only to have exceeded the thickness
study of ancient lake deposits, such as the Mackellar Formation, of upward-coarsening sequences (20 m to 30 m). To ensure that
yields information about depositional processes, temporal and MLIS conditions were bracketed, numbers of IFD were calcu-
spatial distribution of the lake facies and of its fossilized flora lated for mean annual temperatures of –20 °C to 0 °C and mean
and fauna, organic matter preserved in the sediments, and oxygen lake depths of 20 m to 500 m (Fig. 9; only 12 modern lakes have
conditions during deposition. Known characteristics of the MLIS maximum depths >500 m; Kalff, 2002, table 4.5).
based on lithologic, sedimentologic, ichnologic, and organic geo- The major points gleaned from this estimate are that the
chemical data are summarized in Table 4. The challenge is to use MLIS was almost certainly ice covered a significant portion of
those data to constrain limnologic parameters in order to recon- the year, but also it was likely to have been ice free for at least two
struct the setting and functioning of the ancient lake ecosystem. months each year, long enough for photosynthesis to occur and
Ecologically significant characteristics discussed below include the surface water to warm. The most likely range of mean annual
the number of ice-free days; type of stratification and mixing; temperature based on the Permian climate models is –14 °C to
size, shape, and volume of the lake; and productivity. –8 °C. This corresponds to a minimum number of IFD of 80 and
a maximum of ~150, depending on lake depth; the deeper the
Ice-Free Days (IFD) lake, the more IFD (Fig. 9).

The number of days during the year that a lake is not covered Stratification and Mixing
by ice controls the temperature regime, light penetration, produc-
tivity, and stratification. At the continental scale, the number of Stratification and mixing in lakes exert strong control of
IFD on low altitude lakes is controlled primarily by mean air transport of materials (e.g., nutrients, oxygen) vertically within
temperature and secondarily by morphometry (fetch and mean the lake and thus are important in determining the structure of the
depth). A regression equation that predicts the number of IFD biotic communities and lake productivity (Kalff, 2002, p. 154).
based on mean air temperature and mean depth (a proxy for lake Light penetrating the upper layers of the lake during the summer
Mackellar Formation, Transantarctic Mountains 203

250

20 m
50 m
100 m
200 200 m
500 m
Figure 9. Number of ice-free days (IFD)
for the MLIS, calculated using different
mean annual temperatures and various
Ice-free Days/Year

150 lake depths that bracket the depth of the


MLIS and temperature for Permian high
southern latitude (Gibbs et al., 2002;
Kiehl and Shields, 2005). Most likely
the temperature was –8 °C to –14 °C;
100 the most likely depth was 100 m to
300 m. Estimates are based on the fol-
lowing equation derived from data from
75 modern Northern Hemisphere lakes
(Shuter et al., 1983; Kalff, 2002, p. 157):
50 lnIFD = 0.06*(TEMP) + 0.073*(lnzmean)
+ 5.005.

0
-20 -18 -16 -14 -12 -10 -8 -6 -4 -2 0
Mean annual T

both promotes photosynthesis by planktic organisms and warms unknown. What is known for the MLIS is the area of outcrop;
the water, thereby reducing density; postmortem decay of organ- the region considered here is restricted to the Nimrod, Beard-
isms sinking to the bottom depletes oxygen that is not replen- more, and Shackleton Glacier areas and does not include areas of
ished if the water is thermally stratified, leading to low-oxygen outcrop of correlative units (Fig. 2). If rocks exposed at Mount
conditions in the water column and sediments. Bowers (Fig. 2) were deposited in lake-marginal environments,
Was the MLIS stratified or mixed? Several lines of evidence the dimensions of the outcrop of the Mackellar Formation are
based on characteristics of modern lakes suggest that it was sea- ~500 km by 200 km. Used as the maximum effective length and
sonally stratified. The stability of stratification decreases as the maximum effective width, these numbers correspond to a mixing
temperature differential between the water at the surface and at depth of 48 m (Table 5); even if the effective maximum length
depth decreases, a difference that tends to decrease with increas- and width were much larger than that indicated by the outcrop
ing latitude (Geller, 1992). A simple measure of the likelihood area, the mixing depth still would be <65 m (Table 5). However,
of stratification of temperature in a lake is the ratio of the thick- because the age of the Mackellar Formation at different outcrops
ness of the epilimnion, the upper zone of the water column where is poorly constrained, it is possible that the Mackellar facies was
photosynthesis occurs, to the maximum depth of the lake. If this deposited in separate small lakes extant at different times in dif-
ratio is <0.5 the lake undergoes stable seasonal (summer) strati- ferent places along the outcrop belt. The paucity of shoreline
fication. If it is >2.0 there is no stratification; if it is between 0.5 deposits (e.g., wave-influenced versus turbidity-current deposits)
and 2.0 there may be stratification overturned by strong winds or points to one (or a few) large lakes, but the possibility of a series
intermittent stratification (Kalff, 2002, p. 165). of small lakes cannot be ruled out.
The thickness of the epilimnion corresponds to the thickness The depth to the thermocline based on areas of an MLIS with
of the mixed layer and depth to the top of the thermocline. Mixing the dimensions given in Table 5 is shown in Figure 10. For a lake
depth and depth to the thermocline were calculated for the Perm- 500 km by 100 km (100,000 km2) the depth to the thermocline
ian MLIS, using models presented by Kalff (2002, table 11.2) is ~20 m. The major point underscored by Table 5 and Figure 10
based on data from lakes in Poland and Canada (mixing depth) is that depth to thermocline and mixing depth are remarkably
and lakes worldwide (depth to thermocline); Kalff presented no unchanged for large variations in mean effective length and width
models based on modern polar lakes. Mixing depth is dependent and in area. Mixing depth is reduced by half (from ~60 m to
on the maximum effective length and maximum effective width ~30 m when mean length is reduced by 90% and mean width by
of the lake, neither of which is known for the MLIS; the depth 85%) (Table 5); similarly, depth to thermocline varies little until
to the top of the thermocline is based on lake area, which is also lake size becomes very small (<100 km2; Fig. 10).
204 Miller and Isbell

TABLE 5. MIXING DEPTH (zmix) OF LAKE SURFACE WATER AS A Using depth to thermocline or mixing depth as a proxy for
FUNCTION OF MEAN EFFECTIVE LENGTH (MEL) thickness of the epilimnion (ze) and maximum depths (zm) rang-
AND MEAN EFFECTIVE WIDTH (MEW)
ing from 20 m to 500 m, the ze/zm for MLIS consistently is <0.5
Length (MEL) Width (MEW) Mixing depth (zmix)
(km) (km) (m)
for all realistic values. This strongly suggests stable seasonal
1000 200 63 stratification (Kalff, 2002, p. 165). The MLIS is reconstructed as
750 200 58 most likely a dimictic lake that was covered with ice part of the
750 150 56 year and thermally stratified during the summer, with two periods
500 200 51 of mixing occurring after ice melt and after fall cooling of surface
500 150 49
water following warming during the summer. The only uncertain-
500 100 48
400 100 44 ties that arise regarding this interpretation are the applicability of
300 100 40 models that are not based on modern polar lakes and whether or
200 100 36 not in the MLIS there was a sufficiently large temperature differ-
200 50 33 ential between surface and deep water for stratification to occur.
100 50 27
Thus, the possibility that the MLIS was a cold monomictic lake
50 50 23
50 20 20 cannot be completely excluded, although it is less likely because
10 2 10 of the large number of IFD days and the robustness of the model
5 1 7 results that point to a dimictic lake.
Note: Based on following equation derived from characteristics of
88 extant lakes in Poland and Canada; results show mixing depth Productivity
remains between 40 m and 60 m for a wide range of lake dimensions:
0.41
zmix = 4.6*([MEL + MEW]/2) .
TOC in lake sediments commonly is interpreted to reflect
the amount of biomass that sinks to the bottom, and the lake’s
productivity. In oligotrophic (low productivity) systems, TOC in

25

20
Thermocline Depth, m

15

10

0
0 20000 40000 60000 80000 100000 120000 140000 160000 180000 200000
Lake area (sq km)

Figure 10. Depth to thermocline (zt) in modern lakes modeled as a function of lake area (A): zt = ln([A/0.5]/0.043)2.35. Note
that depth to themocline varies relatively little until lake size becomes very small (e.g., <100 km2; Kalff, 2002).
Mackellar Formation, Transantarctic Mountains 205

sediments provides a reliable qualitative record of productivity is constrained only by comparison with modern glacial lakes,
under a variety of mixing regimes (Cohen, 2003, p. 253). How- most of which are <500 m deep. Pleistocene glacial Lake Agas-
ever, in areas of high productivity, patterns of mixing, bacterial siz, the largest proglacial lake known with a surface area of
degradation of organic matter in the water column, and dilution 350,009 km2, had a maximum depth of 200 m (Kalff, 2002). The
of organic matter by clastic sediments are factors controlling MLIS resembles glacial Lake Agassiz in its cratonic setting and
the amount of TOC preserved in the sediment. The absence of in the fact that it is underlain by glacial deposits. Recent evalua-
macroscopic plant fossils in the Mackellar Formation implies tion of the thickness and extent of the late Paleozic glaciation in
that nutrient input from land plants was low (probably too low Antarctica indicates coverage by a single advance of a smaller
to support high productivity), and the low level of bioturbation and thinner ice sheet than that previously envisioned (Isbell et al.,
suggests that biologic mixing in the sediment did not enhance 2003). This ice sheet was unlikely to have scoured to a depth of
bacterial degradation enough to significantly reduce TOC in the 500 m; a maximum lake depth of 200 m is more plausible.
sediment. The low TOC, the postglacial setting, the widespread The relative lack of MLIS deposits indicative of storm or wave
distribution of trace fossils (if low abundance), and the absence activity is perplexing. The fetch of a large lake would be sufficient
of well-preserved allochthonous or autochthonous organic matter to produce long-length waves that would rework bottom sedi-
are all consistent with the MLIS as an oligotrophic, low produc- ments to a significant depth, but hummocky cross-stratification
tivity lake. is not seen, and symmetrical ripples occur at only one locality
(Mount Weeks, Fig. 2). The slope of the bottom may have been
CONCLUSIONS: OVERVIEW OF THE MLIS AND sufficiently steep that only a small area was above storm wave
ITS FUNCTIONING base, but this is unlikely, given the cratonic setting of the MLIS.
Alternatively, lake margin deposits may have been entrained in
The MLIS was situated on a recently deglaciated, cratonic, turbidity currents. It is also possible that major storms occurred
presumably low-relief terrain at southern polar paleolatitude with when the MLIS was covered by ice; this would have been most
a cold but not frigid climate. Braided streams traversed an out- likely if the number of IFD was near the upper end of the range
wash plain before entering the MLIS, loaded with suspended sed- suggested by models of the Permian climate (Fig. 9).
iment, and formed turbidity currents that transported fine sand, Although the morphology of the MLIS remains poorly
silt, and mud away from the lake margin. The MLIS was covered known, integration of sedimentologic, ichnologic, and organic
with ice from a minimum of two months per year to a maximum geochemical data with information derived from studies of mod-
of six months per year. It was stratified in the summer but was ern lakes yields a remarkably clear, if incomplete, picture of the
mixed twice yearly; mixing was enhanced by turbidity currents. functioning of this Permian postglacial lake. Because the record
Bottom waters remained oxic, as indicated by widespread if not of the MLIS is better preserved by the extensively exposed Mack-
pervasive bioturbation. The distribution of bioturbation, a proxy ellar Formation than are other similar, co-eval lakes elsewhere in
for infauna, indicates that animal activity was least present in Gondwana, the reconstruction of the MLIS provides a model of
areas of rapid sedimentation near the lake margin or adjacent to an ancient high latitude lake system that may be applied fruitfully
crevasse splays at turbidite-system channels and was greatest in to other less well-exposed Paleozoic postglacial lake deposits.
shallow littoral zones away from loci of deposition and in the
distal portions of the lake. Some organic matter was delivered to ACKNOWLEDGMENTS
the lake from the watershed, but the absence of plant macrofos-
sils indicates that vegetation probably was not lush in the recently We were supported by U.S. National Science Foundation
deglaciated area. Low TOC in the Mackelllar Formation prob- grant ANT 0126146 and OPP 0440954 to Miller and NSF grant
ably accurately reflects a limited contribution of allochthonous ANT 0126086 to Isbell. We thank Elizabeth Gierlowski-Kordesch
organic matter and low productivity within the lake, although and Kevin Bohacs for helpful comments on an earlier version of
productivity in the surface waters might have been higher than this manuscript.
suggested by the very low TOC values.
The size and depth of the MLIS are poorly constrained. If REFERENCES CITED
the size is approximated by the extent of the Mackellar Forma-
tion outcrop, the MLIS covered ~50,000 km2 in the present-day Aitchison, J.C., Bradshaw, M.A., and Newman, J., 1988, Lithofacies and origin
of the Buckeye Formation: Late Paleozoic glacial and glaciomarine sedi-
Nimrod, Shackleton, and Beardmore Glacier areas. Alternatively, ments, Ohio Range, Transantarctic Mountains, Antarctica: Palaeogeog-
as discussed above, the area may have been covered by a series raphy, Palaeoclimatology, Palaeoecology, v. 64, p. 93–104, doi: 10.1016/
of small time-transgressive lakes. A major factor favoring a large 0031-0182(88)90145-9.
Archanjo, C.J., Silva, M.G., Castro, J.C., Launeau, P., Trindale, R.I.F., and
lake interpretation is the relative lack of shallow-water depos- Macedo, J.W.P., 2006, AMS and grain shape fabric of the late Paleozoic
its that would constitute the dominant facies if there had been diamictites of the southeastern Parana Basin, Brazil: Journal of the Geo-
numerous small lakes. logical Society [London], v. 163, p. 95–106, doi: 10.1144/0016-764904
-155.
The depth of the MLIS was greater than the ~20 m to 30 m Babcock, L.E., Isbell, J.L., Miller, M.F., and Hasiotis, S.T., 2002, New late
thickness of upward-coarsening sequences. The maximum depth Paleozoic conchostracan (Crustacea, Branchiopoda) from the Shackleton
206 Miller and Isbell

Glacier area, Antarctica: Age and paleoenvironmental implications: Jour- Gierlowski-Kordesch, E.H., and Kelts, K.R., 2000, Lake Basins through Space
nal of Paleontology, v. 76, p. 70–75, doi: 10.1666/0022-3360(2002)076 and Time: American Association of Petroleum Geologists Studies in
<0070:NLPCCB>2.0.CO;2. Geology 46, 648 p.
Banks, M.R., and Clarke, M.J., 1987, Changes in the geometry of the Tasma- Horner, T.C., and Krissek, L.A., 1991, Contributions of sedimentologic, ther-
nian basin in the late Paleozoic, in McKenzie, G.D., ed., Gondwana Six: mal alteration, and organic carbon data to paleoenvironmental interpreta-
Stratigraphy, Sedimentology, and Paleontology: American Geophysical tion of fine-grained Permian clastics from the Beardmore Glacier region,
Union, Geophysical Monograph 41, p. 1–14. Antarctica: Washington, D.C., American Geophysical Union, Antarctic
Barrett, P.J., Elliot, D.H., and Lindsay, J.F., 1986, The Beacon Supergroup Research Series, v. 53, p. 33–65.
(Devonian-Triassic) and Ferrar Group (Jurassic) in the Beardmore Glacier Hutton, A., Bharati, S., and Robl, T., 1994, Chemical and petrographic classifica-
area, Antarctica, in Turner, M.D., and Splettstoesser, J.F., eds., Geology of tion of kerogen/macerals: Energy & Fuels, v. 8, p. 1478–1488, doi: 10.1021/
the Central Transantarctic Mountains: Washington, D.C., American Geo- ef00048a038.
physical Union, Antarctic Research Series, v. 36, p. 339–428. Isbell, J.L., 1999, The Kukri erosion surface: A reassessment of its relation-
Berner, R.A., and Raiswell, R., 1984, C/S method for distinguishing freshwater ship to rocks of the Beacon Supergroup in the central Transantarctic
from marine sedimentary rocks: Geology, v. 12, p. 365–368, doi: 10.1130/ Mountains, Antarctica: Antarctic Science, v. 11, p. 228–238, doi: 10.1017/
0091-7613(1984)12<365:CMFDFF>2.0.CO;2. S0954102099000292.
Bohacs, K.M., Carroll, A.R., and Neal, J.E., 2003, Lessons from large lake Isbell, J.L., and Collinson, J.W., 1988, Fluvial architecture of the Fairchild and
systems: Thresholds, nonlinearity, and strange attractors, in Chan, M.A., Buckley Formations (Permian), Beardmore Glacier area: Antarctic Jour-
and Archer, A.W., eds., Extreme Depositional Environments: Mega End nal of the United States, v. 23, no. 5, p. 3–5.
Members in Geologic Time: Geological Society of America Special Paper Isbell, J.L., Miller, M.F., Wolfe, K.L., and Lenaker, P.A., 2003, Timing of late
370, p. 75–90. Paleozoic glaciation in Gondwana: Was glaciation responsible for the
Bradshaw, M.A., Newman, J., and Aitchison, J.C., 1984, Preliminary geologi- development of Northern Hemisphere cyclothems?, in Chan, M.A., and
cal results of the 1983–1984 Ohio Range Expedition, New Zealand: Ant- Archer, A.A., eds., Extreme Depositional Environments: Mega End Mem-
arctic Record (Tokyo), v. 5, p. 1–17. bers in Geologic Time: Geological Society of America Special Paper 370,
Buatois, L.A., and Mangano, M.G., 1995, The paleoenvironmental and paleo- p. 5–24.
ecological significance of the lacustrine Mermia ichnofacies: An arche- Jones, A.T., and Fielding, C.R., 2004, Sedimentological record of the late Paleo-
typical subaqueous nonmarine trace fossil assemblage: Ichnos, v. 4, zoic glaciation in Queensland, Australia: Geology, v. 32, p. 153–156, doi:
p. 151–161, doi: 10.1080/10420949509380122. 10.1130/G20112.1.
Clayton, J.L., and Bostick, N.H., 1986, Temperature effects on kerogen and Kalff, J., 2002, Limnology: Inland Water Ecosystems: Upper Saddle River,
on molecular and isotopic composition of organics in Pierre Shale near New Jersey, Prentice Hall, 592 p.
an igneous dike: Organic Geochemistry, v. 10, p. 135–143, doi: 10.1016/ Kiehl, J.F., and Shields, C.A., 2005, Climate simulation of the latest Perm-
0146-6380(86)90017-3. ian: Implications for mass extinction: Geology, v. 33, p. 757–760, doi: 10
Cohen, A.S., 2003, Paleolimnology: The History and Evolution of Lake Sys- .1130/G21654.1.
tems: New York, Oxford University Press, 500 p.
Knepprath, N.E., Miller, M.F., and Isbell, J.L., 2004, Dense Permian forest with
Collinson, J.W., Vavra, C.I., and Zawiskie, J.M., 1992, Sedimentology of
large trees: Upper Buckley Formation, central Transantarctic Mountains:
the Polarstar Formation, Permian, Ellsworth Mountains, Antarctica, in
Geological Society of America Abstracts with Programs, v. 36, no. 5,
Webers, G.F., Craddock, C., and Splettstoesser, J.F., eds., Geology of the
p. 92.
Ellsworth Mountains, Antarctica: Geological Society of America Memoir
Limarino, C.O., Césari, S.N., Net, L.I., Marenssi, A., Guitierrez, R.P., and
170, p. 63–79.
Tripaldi, A., 2002, The Upper Carboniferous postglacial transgression
Collinson, J.W., Isbell, J.L., Elliot, D.H., Miller, M.F., and Miller, J.M.G.,
in the Paganzo and Río Blanco basins (northwestern Argentina): Facies
1994, Permian–Triassic Transantarctic Basin, in Veevers, J.J., and Powell,
and stratigraphic significance: Journal of South American Earth Sciences,
C.McA., eds., Permian–Triassic Pangean Basins and Foldbelts along the
Panthalassan Margin of Gondwanaland: Geological Society of America v. 15, p. 445–460, doi: 10.1016/S0895-9811(02)00048-2.
Memoir 184, p. 173–222. Lindsay, J.F., 1997, Permian postglacial environments of the Australian plate,
Dickins, J.M., 1996, Problems of a late Paleozoic glaciation in Australia and in Martini, I.P., ed., Late Glacial and Postglacial Environmental Changes:
subsequent climate in the Permian: Palaeogeography, Palaeoclimatology, Oxford, UK, Oxford University Press, p. 5–9.
Palaeoecology, v. 125, p. 185–197, doi: 10.1016/S0031-0182(96)00030-2. Livingstone, D., 1984, The preservation of algal remains in recent lake sedi-
Diessel, C.F.K., 1992, Coal-Bearing Depositional Systems: Berlin, Springer- ments, in Haworth, E.Y., and Lund, J.W.G., eds., Lake Sediments and
Verlag, 721 p. Environmental History: Minneapolis, University of Minnesota Press,
Droser, M.L., and Bottjer, D.J., 1986, A semiquantitative field classification of p. 191–202.
ichnofabric: Journal of Sedimentary Petrology, v. 56, p. 558–559. López-Gamundí, O.R., 1997, Glacial-postglacial transition in the late Paleo-
Eadie, J.M., Bell, G.L., Robbins, J.A., and Myers, P.A., 1992, Carbon flux and zoic basins of southern South America, in Martini, I.P., ed., Late Glacial
remineralization in Lake Michigan: Eos (Transactions, American Geo- and Postglacial Environmental Changes: Oxford, UK, Oxford University
physical Union), v. 73 (Fall 1992 Supplement), p. 197. Press, p. 147–168.
Elliot, D.H., 1975, Gondwana basins in Antarctica, in Campbell, K.S.W., ed., Lowe, D.R., 1982, Sediment gravity flows II. Depositional models with spe-
Gondwana Geology: Canberra, Australian National University Press, cial reference to the deposits of high density turbidity currents: Journal of
p. 493–536. Sedimentary Petrology, v. 52, p. 179–197.
Eyles, C.H., Mory, A.J., and Eyles, N., 2003, Carboniferous-Permian facies Maejima, W., Das, R., Pandya, K.L., and Hayashi, M., 2004, Deglacial control
and tectono-stratigraphic successions of the glacially influenced and on sedimentation and basin evolution of Permo-Carboniferous Talchir
rifted Carnarvon Basin, western Australia: Sedimentary Geology, v. 155, Formation, Talchir Gondwana Basin, Orissa, India: Gondwana Research,
p. 63–86, doi: 10.1016/S0037-0738(02)00160-4. v. 7, p. 339–352.
Fenton, M.M., Moran, S.R., Teller, J.T., and Clayton, L., 1983, Quaternary Miller, J.M.G., 1989, Glacial advance and retreat sequences in a Permo-
stratigraphy and history in the southern part of the Lake Agassiz Basin, Carboniferous section, central Transantarctic Mountains: Sedimentology,
in Teller, J.T., and Clayton, L., eds., Glacial Lake Agassiz: Geological v. 36, p. 419–430, doi: 10.1111/j.1365-3091.1989.tb00617.x.
Association of Canada Special Paper 26, p. 49–74. Miller, M.F., and Collinson, J.W., 1994, Late Paleozoic inland sea filled by fine-
Geller, W., 1992, The temperature stratification and related characteristics of grained turbidites: Mackellar Formation, central Transantarctic Moun-
Chilean lakes in midsummer: Aquatic Sciences, v. 54, p. 37–57, doi: tains, in Deynoux, M., et al., eds., Earth’s Glacial Record: Cambridge,
10.1007/BF00877263. UK, Cambridge University Press, p. 215–233.
Gibbs, M.T., Rees, P.M., Kutzbach, J.E., Ziegler, A.M., Behling, P.J., and Rowley, Miller, M.F., and Labandeira, C.C., 2002, Slow crawl across the salinity
P.B., 2002, Simulations of Permian climate and comparison with climate- divide: Delayed colonization of freshwater ecosystems by invertebrates:
sensitive sediments: Journal of Geology, v. 110, p. 33–55, doi: 10.1086/ GSA Today, v. 12, no. 12, p. 4–10, doi: 10.1130/1052-5173(2002)012
324204. <0004:SCATSD>2.0.CO;2.
Mackellar Formation, Transantarctic Mountains 207

Miller, M.F., and Smail, S.E., 1997, A semiquantitative field method for evalua- Talbot, M.R., 1988, The origins of oil source rocks: Evidence from the lakes of
tion of bioturbation on bedding planes: Palaios, v. 12, p. 391–396, doi: 10 tropical Africa, in Fleet, A.J., Kelts, K., and Talbot, M.R., eds., Lacustrine
.2307/3515338. Petroleum Source Rocks: Geological Society [London] Special Publica-
Miller, M.F., Frisch, R.S., Collinson, J.W., and Dow, W.G., 1988, Permian black tion 40, p. 29–43.
shales of the central Transantarctic Mountains: Proceedings of the 1987 Tasch, P., 1964, Conchostracan trails in bottom clay muds and on turbid water
Eastern Oil Shale Symposium, Kentucky Energy Cabinet Laboratory, surfaces: Transactions of the Kansas Academy of Science, v. 67, p. 126–
p. 193–200. 128, doi: 10.2307/3626685.
Miller, M.F., McDowell, T.A., Smail, S.E., Shyr, Y., and Kemp, N.R., 2002, Teller, J.T., 2001, Formation of large beaches in an area of rapid rebound: The
Hardly used habitats: Dearth and distribution of burrowing in Paleozoic three-outlet control of Lake Agassiz: Quaternary Science Reviews, v. 20,
and Mesozoic stream and lake deposits: Geology, v. 30, p. 527–530, doi: p. 1649–1659, doi: 10.1016/S0277-3791(01)00007-5.
10.1130/0091-7613(2002)030<0527:HUHDAD>2.0.CO;2. Teller, J.T., and Bluemle, J.P., 1983, Geological setting of the Lake Agassiz
Miller, M.F., McDowell, T.A., Berrios, L.A., and Shyr, Y., 2005, Bioturbation region, in Teller, J.T., and Clayton, L., eds., Glacial Lake Agassiz: Geo-
as a proxy for infaunal animal activity in Permian-Jurassic freshwater logical Association of Canada Special Paper 26, p. 7–20.
deposits: Geological Society of America Abstracts with Programs, v. 37, Teller, J.T., and Clayton, L., 1983, Glacial Lake Agassiz: Geological Associa-
no. 7, p. 340. tion of Canada Special Paper 26, 451 p.
Mukhopadhyay, P.K., 1992, Maturation of organic matter as revealed by micro- Teller, J.T., Boyd, M., Yang, Z., Kor, P.S.G., and Fard, A.M., 2005, Alternative
scopic methods: Applications and limitations of vitrinite reflectance routing of Lake Agassiz overflow during the Younger Dryas: New dates,
and continuous spectral and pulsed later fluorescence spectroscopy, in paleotopography, and a re-evaluation: Quaternary Science Reviews, v. 24,
Wolf, K.H., and Chilingarian, G.V., eds., Diagenesis III. Developments p. 1890–1905, doi: 10.1016/j.quascirev.2005.01.008.
in Sedimentology: Amsterdam, Elsevier Science Publication Co., v. 47, Trosdtorf, I., Jr., Rocha-Campos, A.C., dos Santos, P.R., and Tomio, A., 2005,
p. 435–510. Origin of late Paleozoic, multiple, glacially striated surfaces in northern
Powell, C.McA., and Li, Z.X., 1994, Reconstruction of the Panthalassan mar- Paraná Basin (Brazil): Some implications for the dynamics of the Paraná
gin of Gondwana, in Veevers, J.J., and Powell, C.McA., eds., Permian– glacial lobe: Sedimentary Geology, v. 181, p. 59–81, doi: 10.1016/j.sedgeo
Triassic Pangean Basins and Foldbelts along the Panthalassan Margin of .2005.07.006.
Gondwanaland: Geological Society of America Memoir 184, p. 5–9. Visser, J.N.J., 1994, A Permian argillaceous syn- to post-glacial foreland
Powell, M.G., 2005, Climate basis for sluggish macroevolution during the late sequence in the Karoo Basin, South Africa, in Deynoux, M., et al., eds.,
Paleozoic ice age: Geology, v. 33, p. 381–384, doi: 10.1130/G21155.1. Earth’s Glacial Record: Cambridge, UK, Cambridge University Press,
Rees, P.M., Ziegler, A.M., Gibbs, M.T., Kutzbach, J.E., Behling, P.J., and Row- p. 193–203.
ley, D.B., 2002, Permian phytogeographic patterns and climate data/model White, D.S., and Miller, M.F., 2008, Benthic invertebrate activity in lakes:
comparisons: Journal of Geology, v. 110, p. 1–31, doi: 10.1086/324203. Linking present and historical bioturbation patterns: Aquatic Biology,
Scheffler, K., Hoernes, S., and Schwark, L., 2003, Global changes during Car- v. 2, p. 269–277.
boniferous–Permian glaciation of Gondwana: Linking polar and equato-
rial climate evolution by geochemical proxies: Geology, v. 31, p. 605–608,
doi: 10.1130/0091-7613(2003)031<0605:GCDCGO>2.0.CO;2.
Shuter, B.J., Schlesinger, D.A., and Zimmerman, A.P., 1983, Empirical predic-
tors of annual surface water temperature cycles in North American lakes:
Canadian Journal of Fisheries and Aquatic Sciences, v. 40, p. 1838–1845. MANUSCRIPT ACCEPTED BY THE SOCIETY 21 DECEMBER 2009

Printed in the USA

You might also like